Skip to main content

Ab initio Calculations

  • Chapter
Computational Chemistry

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 74.99
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Preview

Unable to display preview. Download preview PDF.

Unable to display preview. Download preview PDF.

References

  1. General discussions of and references to ab initio calculations are found in: (a) I. N. Levine, “Quantum Chemistry,” 4th edn, Prentice Hall, Engelwood Cliffs, New Jersey, 2000. (b) J. P. Lowe, “Quantum Chemistry,” 2nd edn, Academic Press, New York, 1993. (c) F. L. Pilar, “Elementary Quantum Chemistry,” 2nd edn, McGraw-Hill, New York, 1990. (d) An advanced book: A. Szabo and N. S. Ostlund, “Modern Quantum Chemistry,” McGraw-Hill, New York, 1989. (e) J. B. Foresman and Æ. Frisch, “Exploring Chemistry with Electronic Structure Methods,” Gaussian Inc., Pittsburgh, PA, 1996. (f) A. R. Leach, “Molecular Modelling,” Longman, Essex, England, 1996. (g) An important reference is still: W. J. Hehre, L. Radom, P. v. R. Schleyer, and J. A. Pople, “Ab Initio Molecular Orbital Theory,” Wiley, New York, 1986. (h) A recent evaluation of the state and future of quantum chemical calculations, with the emphasis on ab initio methods: M. Head-Gordon, J. Phys. Chem., 1996, 100, 13213. (i) F. Jensen, “Introduction to Computational Chemistry,” Wiley, New York, 1999. (j) M. J. S. Dewar, “The Molecular Orbital Theory of Organic Chemistry,” McGraw-Hill, New York, 1969. This book contains many trenchant comments by one of the major contributors to computational chemistry; begins with basic quantum mechanics and ab initio theory, although it later stresses semiempirical theory, (k) D. Young, “Computational Chemistry. A Practical Guide for Applying Techniques to Real World Problems,” Wiley, New York, 2001. (1) C. J. Cramer, “Essentials of Computational Chemistry,” Wiley, New York, 2002.

    Google Scholar 

  2. D. R. Hartree, Proc. Cambridge Phil. Soc., 1928, 24, 89.

    CAS  Google Scholar 

  3. The relativistic Schrödinger equation is called the Dirac equation; see [1a], pp. 602–604, (b) For a brief discussion of spin-orbit interaction see [1a], loc. cit.

    Google Scholar 

  4. See [1b], pp. 127–132.

    Google Scholar 

  5. J. C. Slater, Phys. Rev., 1930, 35, 210. (b) V. Fock, Z. Physik, 1930, 61, 126.

    Google Scholar 

  6. Reference [1a], pp. 187–189 and 282–285.

    Google Scholar 

  7. Although it is sometimes convenient to speak of electrons as belonging to a particular atomic or molecular orbital, and although they sometimes behave as if they were localized, no electron is really confined to a single orbital, and in a sense all the electrons in a molecule are delocalized; see [1j], pp. 139–143.

    Google Scholar 

  8. See, e.g. [1c], p. 200.

    Google Scholar 

  9. J. A. Pople and D. L. Beveridge, “Approximate Molecular Orbital Theory,” McGraw-Hill, New York, 1970, chapters 1 and 2.

    Google Scholar 

  10. Reference [1b], Appendix 7.

    Google Scholar 

  11. Reference [1a], p. 284.

    Google Scholar 

  12. Reference [1j], chapter 2.

    Google Scholar 

  13. Reference [1a], p. 474.

    Google Scholar 

  14. Reference [1a], chapter 8.

    Google Scholar 

  15. See, e.g. C. L. Perrin, “Mathematics for Chemists,” Wiley-Interscience, New York, 1970, pp. 39–41.

    Google Scholar 

  16. Reference [1b], pp. 354–355.

    Google Scholar 

  17. How do we know that iterations improve psi and epsilon? This is not always the case, see e.g. [1j], p. 35, but in practice “initial guess” solutions to the Hartree-Fock equations usually converge fairly smoothly to give the best wavefunction and orbital energies (and thus total energy) that can be obtained by the HF method from the particular kind of guess wavefunction (e.g. basis set; section 5.2.3.6e).

    Google Scholar 

  18. Reference [1a], pp. 305–315.

    Google Scholar 

  19. C. C. J. Roothaan, Rev. Mod. Phys., 1951, 23, 69; G. G. Hall, Proc. Roy. Soc. (London), 1951, A205, 541.

    Article  CAS  Google Scholar 

  20. Reference [1c], pp. 288–299.

    Google Scholar 

  21. Frequencies and zero-point energies are discussed in L. Radom, P. v. R. Schleyer, and J. A. Pople, “Ab Initio Molecular Orbital Theory,” Wiley, New York, 1986 [1g], section 6.3.

    Google Scholar 

  22. GAUSSIAN 92, Revision F.4: M. J. Frisch, G. W. Trucks, M. Head-Gordon, P. M. W. Gill, M. W. Wong, J. B. Foresman, B. G. Johnson, H. B. Schlegel, M. A. Robb, E. S. Repogle, R. Gomperts, J. L. Andres, K. Raghavachari, J. S. Binkley, C. Gonzales, R. L. Martin, D. J. Fox, D. J. Defrees, J. Baker, J. J. P. Stewart, and J. A. Pople; Gaussian, Inc., Pittsburgh, PA, 1992.

    Google Scholar 

  23. See, e.g. “Interactive Linear Algebra: A Laboratory Course Using Mathcad,” G. J. Porter and D. R. Hill, Springer Verlag, New York, 1996.

    Google Scholar 

  24. Cf. N. S. Ostlund, “Modern Quantum Chemistry,“ McGraw-Hill, New York, 1989 [1d], pp. 152–171.

    Google Scholar 

  25. Reference N. S. Ostlund, “Modern Quantum Chemistry,“ McGraw-Hill, New York, 1989 [1d], Appendix A.

    Google Scholar 

  26. Reference [1a], pp. 494–498.

    Google Scholar 

  27. See [1a-i].

    Google Scholar 

  28. S. F. Boys, Proc. Roy. Soc. (London), 1950, A200, 542.

    CAS  Google Scholar 

  29. Spartan is an integrated molecularmechanics, ab initio and semiempirical program with an outstanding input/output graphical interface that is available in UNIX workstation and PC versions: Wavefunction Inc., 18401 Von Karman, Suite 370, Irvine CA 92715; http://www.wavefun.com..

  30. Reference Æ Frisch, “Exploring Chemistry with Electronic Structure Methods,” Gaussian Inc., Pittsburgh, PA, 1996. [1e], pp. 32–33.

    Google Scholar 

  31. W. J. Hehre, “Practical Strategies for Electronic Structure Calculations,“ Wavefunction, Inc., Irvine, CA, 1995.

    Google Scholar 

  32. Reference L. Radom, P. v. R. Schleyer, and J. A. Pople, “Ab Initio Molecular Orbital Theory,” Wiley, New York, 1986. [1g], pp. 65–88.

    Google Scholar 

  33. J. Simons and J. Nichols, “Quantum Mechanics in Chemistry,“ Oxford University Press, New York, 1997, pp. 412–417.

    Google Scholar 

  34. Reference [1a], pp. 490–493.

    Google Scholar 

  35. M. J. S. Dewar and D. M. Storch, J. Am. Chem. Soc., 1985, 107, 3898.

    CAS  Google Scholar 

  36. S. Inagaki, Y. Ishitani, and T. Kakefu, J. Am. Chem. Soc., 1994, 116, 5954.

    Article  CAS  Google Scholar 

  37. E. Lewars, J. Mol. Struct. (Theochem), 1998, 423, 173.

    Article  CAS  Google Scholar 

  38. The experimental geometries of Me2SO and NSF are taken from L. Radom, P. v. R. Schleyer, and J. A. Pople, “Ab Initio Molecular Orbital Theory,” Wiley, New York, 1986. [1g], Table 6.14.

    Google Scholar 

  39. O. Wiest, D. C. Montiel, and K. N. Houk, J. Phys. Chem. A, 1997, 101, 8378, and references therein.

    Article  CAS  Google Scholar 

  40. C. Van Alsenoy, C.-H. Yu, A. Peeters, J. M. L. Martin, and L. Schüfer, J. Phys. Chem. A, 1998, 102, 2246.

    Google Scholar 

  41. Basis sets without polarization functions evidently make lone-pair atoms like tricoordinate N and tricoordinate O+ too flat: C. C. Pye, J. D. Xidos, R. A. Poirer, and D. J. Burnell, J. Phys. Chem. A, 1997, 101, 3371. Other problems with the 3-21G basis are that cation-metal distances tend to be too short (e.g. W. Rudolph, M. H. Brooker, and C. C. Pye, J. Phys. Chem., 1995, 99, 3793) and that adsorption energies of organics on aluminosilicates are overestimated, and charge separation is exaggerated (private communication from G. Sastre, Instituto de Technologica Quimica, Universidad Polytechnica de Valencia). Nevertheless, the 3–21G basisapparently usually gives good geometries (section5.5.1).

    Article  CAS  Google Scholar 

  42. P. M. Warner, J. Org. Chem., 1996, 61, 7192.

    Article  CAS  Google Scholar 

  43. J. E. Fowler, J. M. Galbraith, G. Vacek, and H. F. Schaefer, J. Am. Chem. Soc., 1994, 116, 9311. (b) G. Vacek, J. M. Galbraith, Y. Yamaguchi, H. F. Schaefer, R. H. Nobes, A. P. Scott, and L. Radom, J. Phys. Chem., 1994, 98, 8660.

    CAS  Google Scholar 

  44. The special theory of relativity (the one relevant to chemistry) and its chemical consequences are nicely reviewed in K. Balasubramanian, “Relativistic Effects in Chemistry,“ Parts A and B, Wiley, New York, 1997.

    Google Scholar 

  45. P. A. M. Dirac, Proc. R. Soc. 1929, A123, 714: “[trelativity is]... of no importance in the consideration of atomic and molecular structure, and ordinary chemical reactions...“

    CAS  Google Scholar 

  46. M. Krauss and W. J. Stevens, Annu. Rev. Phys.Chem., 1984, 35, 357; L. Szasz, “Pseudopotential Theory of Atoms and Molecules,“ Wiley, New York, 1985. (b) Relativistic Dirac-Fock calculations on closed-shell molecules: L. Pisani and E. Clementi, J. Comput. Chem., 1994, 15, 466.

    Article  CAS  Google Scholar 

  47. Gaussian 94 for Windows (G94W): Gaussian 94, Revision E. 1, M. J. Frisch, G. W. Trucks, H. B. Schlegel, P. M. W. Gill, B. G. Johnson, M. A. Robb, J. R. Cheeseman, T. Keith, G. A. Petersson, J. A. Montgomery, K. Raghavachari, M. A. Al-Laham, V. G. Zakrzewski, J. V. Ortiz, J. B. Foresman, J. Cioslowski, B. B. Stefanov, A. Nanayakkara, M. Challacombe, C. Y. Peng, P. Y. Ayala, W. Chen, M. W. Wong, J. L. Andres, E. S. Replogle, R. Gomperts, R. L. Martin, D. J. Fox, J. S. Binkley, D. J. Defrees, J. Baker, J. P. Stewart, M. Head-Gordon, C. Gonzalez, and J. A. Pople, Gaussian, Inc., Pittsburgh PA, 1995. G94 and G98 are available for both UNIX workstations and PCs.

    Google Scholar 

  48. Reference L. Radom, P. v. R. Schleyer, and J. A. Pople, “Ab Initio Molecular Orbital Theory,” Wiley, New York, 1986. [1g], p. 191.

    Google Scholar 

  49. Reference [1a], pp. 444, 494, 602–604.

    Google Scholar 

  50. A detailed review: G. Frenking, I. Antes, M. Böhme, S. Dapprich, A. W. Ehlers, V. Jonas, A. Neuhaus, M. Otto, R. Stegmann, A. Veldkamp, and S. Vyboishchikov, chapter 2 in Reviews in Computational Chemistry, Volume 8, K. B. Lipkowitz and D. B. Boyd, Eds., VCH, New York, 1996. (b) The main points of [51a] are presented in G. Frenking and U. Pidun, J. Chem. Soc., Dalton Trans., 1997, 1653. (c) T. R. Cundari, S. O. Sommerer, L. Tippett, J. Chem. Phys., 1995, 103, 7058.

    Google Scholar 

  51. J. Comp. Chem., 2002, 23, issue no. 8.

    Google Scholar 

  52. W. J. Hehre, W. W. Huang, P. E. Klunzinger, B. J. Deppmeier, and A. J. Driessen, “A Spartan Tutorial,“ Wavefunction Inc., Irvine, CA, 1997. (b) W. J. Hehre, J. Yu, and P. E. Klunzinger, “A Guide to Molecular Mechanics and Molecular Orbital Calculations in Spartan,“ Wavefunction Inc., Irvine, CA, 1997. (c) “A Laboratory Book of Computational Organic Chemistry,“ W. J. Hehre, A. J. Shusterman, and W. W. Huang, Wavefunction Inc., Irvine, C A, 1996.

    Google Scholar 

  53. R. Janoschek, Chemie in unserer Zeit, 1995, 29, 122.

    Article  CAS  Google Scholar 

  54. M. J. S. Dewar, “A Semiempirical Life,” Profiles, Pathways and Dreams series, J. I. Seeman, Ed., American Chemical Society, Washington, D.C., 1992, p. 185.

    Google Scholar 

  55. K. Raghavachari and J. B. Anderson, J. Phys. Chem., 1996, 100, 12960. (b) A historical review: P.-O. Löwdin, Int. J. Quantum Chem., 1995, 55, 77. (c) Fermi and Coulomb holes and correlation: [1c], pp. 296–297.

    Article  CAS  Google Scholar 

  56. Reference [1c], p. 286.

    Google Scholar 

  57. For example, A. C. Hurley, “Introduction to the Electron Theory of Small Molecules,” Academic Press, New York, 1976, pp. 286–288, or W. C. Ermler and C. W. Kern, J. Chem. Phys., 1974, 61, 3860.

    Google Scholar 

  58. P.-O. Löwdin, Advan. Chem. Phys., 1959, 2, 207.

    Google Scholar 

  59. A. P. Scott and L. Radom, J. Phys. Chem., 1996, 100, 16502.

    CAS  Google Scholar 

  60. 368 kJ mol−1: R. T. Morrison and R. N. Boyd, “Organic Chemistry,” 6th edn., Prentice Hall, Engelwood Cliffs, New Jersey, 1992, p. 21; 377 kJ mol−1: K. P. C. Vollhardt and N. E. Schore, “Organic Chemistry,” 2nd edn., Freeman, New York, 1987, p. 75.

    Google Scholar 

  61. For example, R. J. Bartlett and J. F. Stanton, chapter 2 in “Reviews in Computational Chemistry,” Vol. 5, K. B. Lipkowitz and D. B. Boyd, Eds., VCH, New York, 1994.

    Google Scholar 

  62. For example, the helium atom: [1a], pp. 256–259.

    Google Scholar 

  63. Brief introductions to the MP treatment of atoms and molecules: [1a], pp. 563–568; [1b], pp. 369–370; [1f] pp. 83–85.

    Google Scholar 

  64. Reference [1a], chapter 9.

    Google Scholar 

  65. C. Møller and M. S. Plesset, Phys. Rev., 1934, 46, 618.

    Google Scholar 

  66. J. S. Binkley and J. A. Pople, Int. J. Quantum Chem., 1975, 9, 229.

    Article  CAS  Google Scholar 

  67. Reference N. S. Ostlund, “Modern Quantum Chemistry,” McGraw-Hill New York, 1989 [1d], chapter 6.

    Google Scholar 

  68. For example, ref. [1b], pp. 367–368.

    Google Scholar 

  69. For example, ref. N. S. Ostlund, “Modern Quantum Chemistry,” McGraw-Hill New York, 1989 [1d], p. 353, [1f], p. 85.

    Google Scholar 

  70. A. Boldyrev, P. v. R. Schleyer, D. Higgins, C. Thomson, and S. S. Kramarenko, J. Comput. Chem., 1992, 9, 1066. Fluoro-and difluorodiazomethanes are minima by HF calculations, but are not viable minima by the MP2 method.

    Google Scholar 

  71. H 2 C=CHOH reaction: E. Lewars and I. Bonnycastle, J. Mol. Struct. (Theochem), 1997, 418, 17 and references therein. HNC reaction V. S. Rao, A. Vijay, A. K. Chandra, Can. J. Chem., 1996, 74, 1072. CH 3 NC reaction: The reported experimental activation energy is 161 kJ mol−1: F. W. Schneider and B. S. Rabinovitch, J. Am. Chem. Soc., 1962, 84, 4215; B. S. Rabinovitch and P. W. Gilderson, J. Am. Chem. Soc., 1965, 87, 158. The energy of CH3CN relative to CH3NC by a high-level (G2) calculation is −98.3 kJ mol−1 (E. Lewars). An early ab initio study of the reaction: D. H. Liskow, C. F. Bender, and H. F. Schaefer, J. Am. Chem. Soc., 1972, 95, 5178. A comparison of CH3CN, CH3NC, other isomers and radicals, cations and anions: P. M. Mayer, M. S. Taylor, M. Wong, and L. Radom, J. Phys. Chem. A, 1998, 102, 7074. Cyclopropylidene reaction: H. F. Bettinger, P. R. Schreiner, P. v. R. Schleyer, and H. F. Schaefer, J. Phys. Chem., 1996, 100, 16147.

    Article  CAS  Google Scholar 

  72. A superb brief introduction to CI is given in [1a], pp. 444–451, 557–562, and 568–573. (b) A comprehensive review of the developmenl of CI: I. Shavitt, Mol. Phys., 1998, 94, 3. (c) See also [1b], pp. 363–369; [1c], pp. 388–393; [1d], chapter 4; [1g], pp. 29–38.

    Google Scholar 

  73. N. Ben-Amor, S. Evangelisti, D. Maynau, and E. P. S. Rossi, Chem. Phys. Lett., 1998, 288, 348.

    Article  CAS  Google Scholar 

  74. R. B. Woodward and R. Hoffmann, “The Conservation of Orbital Symmetry,” Academic Press, New York, 1970, chapter 6.

    Google Scholar 

  75. Reference Æ. Frisch, “Exploring Chemistry with Electronic Structure Methods,” Gaussian Inc., Pittsburgh, PA, 1996 [1e], pp. 228–236 shows how to do CASSCF calculations. For CASSCF calculations on the Diels-Alder reaction, see Y. Li and K. N. Houk, J. Am. Chem. Soc., 1993, 115, 7478.

    Google Scholar 

  76. Reference N. S. Ostlund, “Modern Quantum Chemistry,” McGraw-Hill New York, 1989 [1d], chapter 6.

    Google Scholar 

  77. A paper boldly titled “Quadratic CI versus Coupled-Cluster theory...”: J. Hrusak, S. Ten-no, and S. Iwata, J. Chem. Phys., 1997, 106, 7185.

    CAS  Google Scholar 

  78. I. L. Alberts and N. C. Handy, J. Chem. Phys., 1988, 89, 2107.

    Article  CAS  Google Scholar 

  79. The water dimer has been extensively studied, theoretically and experimentally: (a) M. Schuetz, S. Brdarski, P.-O. Widmark, R. Lindh, and G. Karlström, J. Chem. Phys., 1997, 107, 4597; these workers report an interaction energy of 20.7 kJ mol−1 (4.94 kcal mol−1). (b) M. W. Feyereisen, D. Feller, and D. A. Dixon, J. Phys. Chem., 1996, 100, 2993; these workers report an interaction energy of 20.9 kJ mol−1 (5.0 kcal mol−1). (c) A. Halkier, H. Koch, P. Jorgensen, O. Christiansen, M. B. Nielsen, and T. Halgaker, Theor. Chem. Acc., 1997, 97, 150. (d) M. S. Gordon and J. H. Jensen, Acc. Chem. Res., 1996, 29, 536.

    Google Scholar 

  80. For discussions of BSSE and the counterpoise method see: T. Clark, “A Handbook of Computational Chemistry,” Wiley, New York, 1985, pp. 289–301. (b) J. M. Martin in “Computational Thermochemistry,” K. K. Irikura and D. J. Frurip, Eds., American Chemical Society, Washington,D.C., 1998, p. 223. (c) References [80] give leading references to BSSE and [80(a)] describes a method for bringing the counterpoise correction closer to the basis set limit. (d) Methods designed to be free of BSSE: G. J. Halasz, A. Vibok, I. Mayer, J. Comput. Chem., 1999, 20, 274.

    Google Scholar 

  81. L. M. Balbes, S. W. Mascarella, D. B. Boyd, chapter 7 in Reviews in Computational Chemistry, Volume 5, K. B. Lipkowitz and D. B. Boyd, Eds., VCH, New York, 1994. (b) A. Tropsha and J. P. Bowen, chapter 17 in “Using Computers in Chemistry and Chemical Education,” T. J. Zielinski and M. L. Swift, Eds., American Chemical Society, Washington D.C., 1997. (c) H.-D. Höltje and G. Folkers, “Molecular Modelling,” VCH, New York, 1997. (d) C. E. Bugg, W. M. Carson, J. A. Montgomery, Scientific American, 1993, December, 92. (e) J. L. Vinter and M. Gardner, “Molecular Modelling and Drug Design,” Macmillan, London, 1994. (f) (e) P. M. Dean, “Molecular Foundations of Drug-Receptor Interactions,” Cambridge University Press, Cambridge, 1987.

    Google Scholar 

  82. Reference J. F. Stanton, chapter 2 in “Review in Computational Chemistry,” Vol. 5, K. B. Lipkowitz and D. B. Boyd, Eds., VCH, New York, 1994 [62], p. 106.

    Google Scholar 

  83. U. Burkert and N. L. Allinger, “Molecular Mechanics,” ACS Monograph 177, American Chemical Society, Washington, D.C., 1982; pp. 6–10. See also B. Ma, J.-H. Lii, H. F. Schaefer, N. L. Allinger, J. Phys. Chem., 1996, 100, 8763; M. Ma, J.-H. Lii, K. Chen, N. L. Allinger, J. Am. Chem. Soc., 1997, 119, 2570.

    Google Scholar 

  84. A. Domenicano and I. Hargittai, Eds., “Accurate Molecular Structures,” Oxford University Press, New York, 1992. (b) A “wake-up call”: V. G. S. Box, Chem. & Eng. News, 2002, February 18, 6.

    Google Scholar 

  85. G. A. Peterson in chapter 13, “Computational Thermochemistry,” K. K. Irikura and D. J. Frurip, Eds., American Chemical Society, Washington, D.C., 1998.

    Google Scholar 

  86. Reference L. Radom, P. v. R. Schleyer, and J. A. Pople, “Ab Initio Molecular Orbital Theory,” Wiley, New York, 1986 [1g], pp. 133–226; note the summary on p. 226.

    Google Scholar 

  87. Observations by the author. Others have noted that planar hexaazabenzene is a relative minimum with Hartree-Fock calculations, but a hilltop at correlated levels, e.g. R. Engelke, J. Phys. Chem., 1992, 96, 10789 (HF/4-31G, HF/4-31G*, MP2/6-31G*).

    Article  CAS  Google Scholar 

  88. Reference Æ. Frisch, “Exploring Chemistry with Electronic Structure Methods,” Gaussian Inc., Pittsburgh, PA, 1996 [1e], p. 118.

    Google Scholar 

  89. Reference Æ. Frisch, “Exploring Chemistry with Electronic Structure Methods,” Gaussian Inc., Pittsburgh, PA, 1996 [1e], pp. 118 (ozone) and 128 (FOOF).

    Google Scholar 

  90. Reference Æ. Frisch, “Exploring Chemistry with Electronic Structure Methods,” Gaussian Inc., Pittsburgh, PA, 1996 [1e], p. 36; other calculations on ozone are on pp. 118, 137 and 159.

    Google Scholar 

  91. Other examples of problems with fluorine and/or other halogens: (a) the G2 method (section 5.5.2) is relatively inaccurate (errors of up to 33 kJ mol−1 for molecules with multiple halogens, cf. the average error of 6.6 kJ mol−1 for a set of 148 molecules: L. A. Curtiss, K. Raghavachari, P. C. Redfern, J. A. Pople, J. Chem. Phys., 1997, 106, 1063. (b) A (very?) good ab initio geometry for H2CCH2 required a very large basis (HF/6-311++G(3df,2pd)): D. Diederdorf, Applied Research Associates Inc., personal communication).

    CAS  Google Scholar 

  92. For example, Fluoroethanes: R. D. Parra and X. C. Zeng, J. Phys. Chem. A, 1998, 102, 654. Fluoroethers: D.A. Good and J. S. Francisco, J. Phys. Chem. A, 1998, 102, 1854.

    Article  CAS  Google Scholar 

  93. C. A. Coulson, “Valence,” 2nd edn, Oxford University Press, London, 1961, p. 91.

    Google Scholar 

  94. “Computational Thermochemistry,” K. K. Irikura and D. J. Frurip, Eds., American Chemical Society, Washington, D.C., 1998.

    Google Scholar 

  95. M. L. McGlashan, “Chemical Thermodynamics,” Academic Press, London, 1979. (b) L. K. Nash, “Elements of Statistical Thermodynamics,” Addison-Wesley, Reading, MA, 1968. (c) A good, brief introduction to statistical thermodynamics is given by K. K. Irikura in [95], Appendix B.

    Google Scholar 

  96. For example, P. W. Atkins, “Physical Chemistry,” 6th edn, Freeman, New York, 1998.

    Google Scholar 

  97. R. S. Treptow, J. Chem. Educ., 1995, 72, 497.

    CAS  Google Scholar 

  98. See K. K. Irikura and D. J. Frurip, chapter 1, S. W. Benson and N. Cohen, chapter 2, and M. R. Zachariah and C. F. Melius, chapter 9, in D. J. Frurip, Eds., America Chemical Society, Washington, D.C., 1998 [95].

    Google Scholar 

  99. The bond energies were taken from M. A. Fox and J. K. Whitesell, “Organic Chemistry,” Jones and Bartlett, Boston, 1994, p. 72.

    Google Scholar 

  100. For good accounts of the history and meaning of the concept of entropy, see: (a) H. C. von Baeyer, “Maxwell’s Demon. Why warmth disperses and time passes,” Random House, New York, 1998. (b) G. Greenstein, “Portraits of Discovery. Profiles in Scientific Genius,” chapter 2 (“Ludwig Boltzmann and the second law of thermodynamics”), Wiley, New York, 1998.

    Google Scholar 

  101. Reference L. Radom, P. v. R. Schleyer, and J. A. Pople, “Ab Initio Molecular Orbital Theory,” Wiley, New York, 1986 [1g], section 6.3.9.

    Google Scholar 

  102. A sophisticated study of the calculation of gas-phase equilibrium constants: F. Bohr and E. Henon, J. Phys. Chem. A, 1998, 102, 4857.

    Article  CAS  Google Scholar 

  103. A very comprehensive treatment of rate constants, from theoretical and experimental viewpoints, is given in J. I. Steinfeld, J. S. Francisco, and W. L. Hase, “Chemical Kinetics and Dynamics,” Prentice Hall, New Jersey, 1999.

    Google Scholar 

  104. For the Arrhenius equation and problems associated with calculations involving rate constants and transition states see J. L. Durant in D. J. Frurip, Eds., America Chemical Society, Washington, D.C., 1998 [95], chapter 14.

    Google Scholar 

  105. Reference [97], p. 949.

    Google Scholar 

  106. P. W. Atkins, “Physical Chemistry,” 4th edn, Freeman, New York, 1990, p. 859.

    Google Scholar 

  107. Reference N. L. Allinger, J. Comp. Chem., 1996, 17, 730 [32], chapter 2.

    CAS  Google Scholar 

  108. Reference N. L. Allinger, J. Comp. Chem., 1996, 17, 730 [32], section 4.2.

    CAS  Google Scholar 

  109. Reference L. Radom, P. v. R. Schleyer, and J. A. Pople, “Ab Initio Molecular Orbital Theory,” Wiley, New York, 1986 [1g], section 7.2.2.

    Google Scholar 

  110. Reference [1g], section 7.2.4. (b) D. B. Chestnut, J. Comput. Chem., 1995, 16, 1227; D. B. Chestnut, J. Comput. Chem., 1997, 18, 584 (c) P. K. Freeman, J. Am. Chem. Soc, 1998, 120, 1619. (d) P. Politzer, M. E. Grice, J. S. Murray, and J. M. Seminario, Can J. Chem., 1993, 71, 1123. (e) For calculation of aromatic character by nonisodesmic methods (G2 calculations on cyclopropenone) see D. W. Rogers, F. L. McLafferty, and A. W. Podosenin, J. Org. Chem, 1998, 63, 7319.

    Google Scholar 

  111. By another approach (diagonal strain energy, relating three-membered rings to essentially strainless six-membered rings), the strain energies of cyclopropane and oxirane have been calculated to be 117 and 105 kJ mol−1, respectively: A. Skancke, D. Van Vechten, J. F. Liebman, and P. N. Skancke, J. Mol. Struct., 1996, 376, 461.

    Article  CAS  Google Scholar 

  112. For example, N. L. Allinger, J. Comp. Chem., 1996, 17, 730 [32], section 4.2. One problem with the 3-21G basis is that it tends to flatten tricoordinate nitrogen too much [42].

    CAS  Google Scholar 

  113. Reference D. J. Frurip, Eds., America Chemical Society, Washington, D.C., 1998 [95], p. 940.

    Google Scholar 

  114. Errors of about 30–60 kJ mol−1 have been noted for isogyric reactions: L. Radom, P. v. R. Schleyer, and J. A. Pople, “Ab Initio Molecular Orbital Theory,” Wiley, New York, 1986 [102], p. 13.

    Google Scholar 

  115. For example K. B. Wiberg and J. W. Ochterski, J. Comput. Chem., 1997, 18, 108.

    Article  CAS  Google Scholar 

  116. General discussions: (a) D. A. Ponomarev and V. V. Takhistov, J. Chem. Educ., 1997, 74, 201. (b) Ref. [1e], pp. 181–184 and 204–207. (c) Ref. [1g], pp. 271, 293, 298, 356. (d) Ref. [1a], pp. 595–597.

    CAS  Google Scholar 

  117. G2 method: L. A. Curtiss, K. Raghavachari, G. W. Trucks, and J. A. Pople, J. Chem. Phys., 1991, 94, 7221. There have been many reviews of the G2 method, e.g. L. A. Curtiss and K. Raghavachari, in [102], chapter 10. (b) G3 method: L. A. Curtiss, K. Raghavachari, P. C. Redfern, V. Rassolov, and J. A. Pople, J. Chem. Phys., 1998, 109, 7764. Variations on G3: G3(MP2), L. A. Curtiss, P. C. Redfern, K. Raghavachari, V. Rassolov, and J. A. Pople, J. Chem. Phys., 1999, 110, 4703; G3(B3) and G3(MP2B3), A. G. Baboul, L. A. Curtiss, P. C. Redfern, and K. Raghavachari, J. Chem. Phys., 1999, 110, 7650.

    Article  CAS  Google Scholar 

  118. J. A. Pople, M. Head-Gordon, D. J. Fox, K. Raghavachari, and L. A. Curtiss, J. Chem. Phys., 1989, 90, 5622.

    CAS  Google Scholar 

  119. Reference Æ. Frisch, “Exploring Chemistry with Electronic Structure Methods,” Gaussian Inc., Pittsburgh, PA, 1996. [1e], chapter 7.

    Google Scholar 

  120. L. A. Curtiss, J. E. Carpenter, K. Raghavachari, and J. A. Pople, J. Chem. Phys., 1992, 96, 9030.

    CAS  Google Scholar 

  121. B. J. Smith and L. Radom, J. Am. Chem. Soc., 1993, 115, 4885.

    CAS  Google Scholar 

  122. Reference D. J. Frurip, Eds., American Chemical Society, Washington, D.C., 1998 [86], p. 244.

    Google Scholar 

  123. J. M. L. Martin in chapter 12, D. J. Frurip, Eds., America Chemical Society, Washington, D.C., 1998 [95].

    Google Scholar 

  124. Reference [97], section 2.8.

    Google Scholar 

  125. M. W. Chase, Jr, J. Phys. Chem. Ref. Data, Monograph 9, 1998, 1–1951. NIST-JANAF Thermochemical Tables, 4th edn.

    Google Scholar 

  126. A. Nicolaides, A. Rauk, M. N. Glukhovtsev, and L. Radom, J. Phys. Chem., 1996, 100, 17460.

    Article  CAS  Google Scholar 

  127. J. Hine and K. Arata, Bull. Soc. Chem. Jpn., 1976, 49, 3089:-201 kJ mol−1; J. H. S. Gree, Chem. Ind. (London), 1960, 1215: −201 ±0.2kJmol−1.

    CAS  Google Scholar 

  128. Very convenient and useful: the NIST (National Institute of Standards and Technology, Gaithersburg, MD, USA) website: http://webbook.nist.gov/chemistry/ (b) S. G. Lias, J. E. Bartmess, J. F. L. Holmes, R. D. Levin, W. G. Mallard, J. Phys. Chem. Ref. Data, 1988, 77, Suppl. 1., American Chemical Society and American Institute of Physics, 1988. (c) J. B. Pedley, “Thermochemical Data and Structures of Organic Compounds,” Thermodynamics Research Center, College Station, Texas, 1994.

  129. Worked examples, with various fine points: K. K. Irikura and D. J. Frurip, in L. Radom, P. v. R. Schleyer, and J. A. Pople, “Ab Initio Molecular Orbital Theory,” Wiley, New York, 1986 [102], Appendix C. (b) Heats of formation of neutral and cationic chloromethanes: C. F. Rodrigues, D. K. Bohme, A. C. Hopkinson, J. Phys. Chem., 1996, 100, 2942. (c) Heats of formation, entropies and enthalpies of neutral and cationic enols: F. Turecek and C. J. Cramer, J, Am. Chem. Soc., 1995, 117, 12243. of neutral and cationic enols: (d) Heats of formation by ab initio and molecular mechanics: D. F. DeTar, J. Org. Chem., 1995, 60, 7125. (e) Heats of formation and antiaromaticity in strained molecules: M. N. Glukhovtsev, S. Laiter, A. Pross, J. Phys. Chem., 1995, 99, 6828. (f) Heats of formation of organic molecules with the aid of ab initio and group equivalent methods: L. R. Schmitz and Y. R. Chen, J. Comput. Chem., 1994, 15, 1437. (f) Isodesmic reactions in ab initio calculation of heat of formation of cyclic C6 hydrocarbons and benzene isomers: Z. Li, D. W. Rogers, F. J. McLafferty, M. Mandziuk, and A. V. Podosenin, J. Phys. Chem. A, 1999, 103, 426. (g) Isodesmic reactions in ab initio calculation of heat of formation of benzene isomers: Y.-S. Cheung, C.-K. Wong, and W.-K. Li, Mol. Struct. (Theochem), 1998, 454, 17.

    Google Scholar 

  130. DeL. F. DeTar, J. Org. Chem., 1998, 102, 5128.

    Google Scholar 

  131. S. S. Shaik, H. B. Schlegel, S. Wolfe, ‘Theoretical Aspects of Physical Organic Chemistry. The SN2 mechanism’, Wiley, New York, 1992; pp. 50–51.

    Google Scholar 

  132. Reference [1l], pp. 474–489. Discussion of computation of activation energies: pp. 495–497.

    Google Scholar 

  133. J. M. Haile, ‘Molecular Dynamics Simulation. Elementary Methods,’ Wiley, New York, 1992. (b) A tutorial in MD is available from the website of S. Deiana and B. Manunza, University of Sassani, Italy: http://antas.agraria.uniss.it (c) An application of MD to chemical reactions: B. K. Carpenter, Ang. Chem. Int. Ed. Engl., 1998, 37, 3341.

    Google Scholar 

  134. Ref. [1a], section 2.5. (b) Ref. [104], section 12.3. (c) R. P. Bell, ‘The Tunnel Effect in Chemistry,’ Chapman and Hall, London, 1980.

    Google Scholar 

  135. S. S. Shaik, H. B. Schlegel, S. Wolfe, ‘Theoretical Aspects of Physical Organic Chemistry. The SN2 mechanism,’ Wiley, New York, 1992; pp. 84–88. (b) Ref. [104], chapters 25, 26, 27. (c) Kinetics of halocarbons reactions: R. J. Berry, M. Schwartz, P. Marshall in [95], chapter 18. (d) Ref. [104], particularly chapters 7, 8, 10, 11, 12. (e) The ab initio calculation of rate constants is given in some detail in these two references: D. M. Smith, A. Nicolaides, B. T. Golding, and L. Radom, J. Am. Chem. Soc., 1998, 120, 10223; J. P. A. Heuts, R. G Gilbert, L. Radom, Macromolecules, 1995, 28, 8771; [11], pp. 471–489, 492–497.

    Google Scholar 

  136. Reference J. S. Francisco, and W. L. Hase, ‘Chemical Kinetics and Dynamics,’ Prentice Hall, New Jersey, 1999 [104], chapter 11.

    Google Scholar 

  137. Reference J. S. Francisco, and W. L. Hase, ‘Chemical Kinetics and Dynamics,’ Prentice Hall, New Jersey, 1999 [104], p. 350.

    Google Scholar 

  138. Reference J. S. Francisco, and W. L. Hase, ‘Chemical Kinetics and Dynamics,’ Prentice Hall, New Jersey, 1999 [104], p. 352, Fig. 11–11.

    Google Scholar 

  139. Calculated from the Arrhenius parameters in [97], p. 949, Table 25.4.

    Google Scholar 

  140. J. F. McGarrity, A. Cretton, A. A. Pinkerton, D. Schwarzenbach, and H. D. Flack, Angew. Chem. int. Ed. Engl., 1983, 22, 405; C.E. Blom and A. Bauder, J. Am. Chem. Soc., 1984, 106, 4029; B. Capon, B. Guo, F. C. Kwok, A. K. Siddhanta, C. Zucco, Acc. Chem. Res., 1988, 21, 135.

    Google Scholar 

  141. D. J. DeFrees and A. D. McLean, J. Chem Phys., 1982, 86, 2835.

    Google Scholar 

  142. W. Runge in ‘The Chemistry of Ketenes, Allenes and Related Compounds,’ S. Patai, Ed., Wiley, New York, 1980, p. 45. (b) E. Hirota and C. Matsumura, J. Chem Phys., 1973, 59, 3038.

    Google Scholar 

  143. O. L. Chapman, Pure Appl. Chem., 1974, 40, 511.

    CAS  Google Scholar 

  144. I. R. Dunkin, ‘Matrix Isolation Techniques: A Practical Approach. The Practical Approach in Chemistry Series,’ Oxford University Press, New York, 1998.

    Google Scholar 

  145. Reference [32], chapter 6.

    Google Scholar 

  146. Reference Æ. Frisch, ‘Exploring Chemistry with Electronic Structure Methods,’ Gaussian Inc., Pittsburgh, PA, 1996 [1e], pp. 146–148, 157–158.

    Google Scholar 

  147. J. B. Foresman, personal communication, October 1998.

    Google Scholar 

  148. Reference [32], pp. 58–62.

    Google Scholar 

  149. For introductions to the theory and interpretation of IR, UV and NMR spectra, see R. M. Silverstein and F. X. Webster, ‘Spectrometric Identification of Organic Compounds,’ 6th edn, Wiley, New York, 1997.

    Google Scholar 

  150. K. P. Huber and G. Herzberg, ‘Molecular Spectra and Molecular Structure. IV. Constants of Diatomic Molecules,’ Van Nostrand Reinhold, New York, 1979.

    Google Scholar 

  151. Reference L. Radom, P. v. R. Schleyer, and J. A. Pople, ‘Ab Initio Molecular Orbital Theory,’ Wiley, New York, 1986 [1g], pp. 234, 235.

    Google Scholar 

  152. For example, ‘... it is unfair to compare frequencies calculated within the harmonic approximation with experimentally observed frequencies...’: A. St-Amant, chapter 2, p. 235 in Reviews in Computational Chemistry, Volume 7, K. B. Lipkowitz and D. B. Boyd, Eds., VCH, New York, 1996.

    Google Scholar 

  153. Reference [1i], pp. 271–274.

    Google Scholar 

  154. J. G. Radziszewski, B. A. Hess, Jr., and R. Zahradnik, J. Am. Chem. Soc., 1992, 114, 52. (b) C. Wentrup, R. Blanch, H. Briel, G. Gross, J. Am. Chem. Soc., 1988, 110, 1874. (c) O. L. Chapman, C. C. Chang, J. Kolc, N. R. Rosenquist, H. Tomioka, J. Am. Chem. Soc., 1975, 97, 6586. (d) O. L. Chapman, K. Mattes, C. L. McIntosh, J. Pacansky, G. V. Calder, and G. Orr, J. Am. Chem. Soc., 1973, 95, 6134.

    Article  CAS  Google Scholar 

  155. A. Komornicki and R. L. Jaffe, J. Chem. Phys., 1979, 71, 2150. (b) Y. Yamaguchi, M. Frisch, J. Gaw, H. F. Schaefer, and J. S. Binkley, J. Chem. Phys., 1986, 84, 2262. (c) M. Frisch, Y. Yamaguchi, H. F. Schaefer, and J. S. Binkley, J. Chem. Phys., 1986, 84, 531. (d) R. D. Amos, Chem. Phys. Lett, 1984, 108, 185. (e) J. E. Gready, G. B. Bacskay, and N. S. Hush, J. Chem. Phys., 1978, 90, 467.

    Article  CAS  Google Scholar 

  156. H. Lampert, W. Mikenda, and A. Karpfen, J. Phys. Chem. A, 1997, 101, 2254. This paper shows actual pictures of experimental and calculated (HF, MP2, DFT) spectra for phenol, benzaldehyde and salicylaldehyde. (b) M. D. Halls and H. B. Schlegel, J. Chem. Phys., 1998, 109, 10587. For a series of small molecules absolute intensities compared with QCISD as benchmark, (c) Ref. [69], pp. 118–121. (d) D. H. Magers, E. A. Salter, R. J. Bartlett, C. Salter, B. A. Hess, Jr., and L. J. Schaad, J. Am. Chem. Soc., 1988, 110, 3435 (comments on intensities on p. 3439).

    CAS  Google Scholar 

  157. For a good review of the cyclobutadiene problem, see B. K. Carpenter in ‘Advances in Molecular Modelling,’ D. Liotta, Ed., JAI Press Inc., Greenwich, Connecticut, 1988.

    Google Scholar 

  158. Theoretical calculation of dipole moments: [1a], pp. 399–402. (b) Measurement and applications of dipole moments: O. Exner, ‘Dipole Moments in Organic Chemistry,’ Georg Thieme Publishers, Stuttgart, 1975.

    Google Scholar 

  159. Tables: A. L. McClellan, ‘Tables of Experimental Dipole Moments,’ vol. 1, W. H. Freeman, San Francisco, CA, 1963; vol. 2, Rahara Enterprises, El Cerrita, CA, 1974.

    Google Scholar 

  160. For example, J. F. Stanton, chapter 2 in ‘Reviews in Computational Chemistry,’ Vol. 5, K. B. Lipkowitz and D. B. Boyd, Eds., VCH, New York, 1994 [62], p. 152.

    Google Scholar 

  161. The reason why an electron pair forms a covalent bond has apparently not been settled. See (a) Ref. [1a], pp. 362–363. (b) G. B. Backsay, and J. R. Reimers, S. Nordholm, J. Chem. Ed., 1997, 74, 1494.

    Google Scholar 

  162. For example, (a) Electron density on an atom: G. W. Wheland and L. Pauling, J. Am. Chem. Soc., 1935, 57, 2086. (b) Pi-bond order: C. A. Coulson, Proc. Roy. Soc., 1939, A169, 413.

    Article  CAS  Google Scholar 

  163. R. S. Mulliken, J. Chem Phys., 1955, 23, 1833. (b) R. S. Mulliken, J. Chem Phys., 1962, 36, 3428. (c) Ref. [1a], pp. 475–478.

    CAS  Google Scholar 

  164. P.-O. Löwdin, Advances in Quantum Chemistry, 1970, 5, 185.

    Google Scholar 

  165. A. E. Reed, L. A. Curtiss, and F. Weinhold, Chem. Rev., 1988, 88, 899.

    Article  CAS  Google Scholar 

  166. T. Brinck, “Theoretical Organic Chemistry,” C. Parkanyi, Ed., Elsevier, New York, 1998. (b) D. S. Marynick, J. Comp. Chem., 1997, 18, 955.

    Google Scholar 

  167. Use of bond orders in deciding if a covalent bond is present: P. v. R. Schleyer, P. Buzek, T. Müller, Y. Apelloig, and H.-U. Siehl, Angew. Chem. Int. Ed. Engl., 1993, 32, 1471. (b) Use of bond order in estimating progress along a reaction coordinate: G. Lendvay, J. Phys. Chem., 1994, 98, 6098.

    Google Scholar 

  168. R. F. W. Bader, “Atoms in Molecules,” Oxford University Press, Oxford, 1990. (b) R. W. F. Bader, P. L. A. Popelier, T. A. Keith, Angew. Chem. Int. Ed. Engl., 1994, 33, 620. Applications of AIM theory: (c) I. Rozas, I. Alkorta, J. Elguero, J. Phys. Chem., 1997, 101, 9457. (d) Can. J. Chem., 1996, 74, issue dedicated to R. F. W. Bader. (e) R. J. Gillespie, and E. A. Robinson, Angew. Chem. int. Ed. Engl., 1996, 35, 495. (f) S. Grimme, J. Am. Chem. Soc., 1996, 118, 1529. (g) O. Mo, M. Yanez, M. Eckert-Maksoc, Z. B. Maksic, J. Org. Chem., 1995, 60, 1638. (h) K. M. Gough and J. Millington, Can. J. Chem., 1995, 73, 1287. (i) R. Glaser and G. S.-C. Choy, J. Am. Chem. Soc., 1993, 115, 2340. (j) J. Cioslowski and T. Mixon, J. Am. Chem. Soc., 1993, 115, 1084.

    Google Scholar 

  169. Reference P. Buzek, T. Müller, Y. Apelloig, and H.-U. Siehl, Angew. Chem. Int. Ed. Engl., 1993, 32, [170a], p. 175.

    Google Scholar 

  170. Reference Æ. Frisch, “Exploring Chemistry with Electronic Structure Methods,” Gaussian Inc., Pittsburgh, PA, 1996. [1e], chapter 9.

    Google Scholar 

  171. T. Helgaker, M. Jaszuński, and K. Ruud, Chem. Rev., 1999, 99, 294.

    Article  Google Scholar 

  172. Ref. Æ. Frisch, “Exploring Chemistry with Electronic Structure Methods,” Gaussian Inc., Pittsburgh, PA, 1996. [1e], pp. 53–54 and 104–105. (b) J. R. Cheeseman, G. W. Trucks, T. A. Keith, and M. J. Frisch, J. Chem. Phys., 1996, 104, 5497.

    Google Scholar 

  173. Review: J. Gauss, Ber. Bunsenges. Phys. Chem., 1995, 99, 1001. (b) J. Gauss, J. Chem. Phys., 1993, 99, 3629. (c) P. v. R. Schleyer, J. Gauss, M. Bühl, R. Greatrex, and M. A. Fox, Chem. Commun., 1993, 1766. (d) Electron correlation and coupling constants: S. A. Perera, M. Nooijen, and R. J. Bartlett, J. Chem. Phys., 1996, 104, 3290.

    CAS  Google Scholar 

  174. A. D. Wolf, V. V. Kane, R. H. Levin, and M. Jones, J. Am. Chem. Soc., 1973, 95, 1680.

    CAS  Google Scholar 

  175. V. I. Minkin, M. N. Glukhovtsev, and B. Ya. Simkin, “Aromaticity and Antiaromaticity: Electronic and Structural Aspects,” Wiley, New York, 1994; chapter 2. (b) A simple magnetic criterion for aromaticity (NICS): P. v. R. Schleyer, C. Maerker, A. Dransfeld, H. Jiao, and N. v. E. Hommes, J. Am. Chem. Soc., 1996, 118, 6317. (c) NICS(2.0): T. Veszprémi, M. Tahahashi, B. Hajgató, J. Ogasawara, K. Skamoto, and M. Kira, J. Phys. Chem., 1998, 102, 10530–10535; R. West, J. J. Buffy, M. Haaf, T. Müller, B. Gehrhus, M. F. Lappert, and Y. Apeloig, J. Am. Chem. Soc., 1998, 120, 1639. (d) Analysis of NICS: I. Morao and F. P. Cossío, J. Org. Chem., 1999, 64, 1868.

    Google Scholar 

  176. Reference [1b], pp. 276–277, 288, 372–373.

    Google Scholar 

  177. R. D. Levin and S. G. Lias, “Ionization Potential and Appearance Potential Measurements, 1971–1981,” National Bureau of Standards, Washington, DC, 1982.

    Google Scholar 

  178. See e.g. [1b], pp. 361–363; [1c], pp. 278–280; [1d], pp. 127–128; [1g], pp. 24, 116. A novel look at Koopmans’ theorem: C. Angeli, J. Chem. Ed., 1998, 75, 1494. (b) T. Koopmans, Physica, 1934, 1, 104.

    Google Scholar 

  179. M. B. Smith and J. March, “March’s Advanced Organic Chemistry,” 5th edn, Wiley, New York, 2001, pp. 10–12.

    Google Scholar 

  180. J. E. Lyons, D. R. Rasmussen, M. P. McGrath, R. H. Nobes, and L. Radom, Ang. Chem. Int. Ed. Engl., 1994, 33, 1667.

    Google Scholar 

  181. L. A. Curtiss and K. Raghavachari, chapter 10 in [102].

    Google Scholar 

  182. L. A. Curtiss, R. H. Nobes, J. A. Pople, and L. Radom, J. Chem Phys., 1992, 97, 6766.

    CAS  Google Scholar 

  183. “Data Visualization in Molecular Science: Tools for Insight and Innovation,” J. E. Bower, Ed., Addison-Wesley, Reading, MA, 1995. (b) “Frontiers of Scientific Visualization,” C. Pickover and S. Tewksbury, Wiley, 1994. (c) http://www.tc.cornell.edu/~richard/.

    Google Scholar 

  184. Gauss View: Gaussian Inc., Carnegie Office Park, Bldg. 6, Pittsburgh, PA 15106, USA.

    Google Scholar 

  185. E. L. Eliel and S. H. Wilen, “Stereochemistry of Carbon Compounds,” Wiley, New York, 1994, pp. 502–507 and 686–690.

    Google Scholar 

  186. E. Lewars, unpublished.

    Google Scholar 

  187. The term is not just whimsy on the author’s part: certain stereoelectronic phenomena arising from the presence of lone pairs on heteroatoms in a 1,3-relationship were once called the “rabbit-ear effect,” and a photograph of the eponymous creature even appeared on the cover of the Swedish journal Kemisk Tidskrift. History of the term, photograph: E. L. Eliel, “From Cologne to Chapel Hill,” American Chemical Society, Washington, DC, 1990, pp. 62–64.

    Google Scholar 

  188. E. Lewars, J. Mol. Struct. (Theochem), 2000, 507, 165; E. Lewars, J. Mol. Struct. (Theochem), 1998, 423, 173.

    Article  CAS  Google Scholar 

  189. P. Politzer and J. S. Murray, chapter 7 in Reviews in Computational Chemistry, Volume 2, K. B. Lipkowitz and D. B. Boyd, Eds., VCH, New York, 1996.

    Google Scholar 

  190. Reference A. J. Shusterman, and W. W. Huang, Wavefunction Inc., Irvine, CA, 1996. [53c] pp. 141–142.

    Google Scholar 

  191. W. J. Hehre, A. J. Shusterman, and J. E. Nelson, “The Molecular Modelling Workbook of Organic Chemistry,” Wavefunction Inc., Irvine, CA, 1998 (book and CD).

    Google Scholar 

  192. M. J. S. Dewar, “A Semiempirical Life,” Profiles, Pathways and Dreams series, J. I. Seeman, Ed., American Chemical Society, Washington, D.C., 1992, p. 129.

    Google Scholar 

  193. Coulson’s remarks: J. D. Bolcer and R. B. Hermann, chapter 1 in Reviews in Computational Chemistry, Volume 5, K. B. Lipkowitz and D. B. Boyd, Eds., VCH, New York, 1996, p. 12. (b) The increase in computer speed is also dramatically shown in data provided in Gaussian News, 1993, 4, 1. The approximate times for a single-point HF/6-31G** calculation on 1,3,5-triamino-2,4,6-trinitrobenzene (300 basis functions) are reported as: ca. 1967, on a CDC 1604, 200 years (estimated); ca. 1992, on a 486 DX personal computer, 20 hours. This is a speed factor of 90,000 in 25 years. The price factor for the machines may not be as dramatic, but suffice it to say that the CDC 1604 was not considered a personal computer.

    Google Scholar 

Download references

Rights and permissions

Reprints and permissions

Copyright information

© 2004 Kluwer Academic Publishers

About this chapter

Cite this chapter

(2004). Ab initio Calculations. In: Computational Chemistry. Springer, Boston, MA. https://doi.org/10.1007/0-306-48391-2_5

Download citation

  • DOI: https://doi.org/10.1007/0-306-48391-2_5

  • Publisher Name: Springer, Boston, MA

  • Print ISBN: 978-1-4020-7285-7

  • Online ISBN: 978-0-306-48391-2

  • eBook Packages: Springer Book Archive

Publish with us

Policies and ethics