全部 标题 作者
关键词 摘要

OALib Journal期刊
ISSN: 2333-9721
费用:99美元

查看量下载量

相关文章

更多...

Philicity and Fugality Scales for Organic Reactions

DOI: 10.1155/2014/541547

Full-Text   Cite this paper   Add to My Lib

Abstract:

Theoretical scales of reactivity and selectivity are important tools to explain and to predict reactivity patterns, including reaction mechanisms. The main achievement of these efforts has been the incorporation of such concepts in advanced texts of organic chemistry. In this way, the modern organic chemistry language has become more quantitative, making the classification of organic reactions an easier task. The reactivity scales are also useful to set up a number of empirical rules that help in rationalizing and in some cases anticipating the possible reaction mechanisms that can be operative in a given organic reaction. In this review, we intend to give a brief but complete account on this matter, introducing the conceptual basis that leads to the definition of reactivity indices amenable to build up quantitative models of reactivity in organic reactions. The emphasis is put on two basic concepts describing electron-rich and electron-deficient systems, namely, nucleophile and electrophiles. We then show that the regional nucleophilicity and electrophilicity become the natural descriptors of electrofugality and nucleofugality, respectively. In this way, we obtain a closed body of concepts that suffices to describe electron releasing and electron accepting molecules together with the description of permanent and leaving groups in addition, nucleophilic substitution and elimination reactions. 1. Introduction The development of reactivity indices to describe organic reactivity has been an active area of research from the dawn of theoretical physical organic chemistry [1–4]. From the earlier semiempirical models proposed in Hückel molecular orbital (HMO) theory, reactivity was described with the aid of static first order reactivity descriptors like atomic charges, free valence index, and bond orders [5, 6]. These reactivity indices were formerly developed around the ground state of reactants. The next generation of reactivity indices began with the elegant theory proposed by Coulson and Longuet-Higgins, in a series of papers describing the response functions, including second order quantities like atomic and bond polarizabilities [7–9]. It is worth emphasizing that, at that time, HMO and Coulson-Longuet-Higgins theories were conceived not as methods to approximately solve the Schr?edinger equation but as models of chemical bond. The third generation of reactivity indices started after the pioneering work of Gilles Klopman, who introduced the concept of charge and frontier controlled reactions, including solvation effects [4]. Nowadays, the treatment of

References

[1]  J. D. Roberts and A. Streitwieser Jr., “Quantum mechanical calculations of orientation in aromatic substitution,” Journal of the American Chemical Society, vol. 74, no. 18, pp. 4723–4725, 1952.
[2]  A. A. Frost and B. Musulin, “A mnemonic device for molecular orbital energies,” The Journal of Chemical Physics, vol. 21, no. 3, pp. 572–573, 1953.
[3]  L. Salem, “Intermolecular orbital theory of the interaction between conjugated systems. I. General theory,” Journal of the American Chemical Society, vol. 90, no. 3, pp. 543–552, 1968.
[4]  G. Klopman, “Chemical reactivity and the concept of charge- and frontier-controlled reactions,” Journal of the American Chemical Society, vol. 90, no. 2, pp. 223–234, 1968.
[5]  E. Z. Hückel, “Quantentheoretische Beitr?ge zum Benzolproblem,” Zeitschrift für Physik, vol. 70, pp. 204–286, 1931.
[6]  W. Heitler and F. London, “Wechselwirkung neutraler Atome und hom?opolare Bindung nach der Quantenmechanik,” Zeitschrift für Physik, vol. 44, no. 6-7, pp. 455–472, 1927.
[7]  C. A. Coulson and H. C. Longuet-Higgins, “The electronic structure of conjugated systems. I. General theory,” Proceedings of the Royal Society of London Series A: Mathematical and Physical Sciences, vol. 191, no. 1024, pp. 39–60, 1947.
[8]  C. A. Coulson and H. C. Longuet-Higgins, “The electronic structure of conjugated systems. II. Unsaturated hydrocarbons and their hetero-derivatives,” Proceedings of the Royal Society of London. Series A. Mathematical and Physical Sciences, vol. 192, no. 1028, pp. 16–32, 1947.
[9]  C. A. Coulson and H. C. Longuet-Higgins, “The electronic structure of conjugated systems. III. Bond orders in unsaturated molecules; IV. Force constants and interaction constants in unsaturated hydrocarbons,” Proceedings of the Royal Society of London A, vol. 193, no. 1035, pp. 447–464, 1948.
[10]  R. G. Parr and P. K. Chattaraj, “Principle of maximum hardness,” Journal of the American Chemical Society, vol. 113, no. 5, pp. 1854–1855, 1991.
[11]  R. G. Parr and R. G. Pearson, “Absolute hardness: companion parameter to absolute electronegativity,” Journal of the American Chemical Society, vol. 105, no. 26, pp. 7512–7516, 1983.
[12]  R. G. Parr, L. V. Szentpály, and S. Liu, “Electrophilicity index,” Journal of the American Chemical Society, vol. 121, no. 9, pp. 1922–1924, 1999.
[13]  R. G. Parr and W. Yang, “Density functional approach to the frontier-electron theory of chemical reactivity,” Journal of the American Chemical Society, vol. 106, no. 14, pp. 4049–4050, 1984.
[14]  J. P. Perdew, R. G. Parr, M. Levy, and J. L. Balduz Jr., “Density-functional theory for fractional particle number: derivative discontinuities of the energy,” Physical Review Letters, vol. 49, no. 23, pp. 1691–1694, 1982.
[15]  R. G. Pearson, “Hard and soft acids and bases,” Journal of the American Chemical Society, vol. 85, no. 22, pp. 3533–3539, 1963.
[16]  R. G. Pearson, “Absolute electronegativity and hardness: application to inorganic chemistry,” Inorganic Chemistry, vol. 27, no. 4, pp. 734–740, 1988.
[17]  R. G. Pearson, “The principle of maximum hardness,” Accounts of Chemical Research, vol. 26, no. 5, pp. 250–255, 1993.
[18]  R. G. Pearson, “Absolute electronegativity and hardness: applications to organic chemistry,” Journal of Organic Chemistry, vol. 54, no. 6, pp. 1423–1430, 1989.
[19]  R. G. Pearson and J. Songstad, “Application of the principle of hard and soft acids and bases to organic chemistry,” Journal of the American Chemical Society, vol. 89, no. 8, pp. 1827–1836, 1967.
[20]  P. K. Chattaraj, H. Lee, and R. G. Parr, “HSAB principle,” Journal of the American Chemical Society, vol. 113, no. 5, pp. 1855–1856, 1991.
[21]  P. W. Ayers and M. Levy, “Perspective on “Density functional approach to the frontier-electron theory of chemical reactivity”,” Theoretical Chemistry Accounts, vol. 103, no. 3-4, pp. 353–360, 2000.
[22]  P. W. Ayers and R. G. Parr, “Variational principles for describing chemical reactions: the Fukui function and chemical hardness revisited,” Journal of the American Chemical Society, vol. 122, no. 9, pp. 2010–2018, 2000.
[23]  M. Gonzalez-Suarez, A. Aizman, and R. Contreras, “Phenomenological chemical reactivity theory for mobile electrons,” Theoretical Chemistry Accounts, vol. 126, no. 1, pp. 45–54, 2010.
[24]  M. Gonzalez-Suarez, A. Aizman, J. Soto-Delgado, and R. Contreras, “Bond Fukui functions as descriptor of the electron density reorganization in π conjugated systems,” The Journal of Organic Chemistry, vol. 77, no. 1, pp. 90–95, 2011.
[25]  R. Ormazábal-Toledo, P. R. Campodónico, and R. Contreras, “Are electrophilicity and electrofugality related concepts? A density functional theory study,” Organic Letters, vol. 13, pp. 822–824, 2011.
[26]  P. R. Campodónico, A. Aizman, and R. Contreras, “Electrophilicity of quinones and its relationship with hydride affinity,” Chemical Physics Letters, vol. 471, no. 1–3, pp. 168–173, 2009.
[27]  R. G. Parr and W. Yang, Density-Functional Theory of Atoms and Molecules, Oxford University Press, New York, NY, USA, 1989.
[28]  T. Koopmans, “über die Zuordnung von Wellenfunktionen und Eigenwerten zu den Einzelnen Elektronen Eines Atoms,” Physica, vol. 1, no. 1–6, pp. 104–113, 1934.
[29]  R. Vianello, N. Peran, and Z. B. Maksi?, “Hydride affinities of some substituted alkynes: prediction by DFT calculations and rationalization by triadic formula,” Journal of Physical Chemistry A, vol. 110, no. 47, pp. 12870–12881, 2006.
[30]  R. Contreras, J. Andres, V. S. Safont, P. Campodonico, and J. G. Santos, “A theoretical study on the relationship between nucleophilicity and ionization potentials in solution phase,” Journal of Physical Chemistry A, vol. 107, no. 29, pp. 5588–5593, 2003.
[31]  P. R. Campodónico, R. Ormazábal-Toledo, A. Aizman, and R. Contreras, “Permanent group effect on nucleofugality in aryl benzoates,” Chemical Physics Letters, vol. 498, pp. 221–225, 2010.
[32]  I. Um, H. Han, J. Ahn, S. Kang, and E. Buncel, “Reinterpretation of curved hammett plots in reaction of nucleophiles with aryl benzoates: change in rate-determining step or mechanism versus ground-state stabilization,” Journal of Organic Chemistry, vol. 67, no. 24, pp. 8475–8480, 2002.
[33]  I. Um, S. Jeon, and J. Seok, “Aminolysis of 2,4-dinitrophenyl X-substituted benzoates and Y-substituted phenyl benzoates in MeCN: Effect of the reaction medium on rate and mechanism,” Chemistry, vol. 12, no. 4, pp. 1237–1243, 2006.
[34]  I. H. Um, J. Y. Lee, M. Fujio, and Y. Tsuno, “Structure-reactivity correlations in nucleophilic substitution reactions of Y-substituted phenyl X-substituted benzoates with anionic and neutral nucleophiles,” Organic & Biomolecular Chemistry, vol. 4, no. 15, pp. 2979–2985, 2006.
[35]  R. Contreras, J. Andrés, P. Pérez, A. Aizman, and O. Tapia, “Theory of non-local (pair site) reactivity from model static-density response functions,” Theoretical Chemistry Accounts, vol. 99, pp. 183–191, 1998.
[36]  W. Yang and W. J. Mortier, “The use of global and local molecular parameters for the analysis of the gas-phase basicity of amines,” Journal of the American Chemical Society, vol. 108, no. 19, pp. 5708–5711, 1986.
[37]  R. R. Contreras, P. Fuentealba, M. Galván, and P. Pérez, “A direct evaluation of regional Fukui functions in molecules,” Chemical Physics Letters, vol. 304, no. 5-6, pp. 405–413, 1999.
[38]  P. Fuentealba, P. Pérez, and R. Contreras, “On the condensed Fukui function,” Journal of Chemical Physics, vol. 113, no. 7, pp. 2544–2551, 2000.
[39]  P. Bultinck, C. Cardenas, P. Fuentealba, P. A. Johnson, and P. W. Ayers, “How to compute the Fukui matrix and function for systems with (Quasi-)degenerate states,” Journal of Chemical Theory and Computation, vol. 10, no. 1, pp. 202–210, 2014.
[40]  A. T. Maynard, M. Huang, W. G. Rice, and D. G. Covell, “Reactivity of the HIV-1 nucleocapsid protein p7 zinc finger domains from the perspective of density-functional theory,” Proceedings of the National Academy of Sciences of the United States of America, vol. 95, no. 20, pp. 11578–11583, 1998.
[41]  P. Pérez, A. Aizman, and R. Contreras, “Comparison between experimental and theoretical scales of electrophilicity based on reactivity indexes,” Journal of Physical Chemistry A, vol. 106, no. 15, pp. 3964–3966, 2002.
[42]  P. Pérez, L. R. Domingo, M. J. Aurell, and R. Contreras, “Quantitative characterization of the global electrophilicity pattern of some reagents involved in 1,3-dipolar cycloaddition reactions,” Tetrahedron, vol. 59, no. 17, pp. 3117–3125, 2003.
[43]  P. R. Campodónico, J. Andrés, A. Aizman, and R. Contreras, “Nucleofugality index in α-elimination reactions,” Chemical Physics Letters, vol. 439, pp. 177–182, 2007.
[44]  P. R. Campodónico, P. Fuentealba, E. A. Castro, J. G. Santos, and R. Contreras, “Relationships between the electrophilicity index and experimental rate coefficients for the aminolysis of thiolcarbonates and dithiocarbonates,” Journal of Organic Chemistry, vol. 70, no. 5, pp. 1754–1760, 2005.
[45]  R. Ormazábal-Toledo, R. Contreras, and P. R. Campodónico, “Reactivity indices profile: a companion tool of the potential energy surface for the analysis of reaction mechanisms. Nucleophilic aromatic substitution reactions as test sase,” Journal of Organic Chemistry, vol. 78, pp. 1091–1097, 2013.
[46]  P. Pérez and R. Contreras, “A theoretical analysis of the gas-phase protonation of hydroxylamine, methyl-derivatives and aliphatic amino acids,” Chemical Physics Letters, vol. 293, no. 3-4, pp. 239–244, 1998.
[47]  K. Neimann and R. Neumann, “Electrophilic activation of hydrogen peroxide: Selective oxidation reactions in perfluorinated alcohol solvents,” Organic Letters, vol. 2, no. 18, pp. 2861–2863, 2000.
[48]  R. Das, J. L. Vigneresse, and P. K. Chattaraj, “Redox and Lewis acid–base activities through an electronegativity-hardness landscape diagram,” Journal of Molecular Modeling, vol. 19, no. 11, pp. 4857–4864, 2013.
[49]  W. Zhang, Y. Zhu, D. Wei, Y. Li, and M. Tang, “Theoretical investigations toward the [4 + 2] cycloaddition of ketenes with N-benzoyldiazenes catalyzed by N-heterocyclic carbenes: mechanism and enantioselectivity,” Journal of Organic Chemistry, vol. 77, no. 23, pp. 10729–10737, 2012.
[50]  A. Cedillo, R. Contreras, M. Galván, A. Aizman, J. Andrés, and V. S. Safont, “Nucleophilicity index from perturbed electrostatic potentials,” Journal of Physical Chemistry A, vol. 111, no. 12, pp. 2442–2447, 2007.
[51]  P. Jaramillo, P. Pérez, R. Contreras, W. Tiznado, and P. Fuentealba, “Definition of a nucleophilicity scale,” Journal of Physical Chemistry A, vol. 110, no. 26, pp. 8181–8187, 2006.
[52]  P. Campodonico, J. G. Santos, J. Andres, and R. Contreras, “Relationship between nucleophilicity/electrophilcity indices and reaction mechanisms for the nucleophilic substitution reactions of carbonyl compounds,” Journal of Physical Organic Chemistry, vol. 17, no. 4, pp. 273–281, 2004.
[53]  P. R. Campodonico, A. Aizman, and R. Contreras, “Group electrophilicity as a model of nucleofugality in nucleophilic substitution reactions,” Chemical Physics Letters, vol. 422, no. 4–6, pp. 340–344, 2006.
[54]  S. Jorge, A. Aizman, R. Contreras, and L. R. Domingo, “On the catalytic effect of water in the intramolecular diels-alder reaction of quinone systems: a theoretical study,” Molecules, vol. 17, no. 11, pp. 13687–13703, 2012.
[55]  R. Ormazábal-Toledo, R. Contreras, R. A. Tapia, and P. R. Campodónico, “Specific nucleophile-electrophile interactions in nucleophilic aromatic substitutions,” Organic and Biomolecular Chemistry, vol. 11, pp. 2302–2309, 2013.
[56]  A. Aizman, R. Contreras, and P. Pérez, “Relationship between local electrophilicity and rate coefficients for the hydrolysis of carbenium ions,” Tetrahedron, vol. 61, no. 4, pp. 889–895, 2005.
[57]  R. Contreras, J. Andrés, L. R. Domingo, R. Castillo, and P. Pérez, “Effect of electron-withdrawing substituents on the electrophilicity of carbonyl carbons,” Tetrahedron, vol. 61, pp. 417–422, 2005.
[58]  S. Karabunarliev, O. G. Mekenyan, W. Karcher, C. L. Russom, and S. P. Bradbury, “Quantum-chemical descriptors for estimating the acute toxicity of electrophiles to the bathed minnow (Pimephales promelas): an analysis based on molecular mechanisms,” Quantitative Structure-Activity Relationships, vol. 15, no. 4, pp. 302–310, 1996.
[59]  P. J. O'Brien, “Molecular mechanisms of quinone cytotoxicity,” Chemico-Biological Interactions, vol. 80, no. 1, pp. 1–41, 1991.
[60]  D. R. Roy, R. Parthasarathi, B. Maiti, V. Subramanian, and P. K. Chattaraj, “Electrophilicity as a possible descriptor for toxicity prediction,” Bioorganic and Medicinal Chemistry, vol. 13, no. 10, pp. 3405–3412, 2005.
[61]  P. K. Chattaraj, U. Sarkar, and D. R. Roy, “Electrophilicity index,” Chemical Reviews, vol. 106, no. 6, pp. 2065–2091, 2006.
[62]  P. Thanikaivelan, V. Subramanian, J. Raghava Rao, and B. Unni Nair, “Application of quantum chemical descriptor in quantitative structure activity and structure property relationship,” Chemical Physics Letters, vol. 323, no. 1-2, pp. 59–70, 2000.
[63]  R. Parthasarathi, V. Subramanian, D. R. Roy, and P. K. Chattaraj, “Electrophilicity index as a possible descriptor of biological activity,” Bioorganic & Medicinal Chemistry, vol. 12, no. 21, pp. 5533–5543, 2004.
[64]  A. Cerda-Monje, A. Aizman, R. A. Tapia, C. Chiappe, and R. Contreras, “Solvent effects in ionic liquids: empirical linear energy-density relationships,” Physical Chemistry Chemical Physics, vol. 14, no. 28, pp. 10041–10049, 2012.
[65]  K. B. Yatsimirskii, “Hydride affinity as a measure of acidity (electrophilicity),” Theoretical and Experimental Chemistry, vol. 17, no. 1, pp. 75–79, 1981.
[66]  X. Zhu, C. Wang, H. Liang, and J. Cheng, “Theoretical prediction of the hydride affinities of various p- and o-quinones in DMSO,” Journal of Organic Chemistry, vol. 72, no. 3, pp. 945–956, 2007.
[67]  E. Buncel and I. Um, “The α-effect and its modulation by solvent,” Tetrahedron, vol. 60, no. 36, pp. 7801–7825, 2004.
[68]  N. J. Fina and J. O. Edwards, “The alpha effect. A review,” International Journal of Chemical Kinetics, vol. 5, no. 1, pp. 1–26, 1973.
[69]  L. R. Domingo, M. J. Aurell, P. Pérez, and R. Contreras, “Quantitative characterization of the global electrophilicity power of common diene/dienophile pairs in Diels–Alder reactions,” Tetrahedron, vol. 58, no. 22, pp. 4417–4423, 2002.
[70]  L. R. Domingo, M. J. Aurell, P. Pérez, and R. Contreras, “Quantitative characterization of the local electrophilicity of organic molecules. Understanding the regioselectivity on Diels-Alder reactions,” The Journal of Physical Chemistry A, vol. 106, no. 29, pp. 6871–6875, 2002.
[71]  A. C. Legon, “Quantitative gas-phase electrophilicities of the dihalogen molecules XY = F2, Cl2, Br2, BrCl and ClF,” Chemical Communications, no. 23, pp. 2585–2586, 1998.
[72]  A. C. Legon and D. J. Millen, “Hydrogen bonding as a probe of electron densities: limiting gas-phase nucleophilicities and electrophilicities of B and HX,” Journal of the American Chemical Society, vol. 109, no. 2, pp. 356–358, 1987.
[73]  M. Baidya, S. Kobayashi, and H. Mayr, “Nucleophilicity and nucleofugality of phenylsulfinate (PhSO2-): A key to understanding its ambident reactivity,” Journal of the American Chemical Society, vol. 132, no. 13, pp. 4796–4805, 2010.
[74]  J. Soto-Delgado, A. Aizman, L. R. Domingo, and R. Contreras, “Invariance of electrophilicity of independent fragments. Application to intramolecular Diels-Alder reactions,” Chemical Physics Letters, vol. 499, no. 4–6, pp. 272–277, 2010.
[75]  J. Soto-Delgado, L. R. Domingo, and R. Contreras, “Quantitative characterization of group electrophilicity and nucleophilicity for intramolecular Diels-Alder reactions,” Organic and Biomolecular Chemistry, vol. 8, no. 16, pp. 3678–3683, 2010.
[76]  T. J. Heckrodt and J. Mulzer, “Total synthesis of elisabethin A: Intramolecular Diels-Alder reaction under biomimetic conditions,” Journal of the American Chemical Society, vol. 125, no. 16, pp. 4680–4681, 2003.
[77]  B. G. Cox and A. J. Parker, “Solvation of ions. XVIII. Protic-dipolar aprotic solvent effects on the free energies, enthalpies, and entropies of activation of an SNAr reaction,” Journal of the American Chemical Society, vol. 95, no. 2, pp. 408–410, 1973.
[78]  I. Um, L. Im, J. Kang, S. S. Bursey, and J. M. Dust, “Mechanistic assessment of S NAr displacement of halides from 1-Halo-2,4-dinitrobenzenes by selected primary and secondary amines: Br?nsted and Mayr analyses,” Journal of Organic Chemistry, vol. 77, no. 21, pp. 9738–9746, 2012.

Full-Text

comments powered by Disqus

Contact Us

service@oalib.com

QQ:3279437679

WhatsApp +8615387084133