Buscar en
Educación Química
Toda la web
Inicio Educación Química The alkyl group is a –I+R substituent
Información de la revista
Vol. 28. Núm. 4.
Páginas 232-237 (Octubre 2017)
Compartir
Compartir
Descargar PDF
Más opciones de artículo
Visitas
13655
Vol. 28. Núm. 4.
Páginas 232-237 (Octubre 2017)
Reflection
Open Access
The alkyl group is a –I+R substituent
El grupo alquilo es un sustituyente –I+R
Visitas
13655
Luis Salvatella
Instituto de Síntesis Química y Catálisis Homogénea-ISQCH, CSIC – Universidad de Zaragoza, Pedro Cerbuna 12, E-50009 Zaragoza, Spain
Este artículo ha recibido

Under a Creative Commons license
Información del artículo
Resumen
Texto completo
Bibliografía
Descargar PDF
Estadísticas
Figuras (4)
Mostrar másMostrar menos
Abstract

Electronic substituent effects are usually classified as inductive (through σ-bonds) and resonance effects (via π-bonds). The alkyl group has been usually regarded as a σ-electron donor substituent (+I effect, according to the Ingold's classification). However, a σ-withdrawing, π-donor effect (–I+R pattern) allows explaining the actual electron-withdrawing behavior of alkyl groups when bound to sp3 carbon atoms as well as their well-known electron-releasing properties when attached to sp2 or sp atoms. Alkyl substitution effects on several molecular properties (dipole moments, NMR, IR, and UV spectra, reactivity in gas phase and solution) are discussed.

Keywords:
Second-year undergraduate
Organic chemistry
Misconceptions
Textbooks
Acids/bases
Covalent bonding
IR spectroscopy
NMR spectroscopy
UV–Vis spectroscopy
Resumen

Los efectos electrónicos del sustituyente se clasifican habitualmente como inductivos (a través de enlaces σ) o de resonancia (mediante enlaces π). El grupo alquilo ha sido considerado habitualmente como un sustituyente dador de densidad electrónica σ (+I, según la clasificación de Ingold). Sin embargo, un patrón σ-aceptor π-dador (–I+R) permite explicar el comportamiento real de los grupos alquilo como atractores de electrones cuando están unidos a átomos de carbono sp3, así como sus conocidas propiedades dadoras de electrones cuando están unidos a átomos sp2 o sp. Se discuten los efectos de sustitución del grupo alquilo en varias propiedades moleculares (momentos dipolares, espectros de RMN, IR y UV, reactividad en fase gas y disolución).

Palabras clave:
Segundo curso de grado
Química orgánica
Errores conceptuales
Libros de texto
Ácidos/bases
Enlace covalente
Espectroscopía IR
Espectroscopía RMN
Espectroscopía UV-Vis
Texto completo
Introduction

Substituent effects constitute a key concept for the comprehension of reactivity and spectroscopic behavior of organic compounds (Krygowski & St¿pień, 2005). In a simple approach, substituent effects can be classified according to the mechanism of interaction with the reactive center as inductive (through σ-bonds) or resonance effects (through π-bonds). Nevertheless, some further terms (such as steric, field or solvent effects) would be required for a thorough description of substituent effects.

Since the Ingold's classification of electronic substituent effects (Ingold, 1953), the alkyl group has been regarded as a σ-donor substituent (+I, in the Ingold's nomenclature) in most Organic Chemistry textbooks (Burrows, Holman, Parsons, Pilling, & Price, 2013; Hornback, 2006; Roos & Roos, 2014; Smith, 2013; Vollhardt & Schore, 2014). Nevertheless, the Eğe's criticisms to such a simplistic viewpoint should be remarked:

“In water, propanoic acid is slightly weaker than acetic acid. The nature of the inductive effect of an alkyl group is debated by chemists. Alkyl groups stabilize carbocations and in that role appear to be electron-releasing. They also increase the basicity of amines, again suggesting that they are electron-releasing. On the other hand, though tert-butyl alcohol (pKa 19) is a weaker acid than ethanol (pKa 17) in water, it is stronger acid in the gas phase. This experimental observation suggests than alkyl groups can stabilize anions as well as cations and that solvation plays an important role in determining relative acidities. Thus a word of caution is necessary. The relative acidities on which the generalizations presented in this chapter are based were determined in water. In the gas phase, reversals in the order of related compounds are often seen.” (Eğe, 1999, p. 107)

Surprisingly, the accumulation of evidences against the assumed +I feature of the alkyl group (Böhm & Exner, 2004; Laurie & Muenter, 1966; Minot, Eisenstein, Hiberty, & Anh, 1980; Sebastian, 1971; Tasi, Mizukami, & Pálinkó, 1997) has shown very little effect on a so widespread view.

Some alkyl substitution effects have been often explained in textbooks in contradictory or enigmatic ways. Thus, chemical shift differences between CH3 and CH2 groups are attributed in the Hornback's book to the fact that “carbon is slightly more electronegative than hydrogen” (Hornback, 2006, p. 549) despite the alkyl group has been previously classified as a weak inductive electron-donating substituent (Hornback, 2006, p. 117). In the Vollhardt's textbook, the relationship between methyl group chemical shifts for a number of CH3X compounds and the X electronegativity is illustrated by a table lacking an entry for X=methyl (Vollhardt & Schore, 2014, p. 389), thus avoiding the inconvenient carbon issue.

I show here that the alkyl group behaves as a –I+R substituent. Although some factors (such as field, steric or solvent effects) are implicitly ignored in this approach, a lot of currently available theoretical and experimental evidences can thus be described in an easy way.

A Cδ––Hδ+ bond polarization has been experimentally observed for methane (Lazzeretti, Zanasi, & Raynes, 1987), consistently with the larger electronegativity of carbon relative to hydrogen, 2.55 vs. 2.20 in the Pauling scale (Allred, 1961). Such a polarization pattern allows predicting the dipole moment direction of simple hydrocarbons through additive models, though quantitative agreement is usually modest (2-methylpropane: 0.3 D estimated vs. 0.132 D experimental) (Dean, 1999).

Since hydrogen is used as the standard in the Ingold's classification of substituents (Krygowski & St¿pień, 2005), the alkyl group should be classified as a –I substituent (hence a σ electron-withdrawing group). Such a role is illustrated in Fig. 1 for the σ bond polarization from whatever atom Y to a methyl group, though the reverse bond polarization is expected when Y is a more electronegative than carbon (e.g., chlorine).

Figure 1.

–I (left) and +R (right) effects of a methyl group bound to an atom Y.

(0,03MB).

A different behavior is found for alkyl groups when attached to sp2 or sp-hybridized atoms due to electron density donation from alkyl C–H or C–C σ bonds to the empty p orbital of the contiguous atom (the simplest π-system), as shown in Fig. 1. Thus, the decrease of gas phase acidity for phenol and benzoic acid through p-methyl substitution (McMahon & Kebarle, 1977) can only be attributed to a significant π-donor effect for methyl substituent (indeed, larger than that for methoxy group). However, the alkyl group should be considered as an atypical π-donor substituent due to the lack of lone electron pairs. Such a σ-bond/π-system interaction, named as hyperconjugation (Mullins, 2012) can readily be explained by analogy with the π-donor behavior of a lone pair-bearing atom (e.g., chlorine) to an empty p orbital, though C–C or C–H bonds (rather than electron lone pairs) of the alkyl group are involved as electron-releasing units in hyperconjugative interactions. Interestingly, πσ* interactions (negative hyperconjugation) are usually negligible for alkyl groups lacking electronegative atoms (Bocca, Pontes, & Basso, 2004).

Some molecular structural features can be rationalized on the basis of the alkyl group properties. For example, the larger CO bond lengths found in methyl ketones (acetone: exp. 1.210Å, calc. 1.193Å) in comparison with the related aldehydes (acetaldehyde: exp. 1.209Å, calc. 1.188Å) (Berry, Waltman, Pacansky, & Hagler, 1995) can be attributed to the stabilization of the zwitterionic resonance form (see Fig. 2) through alkyl group π-donation to the carbonylic carbon atom, thus weakening the double-bond feature of the carbonyl group.

Figure 2.

Neutral (left) and zwitterionic (right) resonance forms of a carbonyl compound.

(0,06MB).

Hyperconjugative interactions are dependent on the arrangement of C–H (or C–C) bonds relative to the p orbital of the contiguous atom Y, the most effective interaction corresponding to a nearly parallel arrangement. For example, the toluene Csp3–H bond nearly perpendicular to the framework plane is slightly longer than the other Csp3–H bonds (by 0.002Å, Hameka & Jensen, 1996). The geometry dependence of hyperconjugation allows explaining the conformational analysis of methyl-substituted unsaturated compounds, such as propene (Liberles, O’Leary, Eilers, & Whitman, 1972) or acetaldehyde (Muñoz-Caro, Niño, & Moule, 1994).

As a well-known consequence of the π-donor behavior of the alkyl group, alkyl substitution yields more electron-rich alkenes and arenes (Libit & Hoffmann, 1974). The high reactivity of an alkyl-substituted arene in a SEAr reaction can thus be attributed to the stabilization of the corresponding Wheland intermediate through π-electron donation.

The –I+R behavior of the alkyl group allows explaining a number of features of alkyl-substituted compounds, such as dipole moments, spectroscopic properties and reactivity (in gas phase and solution media), as shown below.

Dipole moments

The electron-withdrawing behavior of alkyl group in aliphatic compounds is also reflected in dipole moments. Thus, the dipole moment vectors for propane and 2-methylpropane (Tasi et al., 1997), as well as some substituted bicyclo[2.2.2]octanes (Böhm & Exner, 2004) can be attributed to the withdrawing effect (–I) of the methyl group in comparison with hydrogen (see Fig. 3).

Figure 3.

Dipole moments of propane and substituted bicyclo[2.2.2]octanes.

(0,07MB).

In contrast, the π-donor character of methyl group (+R) is required in order to explain the raise of dipole moments of nitrobenzene and benzonitrile through p-methyl substitution (Brown, 1959) (see Fig. 4).

Figure 4.

Dipole moments of benzene derivatives.

(0,07MB).

Molecular dipole moments can be reliably calculated by current computational methods. Interestingly, the calculated dipole moment vectors for a set of simple hydrocarbons (Tasi et al., 1997) have allowed inferring a dual role for the methyl group: electron-withdrawing when attached to sp3 carbon atoms, but electron-donating when bound to sp2 or sp3 carbons.

Such a dual behavior of the alkyl substituent is also observed for heteroatom-bearing compounds. Thus, a gradual dipole moment decrease is observed for successive methyl substitution on ammonia (NH3, 1.47 D; MeNH2, 1.31 D; Me2NH, 1.01 D; Me3N, 0.61 D) (Le Fèvre & Russell, 1947), in agreement with the progressive diminution of the nitrogen electron density (Hehre & Pople, 1970). In contrast, a dipole moment enhancement (from 1.53 D to 1.68 D) (Nelson, Lide, & Maryott, 1967) is found for N,N-dimethyl substitution on aniline (Targema, Obi-Egbedi, & Adeoye, 2013), consistently with the raise of the π-donor character for the amino group (Hinchliffe & Kidd, 1980) due to +R contributions of methyl substituents.

Spectroscopic properties

Spectroscopic properties of many organic compounds can be easily rationalized by assuming a –I+R behavior for the alkyl group as a general feature. Thus, the NMR chemical shift of an atom can be regarded as an experimental measure of the electron density at the corresponding nucleus position though other effects – such as anisotropic magnetic fields – can also be involved. Downfield shifts induced by a methyl substituent on sp3 carbon atoms (+9.6ppm in 13C NMR) or the corresponding bound hydrogen atoms (+0.63ppm in 1H NMR) (Pretsch, Bühlmann, & Badertscher, 2009) are consistent with the behavior of typical –I groups (such as halogen atoms).

Alkyl substitution effects on NMR chemical shifts of alkenes show an electron density decrease in α position (+12.9ppm for 13C NMR; +0.45ppm for 1H NMR), as well as a density raise in β position (–7.4ppm for 13C; –0.31/–0.40ppm for 1H), consistently with a –I+R effect, though anisotropic effects (such as ring currents) may also play a role. Such a –I+R behavior is also found for alkynes, according to 13C NMR spectroscopy (+8.5ppm for α position, –3.6ppm for β position).

A –I+R effect is also found for carbonyl compounds. Thus, NMR data on methyl substitution show an electron density decrease on the carbon atom (13C NMR effects: formaldehyde, +3.5ppm; acetaldehyde, +6.2ppm; formic acid, +10.6ppm; methyl formate, +10.7ppm; N,N-dimethylformamide, +7.4ppm; 1H NMR effect on formaldehyde, +0.2ppm) (Pretsch et al., 2009), as well as a density raise on the oxygen (17O NMR effect for acetaldehyde: –33ppm) (Gerothanassis, 2010).

The dichotomous behavior of alkyl substituents on π-systems (electron density raise for α atom, electron density decrease for β atom) cannot be explained on the basis of a simple behavior (such as a +I effect).

A –I+R behavior (Meier, 2007) is observed through 15N NMR spectroscopy for alkyl substitution on amines and amides depending on the nitrogen hybridization (downfield shifts for aliphatic amines, upfield shifts for Nsp2-bearing compounds – such as anilines and amides).

NMR coupling constants are also dependent on substituent electronic properties (as well as some geometrical features). Thus, a significant decrease is found for 1H–1H coupling constants through methyl substitution (trans, –2.3Hz; cis, –1.6Hz; gem, –0.4Hz), in qualitative agreement with data from typical electron-withdrawing groups, such as the fluorine atom (trans, –6.3Hz; cis, –6.9Hz; gem, –5.7Hz). The positive contribution for methyl-substitution on 13C–1H coupling constants of aliphatic compounds (+1.0Hz), is also qualitatively consistent with those from other –I groups (fluorine, +24Hz).

Infrared spectroscopy is also sensitive to substituent properties, as illustrated by the CO stretching frequency of carbonyl compounds as a function of the corresponding substituent Y, which can be rationalized in terms of resonance forms (Fig. 2). By taking an aliphatic aldehyde (ca. 1725cm–1) as a reference, the redshift (wavenumber decrease) induced by a +I substituent (acetyltrimethylsilane, 1645cm–1: Soderquist & Hsu, 1982) can be attributed to the stabilization of the zwitterionic form. Instead, the blueshift provoked by a –I substituent (acyl chlorides, >1800cm–1: Pretsch et al., 2009) can be explained by means of two alternative or concurrent mechanisms (destabilization of the zwitterionic form and/or contribution of an acylium ion-bearing form). Finally, the redshifts provoked by +R substituents (amides, ca. 1680cm–1: Pretsch et al., 2009) can be attributed to the contribution of a specific resonance form. The slight redshift induced by alkyl group (methyl ketones, ca. 1715cm–1) shows a net electron-donating effect (hence, a predominance of the +R effect over –I properties). The net donor effect of the carbonyl-bound alkyl group is consistent with the larger dipole moment of acetone (2.88 D) relative to formaldehyde (2.33 D) (Nelson et al., 1967).

The alkyl group influence on UV–Vis spectra of many compounds can also be explained in terms of electronic effects. Thus, bathochromic shifts induced by alkyl groups on UV absorption bands of α,β-unsaturated compounds (+10nm in α position, +12nm in β position), conjugated polyenes (+5nm) or benzene derivatives (+3.0nm) are qualitatively consistent with effects of typical π-donor groups (e.g., chlorine).

Gas phase acid–base reactivity

Relative basicities of aliphatic amines in aqueous solution have been attributed to the assumed +I effect of alkyl group (Sorrell, 2006). Interestingly, the irregular basicity order of amines in water (Me2NH>MeNH2>Me3N>NH3, as shown by the pKa values for the corresponding conjugated acids: 10.77>10.62>9.80>9.246) (Dean, 1999) is contaminated by solvent effects as illustrated by the systematic basicity order of amines in gas phase (Me3N>Me2NH>MeNH2>NH3) (Brauman, Riveros, & Blair, 1971). Although the gas phase basicity order can be attributed to the usually assumed +I alkyl effect (Carter, 2007), a decrease of the nitrogen electron density through methyl substitution has been indeed observed by means of Molecular Electrostatic Potential calculations (Baeten, De Proft, & Geerlings, 1995), thus indicating a –I behavior for the methyl group. Actually, the gas phase basicity order of aliphatic amines should be attributed to the increasing stabilization of substituted ammonium ions due to the alkyl group polarizability (Aue, Webb, & Bowers, 1976).

Relative acidities of alcohols in aqueous solution (H2O>MeOH>EtOH>iPrOH>tBuOH) have also been attributed in some textbooks to the assumed alkyl +I effect (Johnson, 1999; Solomons, Fryhle, & Snyder, 2016). Since the reverse acidity order is found in gas phase, relative acidities of alcohols in water should be attributed to the lower magnitudes of solvation enthalpies for larger alkoxide anions (Brauman & Blair, 1969).

The discussion on alkyl group electronic properties can also be applied to carbanions. Thus, the ‘textbook’ stability order for simple carbanions (methyl>ethyl>isopropyl>tert-butyl) has been attributed to the assumed +I inductive effect of alkyl groups (Burrows et al., 2013; Chaloner, 2015; Roos & Roos, 2014; Smith, 2013). However, an irregular order is found for gas phase carbanion stabilities (tBu>Me>iPr>Et), in agreement with the concurrence of two opposed alkyl effects (DePuy et al., 1989): a stabilizing mechanism through alkyl polarizability (that is, n→σ* hyperconjugation) and a destabilizing trend (consistently with a +R role, by assuming a p-like behavior for the carbon lone pair).

The stability of other reaction intermediates can also be assessed on the basis of alkyl group effects. Thus, the well-known stability order for carbocations (tertiary>secondary>primary>methyl) has been sometimes attributed to a positive inductive effect (Chaloner, 2015; Roos & Roos, 2014). Interestingly, hyperconjugation is presented in many textbooks as an alternative explanation for the stability order of carbocations (Brown, Iverson, Anslyn, & Foote, 2013; Burrows et al., 2013) though the usual ambiguous writing prevents ascertaining whether both explanations correspond to either two different descriptions of the same phenomenon or two concurrent mechanisms playing in the same direction. Anyway, the stability order for carbocations should be attributed to hyperconjugation (hence, a+R behavior on a vacant p orbital, the simplest π system), though other interactions (such as alkyl polarizability) are also involved (Aue, 2011).

Free radicals show the same stability order as carbocations, thus indicating stabilization through alkyl substitution. Although such a stability order may be justified on the basis of an assumed +I behavior, the +R effect can be alternatively regarded, analogously to the stabilization of free radicals by lone pair-bearing atoms (Zipse, 2006).

Reactivity in solution

Relative acidities of simple carboxylic acids in aqueous solution (acetic acid>propionic acid>butyric acid) have been used in some textbooks to illustrate the assumed +I effect of the alkyl group (Sorrell, 2006). Interestingly, the reverse order is found when enthalpies are instead regarded (Christensen, Izatt, & Hansen, 1967), thus indicating that the acidity order in aqueous solution should be attributed to hydration entropies. Thus, the significant lattice order of liquid water (vaporization entropy equaling 118.89Jmol–1K–1, in contrast with typical values of ca. 88Jmol–1K–1 for most liquids, Dean, 1999) can introduce sizeable changes on the reaction energetics. In particular, hydration of apolar molecules (or moieties) leads to a further solvent lattice ordering (Blokzijl & Engberts, 1993). As a consequence, alkyl group inductive effects from experimental data in aqueous solution are often masked by hydration entropies (Calder & Barton, 1971). Relative acidities of simple carboxylic acids in gas phase (Yamdagni & Kebarle, 1973) and acetonitrile (Eckert et al., 2009) are consistent with the major role played by hydration entropies.

The lower acidity of pivalic acid in comparison with acetic acid, usually attributed to the assumed +I effect of the alkyl group (Smith, 2008), is reversed when reaction enthalpies are considered (Eckert et al., 2009).

The assumed +I alkyl group effect on the acidity of simple carboxylic acids in aqueous solution can thus be attributed to an artifact derived from solvent effects. Whereas a volume increase of neutral solutes leads to a hydration entropy raise, the reverse relationship is found for ionic species (Graziano, 2009). As a consequence, alkyl substitution (through an increase of the molecular volume) leads to the stabilization (in Gibbs free energy terms) of non-ionized acid in water as well as the destabilization of the corresponding carboxylate anion, thus resulting an acidity decrease.

The larger acidity of formic acid in comparison with acetic acid in aqueous solution (pKa values: 3.751 and 4.756, respectively, Dean, 1999) has also been discussed in many textbooks as an example of the application of inductive effects (Hart, Hadad, Craine, & Hart, 2012; Hornback, 2006; Okuyama & Maskill, 2014; Roos & Roos, 2014). Since very similar reaction enthalpies are involved in the dissociation reactions of formic and acetic acids (Christensen et al., 1967), the larger acidity of formic acid must be indeed attributed to hydration entropy differences.

Conclusions

A clear comprehension of inductive and resonance effects is a major key for a sound learning of Organic Chemistry (Mullins, 2008). Surprisingly, the almost ubiquitous alkyl group has been incorrectly presented in many textbooks as a σ-donor (+I) group. However, a dual behavior is shown by alkyl substituents depending on the hybridization of the neighbor atom. Thus, alkyl groups bound to aliphatic chains behave as σ-acceptors (–I, consistently with the larger electronegativity of carbon relative to hydrogen), whereas those attached to π-systems act as π-donors (+R, due to hyperconjugative interactions). A number of experimental and theoretical data (dipole moments, NMR, IR, and UV spectra, reactivity) agree with such a dual behavior.

The whole analysis of all data considered here allows inferring a small –I effect as well as a significant +R behavior for the alkyl group as a feature valid in all discussions on spectroscopic and reactivity properties of organic compounds.

Conflict of interests

The author declares no conflict of interest.

Acknowledgements

Financial support from the Ministerio de Economía y Competitividad (MINECO) of Spain (Project CTQ2014-52367-R), the Gobierno de Aragón (Consolidated Group E11) and the European Regional Development Fund (ERDF) is gratefully acknowledged.

References
[Allred, 1961]
A.L. Allred.
Electronegativity values from thermochemical data.
Journal of Inorganic and Nuclear Chemistry, 17 (1961), pp. 215-221
[Aue, 2011]
D.H. Aue.
Carbocations.
Wiley Interdisciplinary Reviews: Computational Molecular Science, 1 (2011), pp. 487-508
[Aue et al., 1976]
D.H. Aue, H.M. Webb, M.T. Bowers.
Quantitative proton affinities, ionization potentials, and hydrogen affinities of alkylamines.
Journal of the American Chemical Society, 98 (1976), pp. 311-317
[Baeten et al., 1995]
A. Baeten, F. De Proft, P. Geerlings.
Basicity of primary amines: A group properties based study of the importance of inductive (electronegativity and softness) and resonance effects.
Chemical Physics Letters, 235 (1995), pp. 17-21
[Berry et al., 1995]
R.J. Berry, R.J. Waltman, J. Pacansky, A.T. Hagler.
Ab initio investigation of the conformational energies, rotational barriers, molecular structures, vibrational frequencies, and dipole moments of aldehydes and ketones.
Journal of Physical Chemistry, 99 (1995), pp. 10511-10520
[Blokzijl and Engberts, 1993]
W. Blokzijl, J.B.F.N. Engberts.
Hydrophobic effects. Opinions and facts.
Angewandte Chemie International Edition in English, 32 (1993), pp. 1545-1579
[Bocca et al., 2004]
C.C. Bocca, R.M. Pontes, E.A. Basso.
Implications of hyperconjugative effects on bond lengths of allylic systems: An NBO investigation.
Journal of Molecular Structure (THEOCHEM), 710 (2004), pp. 105-110
[Böhm and Exner, 2004]
S. Böhm, O. Exner.
Prediction of molecular dipole moments from bond moments: Testing of the method by DFT calculations on isolated molecules.
Physical Chemistry Chemical Physics, 6 (2004), pp. 510-514
[Brauman and Blair, 1969]
J.I. Brauman, L.K. Blair.
Gas-phase acidities of amines.
Journal of the American Chemical Society, 91 (1969), pp. 2126-2127
[Brauman et al., 1971]
J.I. Brauman, J.M. Riveros, L.K. Blair.
Gas-phase basicities of amines.
Journal of the American Chemical Society, 93 (1971), pp. 3914-3916
[Brown, 1959]
T.L. Brown.
The electronic properties of alkyl groups. II. The dipole moments of alkyl benzenes and derivatives.
Journal of the American Chemical Society, 81 (1959), pp. 3232-3235
[Brown et al., 2013]
W. Brown, B. Iverson, E. Anslyn, C. Foote.
Organic chemistry.
Cengage Learning, (2013),
[Burrows et al., 2013]
A. Burrows, J. Holman, A. Parsons, G. Pilling, G. Price.
Chemistry3: Introducing inorganic, organic and physical chemistry.
2nd ed., Oxford University Press, (2013),
[Calder and Barton, 1971]
G.V. Calder, T.J. Barton.
Actual effects controlling the acidity of carboxylic acids.
Journal of Chemical Education, 48 (1971), pp. 338-340
[Carter, 2007]
K. Carter.
Organic chemistry.
Global Media, (2007),
[Chaloner, 2015]
P. Chaloner.
Organic chemistry: A mechanistic approach.
CRC Press, (2015),
[Christensen et al., 1967]
J.J. Christensen, R.M. Izatt, L.D. Hansen.
Thermodynamics of proton ionization in dilute aqueous solution. VII. ΔH° and ΔS° values for proton ionization from carboxylic acids at 25°.
Journal of the American Chemical Society, 89 (1967), pp. 213-222
[Dean, 1999]
J.A. Dean.
Lange's handbook on chemistry.
15th ed., McGraw-Hill, (1999),
[DePuy et al., 1989]
C.H. DePuy, S. Gronert, S.E. Barlow, V.M. Bierbaum, R. Damrauer.
The Gas-Phase Acidities of the Alkanes.
Journal of the American Chemical Society, 111 (1989), pp. 1968-1973
[Eckert et al., 2009]
F. Eckert, I. Leito, I. Kaljurand, A. Kütt, A. Klamt, M. Diedenhofen.
Prediction of acidity in acetonitrile solution with COSMO-RS.
Journal of Computational Chemistry, 30 (2009), pp. 799-810
[Eğe, 1999]
S.N. Eğe.
Organic chemistry: Structure and reactivity.
4th ed., Houghton Mifflin College Div, (1999),
[Gerothanassis, 2010]
I.P. Gerothanassis.
Oxygen-17 NMR spectroscopy: Basic principles and applications (Part I).
Progress in Nuclear Magnetic Resonance Spectroscopy, 56 (2010), pp. 95-197
[Graziano, 2009]
G. Graziano.
Hydration entropy of polar nonpolar and charged species.
Chemical Physics Letters, 479 (2009), pp. 56-59
[Hameka and Jensen, 1996]
H.F. Hameka, J.O. Jensen.
Theoretical studies of the methyl rotational barrier in toluene.
Journal of Molecular Structure (THEOCHEM), 362 (1996), pp. 325-330
[Hart et al., 2012]
H. Hart, C.M. Hadad, L. Craine, D.J. Hart.
Organic chemistry: A short course.
Brooks/Cole Cengage Learning, (2012),
[Hehre and Pople, 1970]
W.J. Hehre, J.A. Pople.
The methyl inductive effect on acid-base strengths.
Tetrahedron Letters, 11 (1970), pp. 2959-2962
[Hinchliffe and Kidd, 1980]
A. Hinchliffe, I.F. Kidd.
CH bond dipole.
Journal of the Chemical Society Faraday Transactions 2: Molecular and Chemical Physics, 76 (1980), pp. 172-176
[Hornback, 2006]
J. Hornback.
Organic chemistry.
2nd ed., Thomson Brooks/Cole, (2006),
[Ingold, 1953]
C.K. Ingold.
Structure and mechanism in organic chemistry.
Cornell University Press, (1953),
[Johnson, 1999]
A.W. Johnson.
Invitation to organic chemistry.
Jones & Bartlett Learning, (1999),
[Krygowski and St¿pień, 2005]
T.M. Krygowski, B.T. St¿pień.
Sigma- and pi-electron delocalization: Focus on substituent effects.
Chemical Reviews, 105 (2005), pp. 3482-3512
[Laurie and Muenter, 1966]
V.W. Laurie, J.S. Muenter.
On the inductive effect of methyl groups bonded to saturated systems.
Journal of the American Chemical Society, 88 (1966), pp. 2883-2884
[Lazzeretti et al., 1987]
P. Lazzeretti, R. Zanasi, W.T. Raynes.
On the CH bond dipole moment in alkanes.
Journal of Chemical Physics, 87 (1987), pp. 1681-1684
[Le Fèvre and Russell, 1947]
R.J.W. Le Fèvre, P. Russell.
The dependence on state of the apparent dipole moments of ammonia, methylamine dimethylamine, and trimethylamine.
Transactions of the Faraday Society, 43 (1947), pp. 374-393
[Liberles et al., 1972]
A. Liberles, B. O’Leary, J.E. Eilers, D.R. Whitman.
Methyl rotation barriers and hyperconjugation.
Journal of the American Chemical Society, 94 (1972), pp. 6894-6898
[Libit and Hoffmann, 1974]
L. Libit, R. Hoffmann.
Detailed orbital theory of substituent effects. Charge transfer, polarization, and the methyl group.
Journal of the American Chemical Society, 96 (1974), pp. 1370-1383
[McMahon and Kebarle, 1977]
T.B. McMahon, P. Kebarle.
Intrinsic acidities of substituted phenols and benzoic acids determined by gas-phase proton-transfer equilibria.
Journal of the American Chemical Society, 99 (1977), pp. 2222-2230
[Meier, 2007]
H. Meier.
Nuclear magnetic resonance spectroscopy.
Spectroscopic methods in organic chemistry, 2nd ed., pp. 74-241
[Minot et al., 1980]
C. Minot, O. Eisenstein, P.C. Hiberty, N.T. Anh.
Non-equivalence of the various criteria for alkyl inductive effect.
Bulletin de la Société Chimique de France II, 47 (1980), pp. 119-124
[Mullins, 2008]
J.J. Mullins.
Six pillars of organic chemistry.
Journal of Chemical Education, 85 (2008), pp. 83-87
[Mullins, 2012]
J.J. Mullins.
Hyperconjugation: A more coherent approach.
Journal of Chemical Education, 89 (2012), pp. 834-836
[Muñoz-Caro et al., 1994]
C. Muñoz-Caro, A. Niño, D.C. Moule.
On the origin of the barriers and the structures of acetaldehyde in its ground and first singlet excited state.
Theoretica Chimica Acta, 88 (1994), pp. 299-310
[Nelson et al., 1967]
R.D.J. Nelson, D.R. Lide, A.A. Maryott.
Selected values of electric dipole moments for molecules in the gas phase.
U.S. National Bureau of Standards, (1967),
[Okuyama and Maskill, 2014]
T. Okuyama, H. Maskill.
Organic chemistry: A mechanistic approach.
Oxford University Press, (2014),
[Pretsch et al., 2009]
E. Pretsch, P. Bühlmann, M. Badertscher.
Structure determination of organic compounds: Tables of spectral data.
4th ed., Springer, (2009),
[Roos and Roos, 2014]
G. Roos, C. Roos.
Organic chemistry concepts: An EFL approach.
Academic Press, (2014),
[Sebastian, 1971]
J.F. Sebastian.
The electronic effects of alkyl groups.
Journal of Chemical Education, 48 (1971), pp. 97-98
[Smith, 2008]
J.G. Smith.
Organic chemistry.
3rd ed., McGraw-Hill, (2008),
[Smith, 2013]
M.B. Smith.
March's advanced organic chemistry: Reactions, mechanisms and structure.
7th ed., Wiley, (2013),
[Soderquist and Hsu, 1982]
J.A. Soderquist, G.J.-H. Hsu.
Pure unsolvated (α-methoxyvinyl)lithium and related acyl anion equivalents via the transmetalation of organotin compounds.
Organometallics, 1 (1982), pp. 830-833
[Solomons et al., 2016]
T.W.G. Solomons, C.G. Fryhle, S.A. Snyder.
Organic chemistry.
12th ed., Wiley, (2016),
[Sorrell, 2006]
T.N. Sorrell.
Organic chemistry.
University Science Books, (2006),
[Targema et al., 2013]
M. Targema, N.O. Obi-Egbedi, M.D. Adeoye.
Molecular structure and solvent effects on the dipole moments and polarizabilities of some aniline derivatives.
Computational & Theoretical Chemistry, 1012 (2013), pp. 47-53
[Tasi et al., 1997]
G. Tasi, F. Mizukami, I. Pálinkó.
Analysis of permanent electric dipole moments of aliphatic hydrocarbon molecules.
Journal of Molecular Structure (THEOCHEM), 401 (1997), pp. 21-27
[Vollhardt and Schore, 2014]
P. Vollhardt, N. Schore.
Organic chemistry: Structure and function.
7th ed., Freeman and Company, (2014),
[Yamdagni and Kebarle, 1973]
R. Yamdagni, P. Kebarle.
Intrinsic acidities of carboxylic acids from gas-phase acid equilibria.
Journal of the American Chemical Society, 95 (1973), pp. 4050-4052
[Zipse, 2006]
H. Zipse.
Radical stability – A theoretical perspective.
Topics in Current Chemistry, 263 (2006), pp. 163-189

Peer Review under the responsibility of Universidad Nacional Autónoma de México.

Copyright © 2017. Universidad Nacional Autónoma de México, Facultad de Química
Opciones de artículo
Herramientas