You are on page 1of 9

The Role of the Frontier Orbitals

in Acid–Base Chemistry of Organic


Amines Probed by Ab Initio and
Chemometric Techniques
FELIPE A. LA PORTA, REGIS T. SANTIAGO,
TEODORICO C. RAMALHO, MATHEUS P. FREITAS,
ELAINE F. F. DA CUNHA
Department of Chemistry, Federal University of Lavras, Campus Universitário, CP 3037,
37200-000 Lavras-MG, Brazil

Received 23 October 2009; accepted 15 February 2010


Published online 21 April 2010 in Wiley InterScience (www.interscience.wiley.com).
DOI 10.1002/qua.22676

ABSTRACT: The Frontier effective-for-reaction molecular orbital (FERMO) concept


emerges as a powerful and innovative implement to investigate the role of molecular
orbitals (MOs) applied in the description of breakage and formation of chemical bonds.
In this work, theoretical calculations were carried out for conjugated acids of 18 amines
and their acid–base behavior was analyzed using MO energies. We observed that
highest occupied MO (HOMO) energies are inadequate to describe the acid–base
behavior of these compounds. By using the FERMO concept, the reactions that are
driven by HOMO, and those that are not, can be better explained, independent of the
calculation method used, as independent of the calculation method used, both HF and
Kohn–Sham methodologies lead to the same FERMO. V C 2010 Wiley Periodicals, Inc. Int J

Quantum Chem 110: 2015–2023, 2010

Key words: FERMO; amines; pKb; molecular orbital

molecules. Consequently, a precise knowledge


1. Introduction about the acidity and basicity of organic com-
pounds in various solvents is fundamental for the

P roton transfer is one of the most important


processes in the transformations of organic
study of mechanistic organic chemistry [1, 2].
Among the several compound classes, amines
play an important role in organic chemistry.
Correspondence to: T. C. Ramalho; e-mail: teo@dqi.ufla.br
Contract grant sponsors: FAPEMIG, CNPq.
Amines are relatively weak bases, one of the most
Additional supporting information may be found in the online common organic bases [2], their biological impor-
version of this article. tance is undeniable, because of their appearance

International Journal of Quantum Chemistry, Vol. 110, 2015–2023 (2010)


V
C 2010 Wiley Periodicals, Inc.
LA PORTA ET AL.

in many substances. Another usual way is to clas- The FERMO concept comes from a dose of
sify them as aliphatic, aromatic, and heterocyclic intuition, together with the criteria for the compo-
amines. A basic strength of amines may be con- sition and location to correctly determine the reac-
veniently provided by means of the basicity con- tant MO. This concept can be understood as a
stants (pKb) of their acid conjugates [3]. Evidently, complement to the HOMO–LUMO argument.
basicity is one of the important factors to tune the Another important feature in FERMO is that both
reactivity of such transformations. The basicity of HF and Kohn–Sham orbitals make the same con-
these compounds is related to the lone pair avail- clusions about reactivity [21–24]. A careful study
ability [2]. is necessary to understand when the use of
Molecular orbitals (MOs) and their properties, HOMO energy works or not. Moreover, there is a
such as energies and symmetries, are very useful lack of studies concerning the most common and
for chemists. According to Fukui, the MO proper- important base compounds in organic chemistry,
ties have used the Frontier electron density for the amines, and the relationship of their acid–
predicting the most reactive position in p-electron base behavior with their MO energies. For this
systems. Hoffmann and Woodward set out orbital reason, our primary goal in this work was to
symmetry rules to explain several types of reac- investigate, which is the best MO for describing
tions in conjugated systems, and the Frontier the acid or base character for a set of 18 amines.
MOs gained importance for the better under- Additionally, this work aims to apply the FERMO
standing of chemical reactions. The concept of concept to describe the acid–base behavior (pKb)
Frontier orbital, introduced by Fukui around of a series of amines.
1952, relates reactivity with the properties of two
MOs: highest occupied MO (HOMO) and lowest
unoccupied MO (LUMO). One application of the
HOMO–LUMO argument is the description of the
acid–base behavior of compounds [4–7]. 2. Computational Methods
The calculation of gas-phase acid–base parame-
ters is a well-established methodology. However, In recent years, theoretical methods based on
the theoretical calculation of acid–base parame- the density functional theory (DFT) have emerged
ters, especially in solution, is a great challenge for as an alternative to traditional ab initio methods
quantum chemists [8]. Turning to gas phase, the in the study of structure and reactivity of chemi-
proton affinity (PA) is the most useful acid–base cal systems [26, 27]. All calculations were carried
parameter and many works have shown how PA out with the Gaussian 98 package [28]. Each con-
values correlate with quantum descriptors [9–13]. jugated base of all 18 compounds was fully opti-
The PA of a molecule is a measure of its gas-phase mized by using DFT with the B3LYP functional
basicity. Thus, the higher the PA, the stronger the [29, 30] using the 6-31G(d,p) basis set. No sym-
base and the weaker the conjugate acid in the gas metry constraint was imposed during the opti-
phase. In other words, proton affinities illustrate mization process. Those optimized geometries
the role of hydration in aqueous-phase Brønsted were used in all subsequent calculations. This
acidity. Consequently, all acid–base reactions theoretical level was also used for the frequency
involve proton transfers between combined pairs. calculations. Furthermore, ab initio HF energy
Some studies in the literature show that the calculations were computed using the 6-31G(d,p)
energy of HOMO is directly related to the acid– basis set.
base behavior of some compounds. However, To account for solvent effects (water), single-
HOMO is not the effective Frontier MO at least point energy calculations were performed using
for many compounds; in these cases, the energy the polarizable continuum model [31, 32] at both
of HOMO does not describe correctly the acid– the DFT, with the B3LYP functional, and MP2 lev-
base behavior [4, 14, 15]. els, using the 6-31G basis set.
Thus, given the limitations of HOMO–LUMO The orbital energies obtained from these calcu-
argument and the new approaches proposed in lations were fit to a linear model with experimen-
the literature to understand the chemical reactiv- tal pKb values and the determination coefficients
ity, one must move forward the role of MOs in (r2) were analyzed. The MO figures were pre-
chemistry, giving rise to a new idea: the Frontier pared using the Gauss View 2.1 package [28] with
effective-for-reaction MO (FERMO) [9, 16–25]. a contour value of 0.020.

2016 INTERNATIONAL JOURNAL OF QUANTUM CHEMISTRY DOI 10.1002/qua VOL. 110, NO. 11
ROLE OF THE FRONTIER ORBITALS

TABLE I
Organic amines and their pKb values.

Class Compound Amines pKb Ref.

Aromatic 1 Aniline 9.13 33


2 Benzylamine 4.66 33
3 2-methylaniline 9.55 33
4 3-methylamine 9.29 33
5 4-methylaniline 8.92 33
6 N-methylamine 9.15 33
Heterocyclic 7 2-ethylpyridine 8.11 33
8 2,3-dimethylpyridine 7.43 33
9 2,4-dimethylpyridine 7.01 33
10 2,5-dimethylpyridine 7.60 33
11 2,6-dimethylpyridine 6.65 33
12 3,4-dimethylpyridine 7.54 33
13 3,5-dimethylpyridine 7.85 33
Aliphatic 14 Ethylamine 3.35 34
15 Diethylamine 3.16 34
16 Triethylamine 3.25 34
17 Isobutylamine 3.52 34
18 tert-butylamine 3.30 34

lengths, have to be included in multiple linear


3. Results and Discussion regression analysis to improve the correlation [10–
15]. This procedure weakens the familiar idea that
3.1. HOMO AND CHEMICAL REACTIVITY donor-acceptor reactions are driven by Frontier
Eighteen compounds, classified as aromatic orbital energies.
(1–6), heterocyclic (7–13), and aliphatic (14–18) In line with this, this study is intended to res-
amines, were used in this study, and their experi- cue the role of the Frontier orbitals in acid–base
mental pKb values in water are depicted in Table I. chemistry of organic amines. Thus, the HOMO
Theoretical calculations of acid–base parame- energy values were correlated with the corre-
ters, especially in solution, are a great challenge sponding amine pKb values. From the relation-
for quantum chemists. Then, preliminary to deter- ships, we can observe that different amine classes
mining the solution-phase pKb of an amine, it is (aromatic, heterocyclic, and aliphatic) provided
important to estimate its gas-phase basicity. These different linear models [r2 ¼ 0.918 (aromatic);
results may be used to evaluate the solvent effect 0.099 (heterocyclic); 0.4538 (aliphatic)]. Heterocy-
on basicity parameters using theoretical calcula- clic amines exhibited higher deviations from the
tions of orbital energies. ideal linear model and a very poor linear model
Numerous works have shown how well pKb between HOMO and pKb values.
values correlate with quantum descriptors; the With those correlations, it was not possible to
average local ionization potential [35, 36], atomic distinguish, which is the best orbital to describe
charges and interatomic distances [14, 37], and the acid/base chemical reaction. In addition, from
the HOMO energy [10, 13, 15] are the most com- these data, two questions appear: (1) Can the
mon quantum descriptors used. amine groups be correlated together? and (2)
The relationships between HOMO energies and Would HOMO be the best orbital to describe the
pKa/pKb are often displayed for families of reaction? To find the answer for these questions,
compounds, such as phenols [14] and azines [15]. we have used a new chemically intuitive FERMO
However, for a number of other compounds, concept to describe the acid/base behavior [21–
HOMO energies do not show good correlation 25]. It has been shown elsewhere that the FERMO
with acid/base parameter values, and other quan- idea may be an alternative way to explain chemi-
tum parameters, such as dipole moments, bond cal phenomena when HOMO–LUMO arguments
orders, atomic charges on hydrogen, and bond fail or cannot be applied.

VOL. 110, NO. 11 DOI 10.1002/qua INTERNATIONAL JOURNAL OF QUANTUM CHEMISTRY 2017
LA PORTA ET AL.

FIGURE 1. Representation of the Frontier molecular orbital of amines 1, 13, and 14. [Color figure can be viewed in
the online issue, which is available at www.interscience.wiley.com.]

PN 2
Thus, based on the FERMO idea, we expanded where i¼1 /i (i ¼ N or all) is the sum of the
the correlations to other Frontier orbitals, namely squares of the eigenvalues associated with the
HOMO-1, HOMO-2, and HOMO-3 (see Fig. 1). nitrogen atomic orbital (AO) and all of the AOs in
Furthermore, using the FERMO concept, MO a particular MO, respectively. This calculation
composition and shape are taken into account to was performed on each of the seven HOMOs that
identify the MO, which will actually be involved had significant nitrogen character and these val-
in a given reaction. ues were then average.
On the basis of the orbital composition and When the solvent effects are considered, the
localization, it is possible to get insights about the MO energy positions relative to the one observed
orbital that mainly governs the acid/base reac- in the vacuum are interchanged (see Supporting
tion. The orbital composition is also an important Information for complete results). This is also
indicator for discovering the adequate orbital [21, observed when different basis sets are used. On
22]. The MO shape and the atomic composition the other hand, the solvent effects were neglected
are very important parameters for analyzing in other works correlating MO energies with
FERMO [23, 24]. The percentage of nitrogen char- acid/base parameters values [10, 14]. That
acter in some of the occupied (canonical) MOs in approximation seems valid for some aromatic sys-
the studied compounds was calculated from a full tems previously studied (anilines [10], phenols
population analysis, using Eq. (1): [12], and azines [15]), in which sizes and geome-
tries do not vary strongly. For these reasons, the
P 2 entropy term may be roughly the same along
/
% nitrogen character ¼ P 2N (1) each of these compound families and the solvent
/all effects will also be considered.

2018 INTERNATIONAL JOURNAL OF QUANTUM CHEMISTRY DOI 10.1002/qua VOL. 110, NO. 11
ROLE OF THE FRONTIER ORBITALS

TABLE II LUMOþn (n ¼ 0–4) were taken from the HF calcu-


Nitrogen contribution (%) and orbital energy for lations and matrixed as m rows (amines) by n col-
studied amines using Hartree–Fock approach. umns (energy of the Frontier MOs).
PCA [38] creates p latent variables (Y) as linear
Compound Contribution (%) Fermo (Hartree)
combinations of the original p variables (X), in
1 39.24 0.44053 such a way that new orthogonal axes are built to
2 6.69 0.39392 explain the maximum variance possible in just a
3 37.20 0.43862 few dimensions [Eq. (2)].
4 38.53 0.43220
5 38.35 0.43011 Y i ¼ ei T X (2)
6 44.80 0.41761
7 74.85 0.40465 where the unknown vector ei establishes the ith
8 73.81 0.40178 linear combination, for i ¼ 1,…,p.
9 75.94 0.40249 The PCA allowed clustering the three amine
10 74.05 0.40190 types, namely aliphatic, aromatic, and heterocyclic
11 74.75 0.40028
amines, by using the two first principal compo-
12 73.76 0.40415
13 73.62 0.40313
nents (PC1 and PC2). Compound 2, an aliphatic
14 7.50 0.38174 amine containing an aromatic ring, did not fit in
15 86.49 0.36192 any other group. PC1 was capable of separating
16 88.44 0.34371 aliphatic amines from the aromatic ring-contain-
17 83.44 0.38340 ing amines, whereas PC2 differentiated the heter-
18 80.11 0.37651 ocyclic amines from the aromatic ones [Fig. 2(a)].
This may be rationalized in terms of the higher
energy of the nonresonant lone pair contribution
Upon analyzing orbital localization and compo-
of nitrogen in aliphatic amines to the HOMO-type
sition more deeply, it can be observed that there
orbitals, and to the lower energies of the LUMO-
is a MO with energy quite close to the HOMO
type orbitals in the aromatic compounds (p* con-
energy value and with large nitrogen contribution
tribution). From the loadings analysis, the four
(Table II). Thus, it is supposed that this orbital
MOs contributing importantly for the groups sep-
could describe the acid/base behavior better than
aration in PC1 were HOMO1 > LUMOþ1 >
HOMO. Moreover, it also does not fit our orbital
LUMO > HOMO, whereas HOMO2 exhibited
choice criterion, because it is not common to all
high weight for the separation of heterocyclic
studied compounds and is not mainly located
amines in PC2 [Fig. 2(b)]. The energies of the five
where the reaction takes place, in the nitrogen
most weighted orbitals in PC1 and PC2 were
atom. The HOMO problem arises from the influ-
regressed against the pKb values using multiple
ence of the aromatic p-electrons, and the solution
linear and PLS regressions. The best PLS model
would be finding an MO that does not have this
was found using two latent variables, which
influence. From this point of view, other concepts
explain 96% of data variance, and the calibration
of chemical reactivity could be useful to under-
squared correlation coefficient (r2) was 0.814 (root
stand reactions that were not driven by HOMO–
mean square error of calibration, RMSEC ¼ 1.009),
LUMO properties.
with a leave-one-out cross-validation squared cor-
relation coefficient (q2) of 0.799 (root mean square
3.2. FRONTIERS ORBITAL AND error of validation, RMSEV ¼ 1.05); r2 and q2 are
CHEMOMETRIC TECHNIQUES
improved to 0.877 (RMSEC ¼ 0.826) and 0.954
To understand the reactivity of amines and es- (RMSEV ¼ 0.504) after removing compound 2, an
tablish a relationship with MOs, the amine sets, outlier. The MLR model for the whole series of
namely those aliphatic, aromatic, and heterocyclic, compounds was highly predictive, with an r2 of
were classified according to their Frontier MOs, 0.926 and the following MLR equation [Eq. (3)]:
which were subsequently correlated with the cor-
responding pKb values. In addition, principal com- pKb ¼12:83  34:67  EHOMO2  4:01  EHOMO1
ponent analysis (PCA) and partial least squares
þ 44:54  EHOMO  57:74  ELUMO
(PLS) were used as explorative and regression
tools, respectively. The energies of HOMOn and þ 6:53 LUMOþ1 ð3Þ

VOL. 110, NO. 11 DOI 10.1002/qua INTERNATIONAL JOURNAL OF QUANTUM CHEMISTRY 2019
LA PORTA ET AL.

FIGURE 2. (a) Scores in PC1 and PC2. (b) Loadings in PC1 and PC2.

From the Lewis definition of acid–base reactions, highly correlated energies (Table III). In this way,
which is a further generalization that encom- only HOMO1 and LUMOþ1 were selected to
passes the Brønsted–Lowry definition and the sol- derive the MLR-based model. The MLR squared
vent-system definitions [2]. The Lewis defines a correlation coefficient using HOMO1 and
base (referred to as a Lewis base) to be a com- LUMOþ1 as independent variables was 0.783,
pound that can donate an electron pair, and an and the following model was obtained [Eq. (4)]:
acid (a Lewis acid) to be a compound that can
receive this electron pair [2]. So, the basicity is pKb ¼ 20:81 þ 34:87  EHOMO1  5:61  ELUMOþ1
mainly related to occupied MOs in accordance (4)
with the chemical intuition.
In this point, it is important to mention that the Despite the nice correlation obtained using both
HOMO2 and HOMO exhibited higher weight in regression methods PLS and MLR (see Fig. 3).
the correlation with pKb for occupided MOs. In The MLR model is based on contributions of
line with this, we can then notice that the contri- some MOs apparently not related to basicity,
butions from the HOMO2 and HOMO might according to the chemical intuition (e.g., as is the
determine the acid–base chemistry of organic nitrogen lone pair), making physicochemical inter-
amines. pretation difficult; simple and strongly correlative
The MLR model may be simplified by remov- models may be derived by using the FERMO con-
ing collinear descriptors, that is, those MOs with cept and chemometric techniques, which are

TABLE III
Correlation matrix of selected orbital energies (r2 depicted).

HOMO4 HOMO3 HOMO2 HOMO1 HOMO LUMO LUMOþ1 LUMOþ2 LUMOþ3 LUMOþ4

HOMO4 1.000
HOMO3 0.476 1.000
HOMO2 0.329 0.022 1.000
HOMO1 0.300 0.086 0.501 1.000
HOMO 0.309 0.070 0.253 0.839 1.000
LUMO 0.132 0.002 0.783 0.692 0.434 1.000
LUMOþ1 0.232 0.003 0.846 0.748 0.493 0.966 1.000
LUMOþ2 0.343 0.050 0.490 0.869 0.824 0.629 0.711 1.000
LUMOþ3 0.449 0.073 0.584 0.865 0.824 0.648 0.750 0.943 1.000
LUMOþ4 0.416 0.097 0.582 0.872 0.806 0.661 0.736 0.870 0.865 1.000

2020 INTERNATIONAL JOURNAL OF QUANTUM CHEMISTRY DOI 10.1002/qua VOL. 110, NO. 11
ROLE OF THE FRONTIER ORBITALS

FIGURE 3. MLR and PLS modeling of pKb using HOMO2, HOMO1, HOMO, LUMO, and LUMOþ1 energies.

referred to as those MOs with higher contribution tional, because the LYP correlation functional sig-
of nitrogen AOs (intuitively related to the amines nificantly underestimates the correlation energy of
basicity). the homogeneous electron gas [39].
Thus, motivated by this expectation, a closer On the other hand, for large gap systems, with
examination in Table III reveals that HOMO2 well-localized electrons, the agreement with
and HOMO are MOs with highly correlated ener- experiment and conventional functionals is rea-
gies. This means that HOMO2 and HOMO sonable [39, 43]. In line with that our theoretical
could, in principle, have similar properties, such results using B3LYP combined with the FERMO
as symmetry and ionization potential. In addition, concept could be justified. In these cases, the re-
it is certainly worth noticing those orbitals also spective errors for the constituents of the reaction
show a high nitrogen contribution. cancel, as electrons are localized for all reactants
Furthermore, it is important to note that the and the product [21–25].
FERMO concept leads to same conclusions about
chemical reactivity for both HF and Kohn–Sham
approaches [21–24]. Nowadays, it is well-known 3.3. FERMO AND STATISTICAL
CORRELATIONS
that the Kohn–Sham DFT is a leading method for
electronic structure calculations in chemistry A molecule can have as many FERMOs as it
mainly due to its high efficiency and relatively has reaction sites, and it could be the HOMO or
low computational cost [19, 33]. any other Frontier MO [22]. Reactions involving
Nevertheless, despite the recent improvements electron donation and acceptance are related to
in DFT, there are still difficulties in using DFT to MO energies, as electrons are occupying and will
properly describe intermolecular interactions, espe- occupy a MO and a Frontier orbital, as stated by
cially van der Waals forces (dispersion) and charge Fukui [4–6]. A criterion has been established in
transfer excitations due to lack of exact Hartree– this study to determine the FERMO. In this con-
Fock exchange in some functionals [39–42]. cept, FERMO provides the adequate orbital shape
Currently, the development of new DFT meth- and composition to correlate with reactive indexes.
ods designed to overcome this problem, by altera- The FERMO for aromatic and aliphatic amines
tions to the functional or by the inclusion of addi- is the HOMO, whereas HOMO2 is the invoked
tive terms is a very important research topic. FERMO for heterocyclic amines. Thus, when
Particularly, the hybrid functionals, such as the HOMO fails in describing the reactivity parame-
B3LYP functional used in this work, yields unsatis- ter, another Frontier MO would be responsible for
factory atomization energies [39–41]. For instance, the phenomenon. Using this modification, the cor-
for metals, the error is almost larger than 25%, relation between the orbital energies and pKb val-
and for transition and noble metals, it may ues for the heterocyclic amines enhanced from
increase to 40%.[39, 40] Probably, a larger fraction 0.099 to 0.631. Based on the MO location and
of the error is related to the LYP correlation func- composition, the FERMOs shown in Figure 4

VOL. 110, NO. 11 DOI 10.1002/qua INTERNATIONAL JOURNAL OF QUANTUM CHEMISTRY 2021
LA PORTA ET AL.

place in specific regions of the molecule. Therefore,


currently, the need for inexpensive computational
techniques capable of reliably contributing to the
prediction of the chemical reactivity based on
chemical intuition is emerging. We believe that the
FERMO idea could, in combination with other
approaches, offer contributions in this sense.
Thus, we may notice that the FERMO idea is to
use the MO calculations with a valence bond
theory (VB) interpretation. In MO theory, the elec-
trons in a molecule occupy delocalized orbitals
FIGURE 4. Surface plots for FERMO at the HF level made from linear combination of AOs. However,
for the aromatic (A), heterocyclic (B), and aliphatic (C) it should be kept in mind that the VB approaches
amines. [Color figure can be viewed in the online issue,
are quite useful in analyzing enzymatic reactivity
which is available at www.interscience.wiley.com.]
[44, 45], photochemistry [46–45], chemical dynam-
ics [46], and theories of conductivity, where the
could be found for the aromatic (A), heterocyclic localized representation seemed indispensable.
(B), and aliphatic (C) amines. In fact, it is well known that the VB methods
From Figure 1, it is clear that the HOMO is allow generating new ideas on chemical bonding.
formed mainly by the AO in the ‘Z’ axis of the Recently, there has been an intense surge of con-
nitrogen, whereas the FERMO has a high contri- cepts related to localized representation, for exam-
bution of the AO of nitrogen in the ‘XY’ plan ple, multiple bonding between transition metals
(plan of the aromatic ring) of the amine group. [50, 51] and the development of natural resonance
A direct consequence of the different composi- theory [52].
tions of these orbitals is the position of their node
plains in relation to the plan formed by nitrogen
atom. Thus, the statistical results from chemomet-
ric methods reinforce our rationalization: FERMO 4. Conclusions
could in principle be the orbital to better describe
the acid–base chemistry of organic amines. Overall, the FERMO concept has been success-
The previous results take us to a key question: fully applied to describe the acid–base behavior
why are FERMO energies better quantum descrip- of the series of amines studied, presenting better
tors for pKb values than HOMO energies? The an- correlation with pKb than HOMO. Thus, the
swer to this question resides in the shape of those FERMO concept may be useful to explain acid–
MOs. It is well known that organic amine moi- base reactions in a wide range of applications.
eties from these bases have a specific geometry Amine types were clustered using PCA and the
[1–7] because of the lone pair distribution. We more weighted MOs were accurately regressed
should keep in mind that acid/base reactions are against the pKb values using MLR and PLS, dem-
localized and the basicity of these compounds is onstrating that different MOs other than HOMO
related to the lone pair availability [2]. may be applied to explain the basicity-based reac-
Upon analyzing the localization and the com- tions of amines. Nevertheless, it is the FERMO
position of the orbitals more deeply, it can be concept (the shape of the reactant orbital) that
observed that there is an MO with energy quite really provided chemical insight about which is
close to the HOMO energy value, with highly cor- the MO related to acid–base reactions. It should
related energies (Table III), and with large nitro- be kept in mind that understanding the behavior
gen contribution (Table II). of the MO is crucial to understanding the chemis-
It is well known that in the MO theory, the try. Therefore, these investigations lead to new
electrons in a molecule are in the delocalized orbi- perspectives and ideas about the reactivity, and
tals generated from linear combination of AOs. may be expanded to solve other problems. For
Nevertheless, sometimes these delocalized orbitals example, the principle of hardness and enervation
are quite diffuse leading to a difficult rationaliza- of Pearson may invoke the FERMO concept. The
tion of the chemical reactivity. In contrast, some hardness of a molecule is defined as the differ-
chemical reactions, such as acid–base reaction, take ence in energy between HOMO and LUMO; thus,

2022 INTERNATIONAL JOURNAL OF QUANTUM CHEMISTRY DOI 10.1002/qua VOL. 110, NO. 11
ROLE OF THE FRONTIER ORBITALS

according to that definition, only a single value of 24. Ramalho, T. C.; Martins, T. L. C, Borges, L. E. P. Int
hardness would exist for a given molecule. But as J Quant Chem 2003, 95, 267.
we know, many molecules have soft and hard 25. Ramalho, T. C.; Pereira, D. H. Mol Sim 2009, 35, 1269.
sites and the HOMO–LUMO energy difference is 26. Hohenberg, P.; Kohn, W. Phys Rev 1964, 136, B864.
not enough to describe this type of behavior. Our 27. Kohn, W.; Sham, L. J Phys Rev 1965, 140, A1133.
results clearly reveal the fundamental relation- 28. Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.;
Robb, M. A.; Cheeseman, J. R.; Zakrzewski, V. G.; Mont-
ships between the nature of chemical bonding
gomery, J. A.; Stratmann, R. E.; Burant, J. C.; Dapprich, S.;
and the acid–base chemistry of organic amines. Millam, J. M.; Daniels, A. D.; Kudin, K. N.; Strain, M. C.;
Farkas, O.; Tomasi, J.; Barone, V.; Cossi, M.; Cammi, R.;
ACKNOWLEDGMENT Mennucci, B.; Pomelli, C.; Adamo, C.; Clifford, S.; Ochter-
ski, J.; Petersson, G. A.; Ayala, P. Y.; Cui, Q.; Morokuma,
We thank the Brazilian agency FAPEMIG, K.; Salvador, P.; Dannenberg, J. J.; Malick, D. K.; Rabuck,
CNPq for funding part of this work. CNPq is also A. D.; Raghavachari, K.; Foresman, J. B.; Cioslowski, J.;
gratefully acknowledged for the fellowships (to TCR Ortiz, J. V.; Baboul, A. G.; Stefanov, B. B.; Liu, G.; Lia-
shenko, A.; Piskorz, P.; Komaromi, I.; Gomperts, R.; Mar-
and MPF) and CENAPAD-SP for the computational
tin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C.
facilities. Y.; Nanayakkara, A.; Gonzalez, C.; Challacombe, M.; Gill,
P. M. W.; Johnson, B. G.; Chen, W.; Wong, M. W.; Andres,
J. L.; Head-Gordon, M.; Replogle, E. S.; Pople, J. A. Gaus-
sian 98 (Revision A.11); Gaussian: Pittsburgh, 2001.
References 29. Becke, A. D. J Chem Phys 1993, 98, 5648–5652.
30. Lee, C.; Yang, W.; Parr, R. G. Phys Rev B 1988, 37, 785.
1. Li, J.-N.; Fu, Y.; Liu, L.; Guo, Q. X. Tetrahedron 2006, 62,
31. Miertus, S.; Scrocco, E.; Tomasi, J. E. J Chem Phys 1981, 55, 117.
11801.
32. Miertus, S. Chem Phys 1982, 65, 239.
2. Bruice, P. Y. Quı́mica orgânica, 4th ed.; São Paulo: Pearson
Prentice Hall, 2006; 641. 33. Lide, D. R. CRC Handbook of Chemistry and Physics: A
Ready-Reference Book of Chemical and Physical Data, 88th
3. Solomons, T. W. G.; Fryhle, C. B. Quı́mica orgânica, 8th
ed.; 2007–2008 Boca Raton: CRC Press, 2008.
ed.; Rio de Janeiro: Livros Técnicos e Cientı́ficos, 2005–
2006; 542. 34. Morrison, R. T.; Boyd, R. N. Quı́mica orgânica, 13th ed.;
4. Fukui, K.; Yonezawa, T.; Shingu, H. J Chem Phys 1952, 20, 722. Lisboa: Fundação Calouste Gulbenkian, 1996.
5. Fukui, K.; Yonezawa, T.; Nagata, C. J Chem Phys 1954, 22, 35. Gross, K. C.; Seybold, P. G.; Peralta-Inga, Z.; Murray, J. S.;
1433. Politzer, P. J Org Chem 2001, 66, 6919.
6. Fukui, K. Angew Chem Int Ed 1982, 21, 801. 36. Brinck, T.; Murray, J. S.; Politzer, P. J Org Chem 1991, 56,
5012.
7. Hoffmann, R.; Woodward, R. B. Acc Chem Res 1968, 1, 17.
37. Gross, K. C.; Seybold, P. G. Int J Quant Chem 2000, 80, 1107.
8. Hehre, W. J.; Random, L.; Schleyer, P. P.; People, J. A.
Wiley-Interscience: New York, 1986; 226. 38. Pearson, K. Philos Mag 1901, 6, 559.
9. Vianello, R.; Makisié, Z. B. Tetrahedron 2006, 62, 3402. 39. Paier, J; Martijn., M.; Kresse, G. J Chem Phys 2007, 127,
024103.
10. Pytela, O.; Otyepka, M.; Kulhánek, J.; Otyepková, E.;
Nevëëná, T. J Phys Chem A 2003, 107, 11489. 40. Paier, J.; Marsman, M.; Hummer, K.; Kresse, G.; Gerber, I.
11. Gruber, C.; Buss, V. Chemosphere 1989, 19, 1595. C.; Ángyán, J. G. J Chem Phys 2006, 124, 154709.
12. Citra, M. J. Chemosphere 1999, 38, 191. 41. Ramalho, T. C.; Taft, C. A. J Chem Phys 2005, 123, 54319.
13. Soriano, E.; Cerdán, S. J Mol Struct 2004, 686, 121. 42. Prytz, O.; Flage-Larsen, E. J. Phys Condens Matter 2009, 22,
015502.
14. Gross, K. C.; Seybold, P. G. Int J Quant Chem 2001, 85, 569.
43. Ramalho, T. C.; Oliveira, L. C. A.; Carvalho, K. T. G.; da
15. Machado, H. J. S.; Hinchliffe, A. J Mol Struct Theochem
Cunha, E. F. F.; Souza, E. F. Mol Phys 2009, 107, 171.
1995, 339, 255.
44. Cimiraglia, R.; Malrieu J. P. J Phys B 1985, 18, 3073.
16. Hirao, H.; Ohwada, T. J Phys Chem A 2003, 107, 2875.
45. Nakamura, H.; Truhlar, D. G. J Chem Phys 2001, 115, 10353.
17. Hirao, H.; Ohwada, T. J Phys Chem A 2005, 109, 816.
18. Fujimoto, H.; Mizutani, Y.; Iwase, K. J Phys Chem 1986, 90, 46. Truhlar, D. G. J Comp Chem 2007, 28, 73.
2768. 47. van der Lugt, A.; Oosterhoff, L. J. J Am Chem Soc 1969, 91,
19. Fujimoto, H. Acc Chem Res 1987, 20, 448. 6042.
20. Fujimoto, H.; Satosh, S. J Phys Chem 1994, 98, 1436. 48. Atchity, G. J.; Ruedenberg, K. Theor Chem Acc 1997, 97, 47.
21. Da Silva, R. R.; Ramalho, T. C.; Santos, J. M.; Figueroa- 49. Robb, M. A.; Garavelli, M.; Olivucci, M.; Bernardi, F. Rev
Villar, J. D. J Phys Chem A 2006, 110, 1031. Comp Chem 2000, 15, 87.
22. Da Silva, R. R.; Santos, J. M.; Ramalho, T. C.; Figueroa- 50. Brynda, M.; Gagliardi, L.; Widmark, P. O.; Power, P. P.;
Villar, J. D. J Braz Chem Soc 2006, 17, 223. Ross, B. O. Angew Chem Int Ed 2006, 45, 3804.
23. Da Silva, R. R.; Santos, J. M.; Ramalho, T. C.; Figueroa- 51. Weinhold, F.; Landis, C. Science 2007, 316, 61.
Villar, J. D. J Phys Chem A 2006, 110, 10653. 52. Jug, K.; Hiberty, P. C.; Shaik, S. Chem Rev 2001, 101, 1477.

VOL. 110, NO. 11 DOI 10.1002/qua INTERNATIONAL JOURNAL OF QUANTUM CHEMISTRY 2023

You might also like