Topological Analysis of The Electron Localization Function (ELF) Applied To The Electrophilic Aromatic Substitution

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

852 J. Phys. Chem.

A 2000, 104, 852-858

Topological Analysis of the Electron Localization Function (ELF) Applied to the


Electrophilic Aromatic Substitution
Franck Fuster, Alain Sevin, and Bernard Silvi*
Laboratoire de Chimie Théorique (UMR-CNRS 7616), UniVersité Pierre et Marie Curie, 4 Place Jussieu,
75252-Paris cédex, France
ReceiVed: August 5, 1999; In Final Form: October 20, 1999

The topological analysis of the electron localization function ELF provides a partition of the molecular space
into basins of attractors which have a clear chemical signification. The hierarchy of these basins is given by
the bifurcation of the localization domains. In the case of π-donor substituents (OH, NH2, F, CH3, C6H5, Cl),
the aromatic domain is first opened close to the substituted carbon and then in the vicinity of the meta carbon;
whereas for attractor substituents (CN, CHO, NO2, CF3 and CCl3), it is first opened in the ortho and para
positions. The orienting effects of the electrophilic substitutions are correlated with these bifurcations. The
experimental favored positions always correspond to the locally electronegative carbons (i.e., those which
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

keep their shell structure at the higher ELF values). This suggests that the local Pauli repulsion plays a noticeable
Downloaded via UNIV ESTADUAL DE CAMPINAS on July 6, 2021 at 18:58:48 (UTC).

role in the orienting effects which are complementary to the charge transfer effect involved in standard quantum
chemical pictures.

1. Introduction refers to the isolated aromatic entity. As Bader’s, our approach


Since the early work of Kekulé, aromatic compounds have is only related to the first type of argument. The article is
played a major role in the development of chemical concepts. organized as follows: after a brief methodological introduction,
Though their reactivity, and more especially the so-called elec- the role of various ortho-para and meta orienting substituents
trophilic substitution, has been widely studied and empirically on benzene is investigated in order to establish in which way
classified according to Holleman’s rules1-4 as early as 1925, the topology of the ELF function accounts for the Holleman’s
the progressive appearance of a rationale has followed the pro- rules. The analysis is then generalized to multiple substitutions
gress of the electronic theory of molecules. The close interplay and to polyaromatic substituted hydrocarbons.
between experiment and theory is illustrated by the fundamental 2. Methodology
contribution of Hückel5-7 in the 1930s and Wheland8 and
The quantum mechanical calculations have been performed
Ingold9 in the 1950s and then by the classical MO approach,
with the standard 6-31G** basis set29-31 at the Hartree-Fock
illustrated by the classical works of Coulson,10 Streitwieser,11
and hybrid Hartree-Fock density functional Becke3LYP32-35
Salem,12 and Dewar13 in the 1960s and Epiotis14-17 in the 1970s.
levels using the Gaussian94 package.36 The aim of the topologi-
Recently, new proposals that constitute a revival of the topic
cal analysis of ELF is to provide a mathematical model of the
have appeared on the grounds of Bader’s atoms in molecules
Lewis representation. The simplest mathematical structure
theory18-20 as well as of the molecular electrostatic potential.21
enabling the partition of the molecular space and accounting
Even more recently, many applications of the Becke and
for its evolution upon the variation of parameters (such as the
Edgecombe’s electron localization function (ELF)22 have proven
nuclear coordinates within the Born-Oppenheimer approxima-
to allow for an efficient description of chemical bonding in
tion) is a gradient dynamical system. A gradient dynamical
stable molecules and short-lived intermediates.
system is the gradient vector field of a scalar, continuous,
We propose here an investigation of Holleman’s rules in derivable function named the potential function. The potential
electrophilic aromatic substitution using ELF methodology. Our function carries the physical (or chemical) information. For
goal is dual. First, we hope to illustrate constructive new example, the gradient field of the electron density distribution
concepts which, second, we will use as a pretext to comfort yields a partition into atomic basins.18 Many studies have shown
ELF results with firmly established evidence. that the electron localization function (ELF) of Becke and
Prior to the detailed account of our study, it is worth recalling Edgecombe22 yields a faithful description of the bonding
the general framework which is used in the investigation of consistent with the Lewis approach25-28,37-41 as well as it
electrophilic substitution mechanism. Actually, two different accounts for its evolution during reactive processes.42-45
kinds of approaches are often made. The first one considers The Gaussian wfn output was then treated with the TopMod
the electronic properties of a suitably substituted aromatic ring package written in our group.46 The latter program first
and deduces the preferential site of attack by the incoming calculates the ELF, over the molecular space, according to its
substituent. This method may be referred to as “reactant like”. definition
The second considers the relative stability of ortho, meta, and
para Wheland cationic intermediates and thus deduces the 1

( )
resulting regioselectivity. The latter approach is typically η(r) ) (1)
Dσ(r) 2
“intermediate like”, and upon the assumption that the transition 1+
state closely resembles the intermediate species, it no longer Dσh (r)

* E-mail: [email protected]. Fax: +033 01 4427 4053. where Dσ stands for the excess local kinetic energy due to the
10.1021/jp992783k CCC: $19.00 © 2000 American Chemical Society
Published on Web 01/07/2000
Analysis of the Electron Localization Function J. Phys. Chem. A, Vol. 104, No. 4, 2000 853

Pauli restriction,23 i.e., the difference between the definite


positive kinetic energy density Ts(r) of the actual fermionic
system and that of the von Weizsäcker kinetic energy functional
TuW(r).47 If the wave function is written as a single determinant,
Dσ is expressed in terms of orbital contributions:

1 |∇F(r)|
2
1
D (r) )
σ

2
∑i |∇φi(r)| 2
-
8 F(r)
(2)

Dσh is the kinetic energy of the electron gas having the same
density:

Dσh (r) ) CFF(r)5/3 (3)

where CF is the Fermi constant. Where electrons are alone, or


form pairs of antiparallel spins, the Pauli principle has little
influence on their behavior and the excess local kinetic energy
has a low value and, therefore, ELF is close to 1, whereas at
the boundaries between such regions the probability of finding
parallel spin electrons close together is rather high and the excess
local kinetic energy has a large value and ELF is small. The
topological analysis is carried out in order to partition the
molecular space into basins of attractors denoted by ΩA. Each
basin is characterized by an attractor which is a local maximum
of ELF. One distinguishes core basins, labeled C(Xi), centered
on atoms Xi, and valence basins, between atoms, labeled V(Xi,
Xj, ...). For each basin one defines its synaptic order which is
the number of core basins that it is connected to. Integration of
Figure 1. Localization domain reduction tree-diagram of benzene.
the electronic density over the ΩA basin yields its population.
The dependence of basin populations upon basis set and
correlation effects is consistent with the chemical intuition. On
one hand, the addition of polarization functions increases the
disynaptic (bonding) basin populations at the expense of the
monosynaptic ones (lone pairs), and correlatively the dissocia-
tion energy and the stretching force constants are enhanced. On
the other hand, correlation yields the opposite trend in agreement
with the lowering of the force constants.39 Moreover, there is
an overall very good agreement between the MP2 and DFT
values38,39 though the former method appears to be much more
basis set dependent. In the systems investigated here, the electron
correlation is not expected to play an important role with respect
to the electronic state or to the structure; moreover it has been
shown that the properties related to the ELF topological analysis
are almost converged with a polarized split-valence basis set.28
As it will be seen in the next sections, the value of the ELF
function at some particular critical points is of a great importance
in the present study. These critical points are located by a
steepest ascent search which uses the analytical derivatives of
the norm of the ELF gradient.
Graphical representations of the bonding are obtained by
plotting isosurfaces of the localization function which delimit
volumes within which the Pauli repulsion is rather weak. These
latter, the localization domains, are called irreducible when they
contain one and only one attractor.

3. f-Localization Domains in the Model Study of Benzene


These domains define the bodies limited by a given isosurface
η(r) ) f, enclosing points for which η(r) > f. They are said to Figure 2. ELF isosurface of benzene. Value of ELF is equal to (A)
be reducible when they contain more than one attractor. For a 0.5, (B) 0.64, and (C) 0.65. The gray scale code used for the localization
domain is as follows: core, black; valence protonated, gray; valence
low value of ELF there exists only one reducible localization disynaptic, light gray. This figure, presented here in black and white,
domain which contains all the molecule population. Upon pro- is available in color on the World Wide Web in Supporting Information.
gressive increase of ELF, the overall isosurface splits into reduc- Color code: magenta ) core, red ) valence monosynaptic, green )
ible valence and irreducible core domains. Thus, there exists a valence disynaptic.
854 J. Phys. Chem. A, Vol. 104, No. 4, 2000 Fuster et al.

Figure 3. η ) 0.65 localization domains of monosubstituted benzenes: (a) C6H5NH2, (b) C6H5OH, (c) C6H5F, (d) C6H5CH3, (e) C6H5Cl, (f)
C6H5-C6H5, (g) C6H6, (h) C6H5CCl3, (i) C6H5CF3, (j) C6H5CN, (k) C6H5CHO, (l) C6H5NO2. This figure, presented here in black and white, is
available in color on the World Wide Web in Supporting Information. Color code: magenta ) core, red ) valence monosynaptic, green ) valence
disynaptic.
hierarchy of splittings which can be visualized as a bifurcation various domains yields the bifurcation diagram of Figure 1,
tree diagram, as displayed for the benzene of Figure 1. which summarizes the various preceding steps.
In Figure 2A,B,C are reported the ELF isosurfaces for η(r) The analysis of a bifurcation diagram brings interesting
) 0.5, 0.6, 0.65, respectively. The complete topology of benzene information about the electron localization on the various sites
might be displayed upon progressive variation of ELF. For a of a given aromatic compound. Indeed, electronic pairing is
low value of ELF, e0.5 (Figure 2A), one gets a contour which more pronounced around a site bearing some negative charge.
only reveals the 6-fold shape, all domains being fused. At ELF Consequently, the corresponding localization domains appear
) 0.6 (Figure 2B), the preceding bulk volume has split, at at larger ELF values than for their related less negative
critical ELF ) 0.58 value, into six V(Ci, Hi)i)1,6 protonated counterparts. Thus, the bifurcation diagram distinguishes the
domains and a distorted toroidal volume which is typical of preferential sites of electrophilic attack as these which separate
aromaticity. In coming discussions, it will be referred to as at highest ELF values.
“aromatic domain”. In turn, the aromatic domain splits into six
distinct V(Ci, Ci+1)i)1,5 and V(C1, C6) domains at the critical 4. Monosubstituted Benzene Derivatives
ELF value of 0.6449, as displayed in Figure 2C. The latter value Figure 3 displays the localization domains bounded by the
of ELF will be used for the sake of comparison when dealing η(r) ) 0.650 isosurface of benzene and 11 monosubstituted
with substituted compounds. This progressive splitting of the derivatives C6H5-S. For the sake of the analysis it is useful to
Analysis of the Electron Localization Function J. Phys. Chem. A, Vol. 104, No. 4, 2000 855

TABLE 1: Calculated Electrophilic Substitution Positional


Indices of Monosubsituted Benzenes
6-31G**/HF 6-31G**/Becke3LYP
S ortho meta para ortho meta para
NH2 0.042 -0.021 0.029 0.039 -0.017 0.024
OH 0.035 -0.019 0.024 0.030 -0.012 0.018
F 0.024 -0.012 0.011 0.019 -0.006 0.009
CH3 0.013 -0.007 0.010 0.016 -0.004 0.006
C6H5 0.007 -0.004 0.009 0.009 -0.002 0.004
Cl 0.004 -0.004 0.001 0.008 -0.001 0.002
H 0.0 0.0 0.0 0.0 0.0 0.0
CCl3 -0.009 0.001 -0.011 -0.003 0.0 -0.004
CF3 -0.009 0.001 -0.011 -0.008 0.001 -0.006
CN -0.016 0.003 -0.016 -0.010 0.002 -0.009
CHO -0.018 0.008 -0.017 -0.020 0.004 -0.011
NO2 -0.032 0.009 -0.025 -0.030 0.006 -0.015

partition the molecular space in two parts: on one hand, the Figure 4. σpara - σ′ vs RIpara(S): 9 Hartree-Fock, 1 Becke3LYP.
subspace formed by the basins which only belong to the
substituent S and, on the other hand, the remaining space
involving the phenyl group basins. The V(C, S) disynaptic basin
linking the substituent to the phenyl skeleton belongs to the
latter subspace. In the phenyl subspace the number and the type
of the basin are identical, therefore all the ELF gradient fields
are isomorphic. This means that the ELF gradient field within
the phenyl subspace is structurally stable with respect to the
substitution which can be consequently considered as a weak
perturbation. Moreover, the isomorphism implies that the critical
points located in the phenyl subspace are topological invariants.
The phenyl group localization domains as displayed in Figure
3 show two kinds of bifurcation diagram according to the nature
of the substituent. For ortho-para orienting substituents, namely
S ) NH2, OH, F, Cl, CH3, and C6H5, the first splitting of the
aromatic domain occurs at the level of the carbons in meta posi-
tion whereas for meta orienting substituents as S ) CCl3, CF3,
CN, CHO, and NO2 it occurs at the level of the ortho and para Figure 5. σmeta - σ′ vs RImeta(S): 9 Hartree-Fock, 1 Becke3LYP.
carbons. From a qualitative point of view, the graphical represen-
tation clearly indicates which carbons are the reactive sites. In made by examining the substituent localization domains of the
turn, the information provided by Figure 3 can be expressed in substituents in Figure 3. The cores of the most electronegative
terms of numbers by considering the values η(Ci; S) of the atoms are always surrounded by a single reducible localization
localization function at the (3, -1) critical points located on domain which reproduces the atomic valence shell. Of course,
the separatrices of the V(Ci-1, Ci) and V(Ci, Ci+1) basins of the this domain gives rise to irreducible localization domains at
S-substituted derivative. These values rule the bifurcation higher values of the ELF function. The larger the electronega-
diagram of the aromatic domain and are expected to enable a tivity, the higher will be the ELF value enabling the corre-
quantitative comparison of the effects of the substituents. Instead sponding splitting of the domain. The ELF value at the (3, -1)
of the η(Ci) themselves it is more convenient to introduce critical point corresponding to the complete reduction of
electrophilic substitution positional indices defined as localization around a given center is a measure of its electrone-
gativity. The electrophilic substitution positional indices are
RIc(S) ) η(Ci; S) - η(Ci; H) (4) therefore related to the in situ electronegativity of the carbons.
The effect of substituent on meta and para electrophilic
in which the subscript c denotes the position of the carbon substitutions has been phenomenologically rationalized by the
labeled by i, i.e., ortho, meta, or para. The electrophilic Hammett equation48 which expresses the rates or the equilibria
substitution positional index is just the difference of the ELF in terms of two parameters, σ and F. The former characterizes
values at equivalent topological invariants in the S-derivative the substituent and the site where the reaction takes place; the
and in benzene. They are local measures of the perturbation of latter depends upon the nature of the reaction and the conditions.
the electron localization function by the substituent. The values of σ which account for the reactivity of the para
The electrophilic substitution positional indices calculated at and meta positions hamper its direct use to establish a positional
the 6-31G**/HF and Becke3LYP levels are listed in Table 1. hierarchy of the substituents; for example, though NH2 and F
As a general rule, for a given species the ortho and para indices are both para-director, their σpara constants have opposite signs
have the same sign and the meta index has the opposite sign. A because the substitution is easier for aniline than for fluoroben-
positive index betokens a favored electrophilic substitution zene. This difficulty can be removed either by considering the
position. Though there are differences between the Hartree- differences (σmeta - σpara) or by the introduction of the so-called
Fock and Becke3LYP values for corresponding indices, the two polar substituent constant σ′ 49 which is used to form (σpara -
approaches yield the same order with respect to the nature of σ′) and (σmeta - σ′). Figures 4 and 5 compare our reactivity
the substituent. The chemical interpretation of the electrophilic indices RIpara and RImeta to these latter differences. A rather
substitution positional indices follows a remark which can be satisfactory linear correlation is observed which enables the
856 J. Phys. Chem. A, Vol. 104, No. 4, 2000 Fuster et al.

Figure 6. σmeta - σpara vs RIpara(S) - RImeta(S): 9 Hartree-Fock, 1


Becke3LYP.

Figure 8. Deformation electron density isosurfaces ∆F ) (0.0005.


Full volume electron density gain, wire frame electron density loss:
(a) aniline φ ) 0.0, (b) aniline φ ) 90.0, (c) toluene. This figure,
presented here in black and white, is available in color on the World
Wide Web in Supporting Information.

RIortho might be among the causes of the breakdown of the


Hammett equation for ortho substitution. In the case of aniline
the V(N) contribution appears to be the dominant one at the
equilibrium geometry (φ ) 0°), that due to the V(C, N) basin
being very weak. In the case of toluene, the η(Ci) are
independent of the rotation of the methyl group. This follows
from the specific forms of the η(Ci)’s independently of the
strength of the hyperconjugacy:

Figure 7. Angular dependence η(Ci) vs φ for aniline: full line, ortho; cos2 φ + cos2(φ + 120) + cos2(φ - 120) ) 3/2 (7)
dashed line, para; dotted line, meta.
cos4 φ + cos4(φ + 120) + cos4(φ - 120) ) 9/4 (8)
proposal of the following approximate relationships: The electron transfers between the substituent and the phenyl
group are evidenced by the electron density deformation (F-
σpara - σ′ ≈ -40.2RIpara (5) (C6H5S) - F(C6H6)). Figure 8 displays the isosurfaces ∆F )
σmeta - σ′ ≈ -17.8RIpara (6) (0.0005 e- bohr-3 for aniline (φ ) 0, 90.0) and for toluene.
For aniline the largest transfers occur at φ ) 0. As expected
Finally, the behavior of (σpara - σmeta) with respect to (RIpara from molecular orbital theory arguments it mostly involves the
- RImeta), Figure 6, looks very similar to the rate constant curves π-system with losses on the meta carbons and gains on the ortho
of the solvolysis of aryl-propyl tosylates discussed by Raber and para ones. Each corresponding isosurface is made of two
et al.50 lobes centered on a carbon which strongly suggests the density
In the standard molecular orbital framework, the substituent of the 2pz atomic orbital. The σ-transfer is weaker and in phase
effect is the sum of two contributions, one arising from the opposition: gains in meta, losses in ortho-para. When the lone
π-system the other from σ-orbitals. In our approach, valence pair axis is in the ring plane, the size of all lobes is decreased
basin contributions are considered instead of molecular orbital with respect to the φ ) 0 situation. The remaining π-transfers
contributions. The importance of the lone pair basin of the are due to hyperconjugacy whereas the weakening of the σ one
substituent can be discussed in the case of aniline which has might be due to the participation of the lone pair. The toluene
one lone pair borne by the nitrogen atom. Figure 7 shows the case unequivocally illustrates the hyperconjugacy effect.
variation of η(Ci) for the ortho, meta, and para position as a
5. Polysubstituted and Polyaromatic Derivatives
function of the dihedral angle φ made by the plane containing
the C-N bond and the V(N) attractor with a plane perpendicular In the previous section it has been shown that the substitution
to the ring. The three curves have a periodicity of 180°. The of a hydrogen by another functional group weakly perturbs the
para and meta curves have their maximal amplitudes at φ ) localization gradient field of the phenyl group. The electrophilic
0° whereas as the two ortho positions are not equivalent because substitution positional indices are the linear response coefficient
the maxima are at φ ∼ (10°. The decomposition in terms of of the ELF function at the (3, -1) critical points of the aromatic
even powers of cos φ shows that the meta and para curves domain. The addition of another substituent is expected to also
behave essentially as cos2 φ and the ortho one as cos4(φ + be a weak perturbation for which the linear response ap-
10). This different behavior of the ortho η(Ci) and therefore of proximation holds. The reactivity indices of a polysubstituted
Analysis of the Electron Localization Function J. Phys. Chem. A, Vol. 104, No. 4, 2000 857

TABLE 2: Calculated and Estimated Electrophilic TABLE 4: Additivity of the Electrophilic Substitution
Substitution Positional Indices of Polysubstituted Benzenes Positional Indices of Fluoro Naphthalenes
molecule c1 c2 S1 S2 calc est 1,4-C10H6F2
o-C6H4F2 meta ortho F F 0.012 0.013 position 1-C10H7F 4-C10H7F calc est
meta para F F 0.003 0.003
C1 0.002
para meta F F 0.003 0.003
C3 -0.007 0.026 0.020 0.019
ortho meta F F 0.012 0.013
C4 -0.001
m-C6H4F2 ortho ortho F F 0.039 0.038
C5 -0.003 0.012 0.010 0.009
para ortho F F 0.028 0.028
C6 -0.001 -0.005 -0.006 -0.006
meta meta F F -0.012 -0.012
C8 0.005 -0.002 0.003 0.003
ortho para F F 0.028 0.028
C9 -0.005 0.001 -0.004 -0.004
p-C6H4F2 ortho meta F F 0.013 0.013
C10 0.024 -0.004 0.027 0.020
m-C6H4(CH3)2 ortho ortho CH3 CH3 0.027 0.026
ortho para CH3 CH3 0.023 0.023
meta meta CH3 CH3 -0.013 -0.014 cycle itself acts as a substituent. Nonetheless, a qualitative
para ortho CH3 CH3 0.023 0.023 reasoning holds. The values of Table 1 show that for ortho/
o-C6H4OHCl ortho meta OH Cl 0.029 0.031 para director substituents, the ortho electrophilic substitution
meta para OH Cl -0.020 -0.018 positional index has always the largest absolute value. For the
meta ortho OH Cl -0.015 -0.015
para meta OH Cl 0.021 0.020 C3 carbon the total contribution of the other cycle is of the form
RIortho + RImeta and larger than that of C4, namely RIpara + RImeta.
TABLE 3: Value of the ELF Function at the (3, -1) The same arguments hold to explain why C3 is also the reactive
Critical Points of the Aromatic Domains of the Naphthalene center in anthracene.
and Anthracene Molecules For substituted and polysubstituted polyaromatic molecules
position naphthalene anthracene it is convenient to generalize the electrophilic substitution
C3 0.6704 0.6836 positional index concept. The definition is the same as previ-
C4 0.6674 ously cited (see eq 4), the η(Ci; H) being the values of the
C5 0.6674 0.6691 hydrogenated compound. Table 4 reports the calculated values
C6 0.6704 0.6650 of these indices for the 1- and 4-fluoro naphthalenes and for
the 1,4-difluoro naphthalene. In the monosubstituted molecules,
benzene should therefore be a sum of monosubstituted derivative
the largest indices correspond to the ortho-like position within
indices:
a cycle; they are of the same order of magnitude as fluoro
RIc1,c2,...(S1,S2,...) ≈ RIc1(S1) + RIc2(S2) + ... (9) benzene: slightly larger in 1-fluoro naphthalene in which there
is only one ortho-like position whereas in 4-fluoro naphthalene
the average of the two ortho-like indices is the ortho value of
The notation is consistent with that used eq 4 for monosubsti-
fluoro benzene. The transferability of the fluoro-benzene indices
tuted derivatives: the successive ci indices are the positions with
to fluoro naphthalene is very good for the meta positions. The
respect to the successive substituents Si. The additivity of the
fulfillment of the additivity property is verified by the estimated
electrophilic substitution positional indices has been verified
values given in the last column of Table 4.
on the three difluoro-benzene isomers, on the meta dimethyl
benzene, and on the ortho chloro phenol. Table 2 compares the 6. Conclusion
electrophilic substitution positional indices of these molecules The results presented in this report show that the topological
explicitly calculated on the disubstituted molecules themselves analysis of the electron localization function is a sound basis
and estimated by eq 9 with the values of Table 1. There is an for the determination of the electrophilic substitution site and,
excellent agreement between the two series of values. The more generally, for the localization of the reactive sites in
additivity of the electrophilic substitution positional indices is molecules. The ELF analysis provides ready-for-use qualitative
consistent with the correlation with the Hammett constants since visual information through the graphical representation of the
these latter are also additive. localization domains. Quantitatively, the electrophilic substitu-
The generalization of the method to polyaromatic molecules tion positional indices make a link between the results of the
is straightforward. The carbon labels of naphthalene and calculation and the Hammett constants. As these latter, they
anthracene are as follows: are additive, which makes possible reliable predictions on a large
number of polysubstituted derivatives without explicit calcula-
tions of their wave functions. This method complements the
description of the reactivity already provided by the frontier
orbital and the atoms in molecules theories. The use of the ELF
function emphasizes the importance of the local Pauli repulsion
as a token of local electronegativity.
The ELF values at the (3, -1) critical points of the aromatic Acknowledgment. The authors gratefully acknowledge Prof.
domains are listed in Table 3. In agreement with experiment, J. Molina’s comments on the draft manuscript. The data analyzer
the electrophilic substitution center of naphthalene is the carbon software SciAn51 has been used to produce Figures 2, 3, and 8.
C3. Each cycle of the naphthalene can be considered as a
Supporting Information Available: The color version of
disubstituted benzene on the carbon C1 and C6; however, eq 9
Figures 2, 3, and 8. This material is available free of charge
cannot legitimately be applied to estimate the ELF values at
via the Internet at http://pubs.acs.org.
the critical points because their deviations from the benzene
value are a sum of two contributions: one arising from the other References and Notes
cycle which is accounted for by the electrophilic substitution (1) Holleman, A. F. Rec. TraV. Chim. Pays-Bas 1906, 24, 140.
positional indices, the other one being due to the fact that each (2) Holleman, A. F. Chem. ReV. 1925, 1, 187.
858 J. Phys. Chem. A, Vol. 104, No. 4, 2000 Fuster et al.

(3) Holleman, A. F.; Wibaut, J. P. Organic Chemistry, a translation (31) Frisch, M. J.; Pople, J. A.; Binkley, J. S. J. Chem. Phys. 1984, 80,
from the sixteenth deutch edition; Elsevier: Amsterdam, 1951. 3265.
(4) Norman, R. O. C. Electrophilic substitution in Benzenoid com- (32) Becke, A. D. J. Chem. Phys. 1993, 98, 5648.
pounds; Elsevier: Amsterdam, 1965. (33) Becke, A. D. Phys. ReV. 1988, A38, 3098.
(5) Hückel, E. Z. Phys. 1931, 70, 204. (34) Lee, C.; Yang, Y.; Parr, R. G. Phys. ReV. 1988, B37, 785.
(6) Hückel, E. Z. Phys. 1931, 72, 310. (35) Miechlich, B.; Savin, A.; Stoll, H.; Preuss, H. Chem. Phys. Lett.
(7) Hückel, E. Z. Phys. 1932, 76, 628. 1989, 157, 200.
(8) Wheland, G. W. Resonance in Organic Chemistry; Wiley: New (36) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Gill, P. M. W.;
York, 1955. Johnson, B. G.; Robb, M. A.; Cheeseman, J. R.; Keith, T.; Petersson, G.
(9) Ingold, C. K. Structure and Mechanisms in Organic Chemistry; A.; Montgomery, J. A.; Raghavachari, K.; Al-Laham, M. A.; Zakrzewski,
Bell and Sons: London, 1953. V. G.; Ortiz, J. V.; Foresman, J. B.; Cioslowski, J.; Stefanov, B. B.;
(10) Coulson, C. A. Valence; Clarendon: Oxford, 1952. Nanayakkara, A.; Challacombe, M.; Peng, C. Y.; Ayala, P. Y.; Chen, W.;
(11) Streitweiser, A., Jr. Molecular Orbital Theory for Organic Chemists; Wong, M. W.; Andres, J. L.; Replogle, E. S.; Gomperts, R.; Martin, R. L.;
Wiley: New York, 1961. Fox, D. J.; Binkley, J. S.; Defrees, D. J.; Baker, J.; Stewart, J. P.; Head-
(12) Salem, L. The Molecular Orbital Theory of Conjugated Systems; Gordon, M.; Gonzalez, C.; Pople, J. A. Gaussian 94, ReVision D.4; Gaussian
W. A. Benjamin Inc.: Reading, 1966. Inc.: Pittsburgh, PA, 1995.
(13) Dewar, M. J. S. The Molecular Orbital Theory of Organic (37) Alikhani, M. E.; Bouteiller, Y.; Silvi, B. J. Phys. Chem. 1996, 100,
Chemistry; McGraw-Hill: New York, 1968. 16092.
(14) Epiotis, N. D. J. Am. Chem. Soc. 1973, 95, 3188. (38) Fourré, I.; Silvi, B.; Chaquin, P.; Sevin, A. J. Comput. Chem. 1999,
(15) Epiotis, N. D.; Cherry, W. J. Am. Chem. Soc. 1976, 98, 4361. 20, 897.
(16) Epiotis, N. D.; Cherry, W. J. Am. Chem. Soc. 1976, 98, 4365. (39) Llusar, R.; Beltrán, A.; Andrés, J.; Noury, S.; Silvi, B. J. Comput.
(17) Epiotis, N. D.; Cherry, W. J. Am. Chem. Soc. 1976, 98, 5432. Chem. 1999, 20, 1517.
(18) Bader, R. F. W. Atoms in Molecules: A Quantum Theory; Oxford (40) Beltrán, A.; Andrés, J.; Noury, S.; Silvi, B. J. Phys. Chem. A 1999,
Univ. Press: Oxford, 1990. 103, 3078.
(19) Bader, R. F. W.; Chang, C. J. Phys. Chem. 1989, 93, 2946. (41) Berski, S.; Silvi, B.; Latajka, Z.; Leszczynski, J. J. Chem. Phys.
(20) Bader, R. F. W.; Chang, C. J. Phys. Chem. 1989, 93, 5095. 1999, 111, 2542.
(21) Cubero, E.; Orozco, M.; Luque, F. J. J. Phys. Chem. A 1999, 103, (42) Krokidis, X.; Noury, S.; Silvi, B. J. Phys. Chem. A 1997, 101, 7277.
315. (43) Krokidis, X.; Goncalves, V.; Savin, A.; Silvi, B. J. Phys. Chem. A
(22) Becke, A. D.; Edgecombe, K. E. J. Chem. Phys. 1990, 92, 5397. 1998, 102, 5065.
(23) Savin, A.; Becke, A. D.; Flad, J.; Nesper, R.; Preuss, H.; von (44) Krokidis, X.; Silvi, B.; Alikhani, M. E. Chem. Phys. Lett. 1998,
Schnering, H. G. Angew. Chem., Int. Ed. Engl. 1991, 30, 409. 292, 35.
(24) Savin, A.; Jepsen, O.; Flad, J.; Andersen, O. K.; Preuss, H.; von (45) Krokidis, X.; Vuilleumier, R.; Borgis, D.; Silvi, B. Mol. Phys. 1999,
Schnering, H. G. Angew. Chem., Int. Ed. Engl. 1992, 31, 187. 96, 265.
(25) Silvi, B.; Savin, A. Nature 1994, 371, 683. (46) Noury, S.; Krokidis, X.; Fuster, F.; Silvi, B. Topmod package,
(26) Savin, A.; Silvi, B.; Colonna, F. Can. J. Chem. 1996, 74, 1088. 1997.
(27) Savin, A.; Nesper, R.; Wengert, S.; Fässler, T. F. Angew. Chem., (47) von Weizsácker, C. F. Z. Phys. 1935, 96, 431.
Int. Ed. Engl. 1997, 36, 1809. (48) Hammett, L. P. Chem. ReV. 1935, 35, 125.
(28) Noury, S.; Colonna, F.; Savin, A.; Silvi, B. J. Mol. Struct. 1998, (49) Roberts, J. D.; Moreland, W. T. J. Am. Chem. Soc. 1953, 75, 2167.
450, 59. (50) Raber, D. J.; Harris, J. M.; Schleyer, P. v. R. J. Am. Chem. Soc.
(29) Hehre, W. J.; Ditchfield, R.; Pople, J. A. J. Chem. Phys. 1972, 56, 1971, 93, 4829.
2257. (51) Pepke, E.; Murray, J.; Lyons, J.; Hwu, T.-Z. Scian; Supercomputer
(30) Clark, T.; Chandrasekhar, J.; Spitznagel, G. W.; Schleyer, P. v. R. Computations Research Institute; Florida State University: Tallahassee, FL,
J. Comput. Chem. 1983, 4, 294. 1993.

You might also like