07.01.2015 Views

Linear Response and Measures of Electron Delocalization ... - CRM2

Linear Response and Measures of Electron Delocalization ... - CRM2

Linear Response and Measures of Electron Delocalization ... - CRM2

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Current Organic Chemistry, 2011, 15, 3609-3618 3609<br />

<strong>Linear</strong> <strong>Response</strong> <strong>and</strong> <strong>Measures</strong> <strong>of</strong> <strong>Electron</strong> <strong>Delocalization</strong> in Molecules<br />

János G. Ángyán*<br />

Équipe Modélisation Quantique et Cristallographique, <strong>CRM2</strong>, UMR 7036, Institut Jean Barriol, Nancy-University, CNRS, B.P. 239,<br />

F-54506 V<strong>and</strong>oeuvre-lès-Nancy, France<br />

Abstract: The concept <strong>of</strong> localization <strong>and</strong> delocalization in molecules is discussed in terms <strong>of</strong> the response <strong>of</strong> the electronic system to an<br />

external perturbation. It is argued that both the spatial organization <strong>of</strong> electrons in pairs <strong>and</strong> the spatial distribution <strong>of</strong> the response<br />

intensity, reflect main features <strong>of</strong> the correlated motion <strong>of</strong> electrons, ultimately described by the pair distribution function <strong>of</strong> electrons.<br />

Various measures, derived from the linear charge density response function, are able to characterize localization in a rigorous way, in<br />

close analogy to the approach followed in solid state physics.<br />

Keywords: <strong>Electron</strong> localization, linear response, Resta localization index, bond order, localized orbitals, exchange hole, distributed<br />

polarizabilities.<br />

1. INTRODUCTION<br />

<strong>Electron</strong> localization/delocalization has <strong>of</strong>ten been a<br />

controversial subject in chemistry <strong>and</strong> physics, perhaps because this<br />

concept is sometimes used only in a relatively loose, qualitative<br />

sense in the chemical literature. However, theoretical chemists have<br />

always attributed a great importance to a better underst<strong>and</strong>ing <strong>of</strong><br />

this notion <strong>and</strong> from the early days <strong>of</strong> electronic structure theory a<br />

considerable effort has been spent to extract precise pieces <strong>of</strong><br />

information about the localization <strong>of</strong> electrons from the wave<br />

function. This subject has been reviewed abundantly in 2005, in a<br />

series <strong>of</strong> papers <strong>of</strong> a special issue <strong>of</strong> the Chemical Reviews, see e.g.<br />

Refs [1, 2]. Other interesting review papers [3, 4] <strong>and</strong> critical<br />

accounts [5, 6] were published in the same year. One should also<br />

mention the Special issue "90 years <strong>of</strong> chemical bonding" <strong>of</strong> the<br />

Journal <strong>of</strong> Computational Chemistry, with several contributions<br />

closely related to the subject <strong>of</strong> the present article, e.g. [7-10]. The<br />

field continues to develop rapidly <strong>and</strong> a considerable number <strong>of</strong><br />

new ideas about the characterization <strong>of</strong> electron localization has<br />

appeared in the past few years.<br />

The purpose <strong>of</strong> the present contribution is not to provide an<br />

exhaustive review <strong>of</strong> these different developments, rather to give a<br />

somewhat specific view <strong>of</strong> the problem <strong>of</strong> electron localization,<br />

which is probably less familiar to the chemical community. It will<br />

be argued that there is a straight relationship between the localized<br />

or delocalized nature <strong>of</strong> an electronic system <strong>and</strong> its response<br />

properties, in particular its charge density response properties.<br />

Although this viewpoint seems to be a widely accepted one in the<br />

solid state physics community, following mainly the pioneering<br />

work <strong>of</strong> Resta <strong>and</strong> others [11-13], according to the author's opinion<br />

it has still not received the deserved audience among chemists.<br />

Resta's overview paper [14], addressed specifically to the chemists'<br />

community, written in a clear <strong>and</strong> accessible style, seems to have<br />

remained practically without any echo. It can be hoped that this<br />

situation will change in the future <strong>and</strong> the usefulness <strong>of</strong> this<br />

viewpoint, stressing the links between the theory <strong>of</strong> electric<br />

polarization <strong>and</strong> electron localization/delocalization, will be<br />

recognized in the world <strong>of</strong> finite molecular systems.<br />

*Address correspondence to this author at the Équipe Modélisation Quantique et<br />

Cristallographique, <strong>CRM2</strong>, UMR 7036, Institut Jean Barriol, Nancy-University,<br />

CNRS, B.P. 239, F-54506 V<strong>and</strong>oeuvre-lès-Nancy, France; Tel: +33 3 83 68 48 74;<br />

Fax: +33 3 83 68 43 00; E-mail: Janos.Angyan@crm2.uhp-nancy.fr<br />

A measure <strong>of</strong> central importance, introduced by Resta, is the<br />

electron localization tensor, defined as the fluctuation <strong>of</strong> the<br />

electron position operator or, in more mathematical terms, the<br />

second cumulant moment <strong>of</strong> the electron distribution. This quantity<br />

is related to the imaginary part <strong>of</strong> the frequency-dependent dipole<br />

polarizability tensor by a well-known sum rule, the zerotemperature<br />

form <strong>of</strong> the fluctuation-dissipation theorem. The<br />

underlying physics is quite simple: an electronic system which<br />

undergoes strong spontaneous quantum fluctuations is, par<br />

excellence, delocalized. In the mean time, such a system is ready to<br />

respond to external perturbations, in other words, it is strongly<br />

polarizable. Moreover, as we shall see, not only the intensity <strong>of</strong><br />

response <strong>and</strong> the delocalization <strong>of</strong> electrons can be put in parallel,<br />

but the spatial distribution <strong>of</strong> the response <strong>and</strong> the organization <strong>of</strong><br />

electrons in opposite-spin pairs are also expected to follow similar<br />

trends. A further remarkable point is that the electronic<br />

polarizability is, in principle, an experimentally accessible physical<br />

property, therefore it might be possible to characterize the<br />

localizability <strong>of</strong> electrons on experimental grounds.<br />

The localization tensor is only one <strong>of</strong> the possible polarizationbased<br />

measures <strong>of</strong> localizability. As it will be shown, other<br />

quantities, like indices characterizing delocalization between<br />

domains as well as some variants <strong>of</strong> pointwise electron localization<br />

functions can be related to the linear response function <strong>of</strong> the<br />

electronic system. The present discussion will be focused on these<br />

relatively disparate subjects linked together by their relationship to<br />

the linear response properties <strong>of</strong> the electronic system.<br />

It is important to stress that the linear response viewpoint is not<br />

contradictory to the common interpretation <strong>of</strong> electron localization<br />

in terms <strong>of</strong> the Pauli exclusion principle <strong>and</strong> <strong>of</strong> the properties <strong>of</strong> the<br />

Fermi-hole. Since the linear response is related to the paircorrelation<br />

<strong>of</strong> electrons, which is strongly dominated by the Fermicorrelation,<br />

consequence <strong>of</strong> the Pauli principle, it is natural that<br />

organization <strong>of</strong> electrons in pairs be also reflected, in a certain way,<br />

by the response properties.<br />

The paper is organized as follows. In Section 2 the linear charge<br />

density response function <strong>and</strong> the fluctuation-dissipation theorem is<br />

introduced, while in Section 3 Resta's localization tensor is derived<br />

therefrom. In Section 4, it is shown how delocalization/localization<br />

indices are related to the charge-flow polarizabilities between<br />

appropriately selected regions. The relationship <strong>of</strong> some orbital<br />

localization criteria to the polarizability is briefly discussed in<br />

1385-2728/11 $58.00+.00 © 2011 Bentham Science Publishers


3610 Current Organic Chemistry, 2011, Vol. 15, No. 20 János G. Ángyán<br />

Section 5. In Section 6, the local measures <strong>of</strong> the electron<br />

localization are overviewed from the polarizability viewpoint, <strong>and</strong><br />

finally we conclude by summarizing the main points <strong>of</strong> the present<br />

contribution.<br />

2. FROM THE POLARIZABILITY TO THE EXCHANGE-<br />

CORRELATION HOLE<br />

The response <strong>of</strong> an electronic system to an external, possible<br />

time-dependent, electric perturbation is characterized by a linear<br />

response function, (r, r ;) , which can be regarded as a<br />

generalization <strong>of</strong> its polarizability:<br />

(r;)= dr(r, r ;)V( r ;), . (1)<br />

where V( r ;) is a perturbing potential at point r oscillating by a<br />

frequency , (r;) is the induced charge density at r <strong>and</strong> the<br />

linear charge density response function is defined as<br />

(r,r ' ;)= lim 1<br />

0 + n0<br />

<br />

0| ˆ(r)|nn | ˆ(r' )|0<br />

<br />

+ i 0n<br />

0| ˆ(r' )|nn | ˆ(r)|0<br />

i + 0n<br />

<br />

<br />

. (2)<br />

The zero-temperature fluctuation-dissipation theorem<br />

establishes a relationship between the imaginary part <strong>of</strong> the charge<br />

density response function <strong>and</strong> the charge density fluctuation<br />

autocorrelation function:<br />

<br />

<br />

d m(r, r ;)=S(r, r ). (3)<br />

0<br />

The charge density autocorrelation function, sometimes called<br />

static form factor, is a ground state expectation value,<br />

S(r, r )=0| ˆ(r) r )|0, (4)<br />

where ˆ(r)= ˆ(r)(r) is the charge density fluctuation operator<br />

taken with respect to the mean ground state density, (r) . The<br />

function S(r, r ) is closely related to the exchange-correlation hole,<br />

h xc<br />

(r, r )= 2<br />

(r, r )/(r)( r ) :<br />

S(r, r )=(r)h xc<br />

(r, r )+ (r)(r r ), (5)<br />

as it can be easily proven by considering the definition <strong>of</strong> the pair<br />

density operator, ˆ 2<br />

(r, r )= ˆ(r) ˆ( r )ˆ(r)(r r ) , <strong>and</strong> inserting it in<br />

the definition <strong>of</strong> the form factor.<br />

An approximate relationship can be established between the xchole<br />

density <strong>and</strong> the static charge density response,<br />

(r, r )=(r, r ;0). Let us define a position dependent parameter<br />

(r, r ) such that<br />

(r, r )= 2 <br />

S(r, r )<br />

(r, r ) . (6)<br />

(r, r ) plays the role <strong>of</strong> an average energy denominator for a pair<br />

<strong>of</strong> coordinates <strong>and</strong> allows us to express the frequency dependent<br />

polarizability in a single-pole form, c.f. Eq. (7) <strong>of</strong> Ref. [15]. If one<br />

neglects the spatial dependence <strong>of</strong> (r, r ) by replacing it by a<br />

space average, (r, r ) , an approximate proportionality can be<br />

established between the static polarizability <strong>and</strong> the xc-hole density:<br />

(r, r ) (r)h xc<br />

(r, r )+ (r)(r r ). (7)<br />

The qualitative meaning <strong>of</strong> the above relationship is that in a<br />

point r the amplitude <strong>of</strong> the electronic charge density response to<br />

an external perturbation at r can be predicted from the knowledge<br />

<strong>of</strong> the exchange-correlation hole belonging to the reference point<br />

r . The expected response will be different from zero only in those<br />

points <strong>of</strong> the space where the xc-hole is non-vanishing. In other<br />

words the polarization <strong>of</strong> the electrons remains confined within the<br />

space occupied by the xc-hole. If the xc-hole extends over the<br />

whole molecular framework, the electron is strongly delocalized.<br />

This would be a finite, molecular analog <strong>of</strong> the conducting state in<br />

an infinite solid. On the contrary, if the xc-hole occupies only a<br />

limited portion <strong>of</strong> the molecular space, one should consider the<br />

electron as localized. Therefore, as far as one underst<strong>and</strong>s the<br />

spatial distribution <strong>of</strong> the xc-hole, one has a key to predict the<br />

charge density response <strong>and</strong> vice versa. Both <strong>of</strong> these quantities,<br />

response function <strong>and</strong> Fermi-hole, provide appropriate bases for our<br />

underst<strong>and</strong>ing <strong>of</strong> electron localization.<br />

A related example for using a mapping between response<br />

functions <strong>and</strong> the Fermi-hole density has been given by the<br />

interpretation <strong>of</strong> the <strong>of</strong> Fermi-contact contributions to the nuclear<br />

spin-spin coupling constants [16-19]. Visualization <strong>of</strong> the response<br />

spin densities <strong>and</strong> <strong>of</strong> the Fermi-hole belonging to a reference<br />

electron placed at an atomic position has convincingly<br />

demonstrated the relationship between these two quantities.<br />

Furthermore, a proportionality has been found between the spinspin<br />

coupling constants <strong>and</strong> the delocalization index between the<br />

atom pair, measuring the sharing <strong>of</strong> the Fermi-hole between the<br />

atomic volumes [16].<br />

The charge density polarizability function as well as the xc-hole<br />

are complicated two-variable functions, which are not really<br />

appropriate to obtain a quick insight <strong>and</strong> underst<strong>and</strong>ing. Further<br />

analysis <strong>of</strong> these functions is needed either in terms <strong>of</strong> integrated<br />

quantities or by using a kind <strong>of</strong> coarse-grain representation, e.g. by<br />

space domains or other kind <strong>of</strong> partitions <strong>of</strong> the molecule. In the<br />

following sections it will be shown how these functions are related<br />

to various simpler measures <strong>of</strong> electron localizability.<br />

3. RESTA LOCALIZATION TENSOR<br />

A useful global measure <strong>of</strong> the electron localizability has been<br />

defined by Resta [12-14], via the second cumulant moment <strong>of</strong> the<br />

electron distribution. It can be easily shown that the following<br />

expectation value, describing the electron position fluctuation per<br />

electron,<br />

N<br />

ˆr ˆr <br />

c<br />

= 1 N {0| ˆr <br />

(i)ˆr <br />

( j)|00| ˆr <br />

(i)|00| ˆr <br />

( j)|0}, (8)<br />

i, j=1<br />

i=1<br />

j=1<br />

is strictly equivalent to the second moment <strong>of</strong> the static form factor<br />

normalized to the number <strong>of</strong> electrons, N :<br />

N<br />

ˆr ˆr <br />

c<br />

= 1 '<br />

drdrr <br />

r <br />

S(r, r ). (9)<br />

N<br />

In the above equations ˆr <br />

(i) the = x, y,z component <strong>of</strong> the<br />

i -th electronic position vector. We can see that Eq. (9) establishes<br />

a direct link between fundamental two-variable functions discussed<br />

in the preceding section. The intensive quantity, ˆr ˆr <br />

c<br />

, is called<br />

also second cumulant moment per electron, or localization tensor<br />

[14].<br />

An explicitly origin-independent equivalent form <strong>of</strong> the<br />

localization tensor can be obtained by using the chargeconservation<br />

sum rule <strong>of</strong> the form factor:<br />

ˆr ˆr <br />

c<br />

= 1 dr dr (r r ) <br />

(r r') <br />

S(r, r ), (10)<br />

2N<br />

N


<strong>Linear</strong> <strong>Response</strong> <strong>and</strong> <strong>Measures</strong> <strong>of</strong> <strong>Electron</strong> <strong>Delocalization</strong> Current Organic Chemistry, 2011, Vol. 15, No. 20 3611<br />

Table 1. Selected "Experimental" Resta Localization Indices, r 2 exp c<br />

= 3 S DODS 2 N 1<br />

, from DOSD Data. The Systems are Listed in the Order <strong>of</strong> Increasing<br />

Localization Index. The Degree <strong>of</strong> <strong>Delocalization</strong>, as Measured by this Quantity, is Supposed to Increase from the the Top to the Bottom <strong>of</strong><br />

the Table<br />

Molecule<br />

r 2 c<br />

exp<br />

Ref. Molecule r 2 c<br />

exp<br />

Ref.<br />

Ne 0.5702 [22] H 2CO 1.1766 [21]<br />

SF 6 0.6739 [23] SiH 4 1.3233 [24]<br />

SiF 4 0.6951 [25] CH 3OCH 3 1.3321 [26]<br />

Ar 0.7307 [22] C 2H 5OC 2H 5 1.3900 [26]<br />

Cl 2 0.8523 [27] CH 3OC 3H 7 1.3975 [26]<br />

N 2 1.0161 [28] NH 3 1.3978 [28]<br />

CO 1.0422 [29] C 2H 2 1.4357 [30]<br />

H 2S 1.0942 [25] C 2H 4 1.5290 [31]<br />

H 2O 1.0974 [29] C 2H6 1.5625 [31]<br />

He 1.1168 [21] H 2 2.3250 [28]<br />

<strong>and</strong> Eq. (10) can be further transformed to a form, which is easy to<br />

interpret as the density-weighted second moment <strong>of</strong> the xc-hole<br />

function:<br />

ˆr ˆr <br />

c<br />

= 1 dr(r)dr (r r ) <br />

(r r') <br />

h xc<br />

(r, r ), (11)<br />

2N<br />

establishing a relationship between the xc-hole (dominated by the<br />

exchange- or Fermi-hole) <strong>and</strong> the Resta localization tensor. Clearly,<br />

<strong>and</strong> xc-hole with large second moment components can be<br />

considered as an indication for the presence <strong>of</strong> strongly delocalized<br />

electrons.<br />

The expression <strong>of</strong> the second cumulant moment in terms <strong>of</strong> the<br />

frequency-dependent dipole polarizability tensor, <br />

, is<br />

essentially the same as that <strong>of</strong> the S 1<br />

, the –1 order dipole oscillator<br />

strength sum rule (see e.g. Ref. [20]). Comparing<br />

<br />

dm<br />

<br />

0 <br />

()=N r <br />

r <br />

c<br />

(12)<br />

with the usual form <strong>of</strong> the sum rule, one finds a simple<br />

proportionality between dipole oscillator strength sum rule S 1<br />

<strong>and</strong><br />

the trace <strong>of</strong> the localization tensor, which can be called Resta<br />

localization index, r 2 c<br />

: r 2 c<br />

= 3<br />

2N S 1<br />

. Thus it can be expected<br />

that the spherical average <strong>of</strong> the localization tensor is an<br />

information accessible from experimental data, e.g. via the dipole<br />

oscillator strength distributions (DOSD) reported in the literature.<br />

The existence <strong>of</strong> such a relationship is quite remarkable, because it<br />

makes possible an experimental characterization <strong>of</strong> a fundamentally<br />

conceptual quantity, the electron localizability. A set <strong>of</strong> illustrative<br />

examples is presented in Table 1, partly relying on a compilation <strong>of</strong><br />

experimental data by Olney [21] <strong>and</strong> using also some more recent<br />

DOSD determinations.<br />

The S 1<br />

oscillator strength sum rule, <strong>and</strong> therefore the trace <strong>of</strong><br />

the localization tensor, is connected to the second moment <strong>of</strong> the<br />

intracule <strong>and</strong> extracule densities by rigorous relationships, which<br />

has been derived for the case <strong>of</strong> many-electron atoms [32]. It can be<br />

shown also that S 1<br />

(<strong>and</strong> the trace <strong>of</strong> the localization tensor) is<br />

proportional to the the negative <strong>of</strong> the mean squared distance<br />

between electrons.<br />

The Resta localization index, in a somewhat paradoxical<br />

manner, increases with the degree <strong>of</strong> delocalization. In view <strong>of</strong> this<br />

fact, it may seem legitimate to call this quantity rather<br />

"delocalization index". Alternatively one might choose e.g. the<br />

inverse <strong>of</strong> r 2 c<br />

, which would increase by increasing degree <strong>of</strong><br />

localization <strong>of</strong> the electron. However, such a change <strong>of</strong><br />

nomenclature would lead to confusion with earlier works, therefore<br />

the original terminology has been retained in this work.<br />

The discussion up to now has been about the in principle exact<br />

xc-hole, which supposes that both the ground state wave function in<br />

the expectation values <strong>and</strong> the linear response function appearing in<br />

the fluctuation-dissipation theorem are exact. Usually one has only<br />

some approximations available for these quantities, e.g. in the form<br />

<strong>of</strong> an independent particle method, like Hartree-Fock (HF) or<br />

Kohn-Sham (KS). In these cases, the single determinant wave<br />

function provides the exchange (Fermi) hole, which can be<br />

calculated from the first order density matrix. In principle, the same<br />

quantity can be obtained from the fluctuation-dissipation formula,<br />

using the the non-interacting (uncoupled Hartree-Fock, UCHF)<br />

response function. To reproduce expectation value results in this<br />

way, extremely large basis sets would be needed due to the slow<br />

convergence <strong>of</strong> the resolution <strong>of</strong> identity in Gaussian basis sets. A<br />

further possibility is to use the interacting, TDCHF (time-dependent<br />

coupled Hartree-Fock) polarizabilities in the fluctuation-dissipation<br />

formula, giving access to a low-order approximation <strong>of</strong> the<br />

correlation hole, but the problem <strong>of</strong> the basis set incompleteness<br />

remains still a crucial one. As far as direct correlation effects on the<br />

expectation value <strong>of</strong> the second cumulant moment are concerned, a<br />

few recent full CI calculations have been reported for linear Li n <strong>and</strong><br />

Be n chains in order to monitor the appearance <strong>of</strong> a metal-insulator<br />

type phase transitions as a function <strong>of</strong> the interatomic spacings<br />

[33, 34].<br />

The localization tensor itself can be related to the dynamic<br />

dipole polarizability tensor <strong>and</strong> it is approximately proportional to<br />

the static polarizability tensor <strong>of</strong> the system. The tensorial character<br />

<strong>of</strong> the localization function reflects the fact that the localizability<br />

can be different in different directions. For instance, one expects<br />

higher mobility <strong>and</strong> larger tensor elements in an aromatic plane <strong>of</strong> a<br />

molecule than in the perpendicular direction. A few examples <strong>of</strong><br />

theoretically calculated localization tensor are listed in Table 2,<br />

obtained at the HF/6-311G** level, using a development version <strong>of</strong><br />

the MOLPRO program package [35]. It is interesting that aromatic


3612 Current Organic Chemistry, 2011, Vol. 15, No. 20 János G. Ángyán<br />

Table 2. The trace <strong>of</strong> the Resta Localization Tensor r 2 theor<br />

c<br />

Calculation were done at the HF/6-311G** Level<br />

<strong>and</strong> the Corresponding Anisotropies, Defined as 2 = 1 2 [3Tr (r r 2 ) Tr(r<br />

c<br />

<br />

r <br />

c<br />

) 2 ].<br />

Molecule r 2 theor<br />

c<br />

Molecule r 2 theor<br />

c<br />

<br />

Ne 0.5856 0.0000 CHCH 1.8202 0.3610<br />

SF 6 0.8250 0.0000 CH 3-CH 3 1.8400 0.0056<br />

F 2 0.8470 0.1932 C 8H + 9 (homotropylium) 1.8424 0.1907<br />

Ar 0.9105 0.0000 CH 2=CH-CCH 1.8446 0.2991<br />

He 1.1765 0.0000 CH 2=CH 2 1.8510 0.1888<br />

H 2O 1.1791 0.0971 C 6H 6 1.8599 0.2171<br />

CO 1.2609 0.1489 CH 2-CH=CH-CH 2 1.8660 0.2613<br />

ClNO 1.2983 0.2473 C 10H 8 (naphtalene) 1.8711 0.2594<br />

N 2 1.3389 0.2720 CH 4 1.8960 0.0000<br />

H 2C=O 1.3840 0.1627 C 16H 10 (pyrene) 1.9028 0.2904<br />

CS 2 1.5537 0.4480 C 24H 12 (coronene) 1.9281 0.3224<br />

NH 3 1.5657 0.0602 CH 2=C=C=CH 2 1.9412 0.4889<br />

H 2C=NH 1.6315 0.1913 B 2H 6 2.1485 0.0822<br />

CH=C=NH 2 1.6761 0.2874 H 2 2.5961 0.3855<br />

systems, which are conventionally considered as strongly<br />

delocalized, have a mean localization tensor around 1.86-1.90 <strong>and</strong><br />

show a relatively moderate anisotropy. The methane molecule is<br />

characterized by a value <strong>of</strong> 1.896, slightly larger than that <strong>of</strong> the<br />

benzene <strong>and</strong> naphtalene. The presence <strong>of</strong> heteroatoms, i.e. a polar<br />

character <strong>of</strong> the electronic structure decreases dramatically the<br />

mean value <strong>of</strong> the localization tensor, indicating a stronger<br />

localization <strong>of</strong> electrons in these systems.<br />

4. <strong>Delocalization</strong> Indices <strong>and</strong> Distributed Polarizabilities<br />

While in the previous section we have seen that a global<br />

characterization <strong>of</strong> the localizability <strong>of</strong> electrons can be achieved<br />

via the Resta localization tensor, which is closely related to the total<br />

molecular dipole polarizability, in this section we continue our<br />

analysis in terms <strong>of</strong> a coarse grain, domain-based representation <strong>of</strong><br />

the full nonlocal charge density response function.<br />

Consider a partition <strong>of</strong> the full molecular space in disjoint, nonoverlapping<br />

domains, { a<br />

} . Although at this point the explicit<br />

nature <strong>of</strong> this partition is irrelevant, in order to fix the ideas, we can<br />

think about Bader's atoms in molecule (AIM) topological<br />

partitioning, as an example. Following Stone's concept <strong>of</strong><br />

distributed polarizability [36], the total charge density response can<br />

be partitioned to multipolar contributions assigned to pairs <strong>of</strong><br />

domains. The appropriate choice <strong>of</strong> the domains is quite important<br />

in order to have physically meaningful polarizability values. For<br />

instance, the Bader's AIM scheme seems to be a convenient choice,<br />

because <strong>of</strong> the excellent transferability properties <strong>of</strong> the atomic<br />

domains obtained in this framework [37-39]. The lowest-order<br />

multipolar contribution to the distributed polarizabilities is due to<br />

the variation <strong>of</strong> the domain net charges under the effect <strong>of</strong> an<br />

external potential or field: electric charges flow from one domain<br />

(atom) to another one, driven by the electrostatic field (potential<br />

difference) acting between them. The charge-flow polarizability is<br />

given by<br />

ab<br />

()= dr<br />

a<br />

dr(r, r ;), (13)<br />

b<br />

<strong>and</strong> it can be calculated at any level <strong>of</strong> response theory, provided<br />

one has transition matrix elements <strong>of</strong> the electron population<br />

operator over the various domains [37, 38]. Since the net electric<br />

charge <strong>of</strong> the molecule should be invariant under the effect <strong>of</strong> an<br />

external field, in addition to the symmetry property, ab<br />

= ba<br />

, the<br />

following charge-conservation sum rule holds:<br />

ab<br />

()=0. (14)<br />

b<br />

The charge conservation rule implies that the sum <strong>of</strong> the<br />

negative two-center charge-flow polarizabilities compensates<br />

exactly the on-site charge-charge polarizability, aa<br />

= <br />

ba ab<br />

.<br />

The fluctuation-dissipation theorem can be applied to the<br />

charge-flow polarizabilities ab<br />

() leading to the interdomain<br />

correlation <strong>of</strong> fluctuations (covariances) <strong>of</strong> the number <strong>of</strong> electrons,<br />

described by the domain (atomic) population operators, ˆNa<br />

<br />

<br />

d m ()= ˆN<br />

0 ab a<br />

ˆNb c<br />

. (15)<br />

ˆN a<br />

ˆNb c<br />

can be regarded as a domain decomposed form-factor,<br />

ˆN a<br />

ˆNb c<br />

= dr drS(r, r )= ab<br />

N a<br />

I ab<br />

. (16)<br />

a<br />

b<br />

The delocalization index, I ab<br />

, between different domains is the<br />

covariance <strong>of</strong> the populations (see e.g. Ref. [2])<br />

I ab<br />

= dr(r)<br />

a<br />

drh xc<br />

(r, r ), (17)<br />

<strong>and</strong> reflects the number <strong>of</strong> electrons shared by the two domains. In<br />

the a=b case I aa<br />

is related to the variance <strong>of</strong> the electron<br />

b


<strong>Linear</strong> <strong>Response</strong> <strong>and</strong> <strong>Measures</strong> <strong>of</strong> <strong>Electron</strong> <strong>Delocalization</strong> Current Organic Chemistry, 2011, Vol. 15, No. 20 3613<br />

F<br />

Cl<br />

O<br />

O<br />

1.3489<br />

1.7571 1.4697 1.2269<br />

1.1630 N<br />

1.1738<br />

N<br />

P<br />

N<br />

O<br />

133.25<br />

o<br />

O<br />

O<br />

129.77<br />

o<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

S<br />

O<br />

1.1569<br />

1.1601<br />

1.7375<br />

1.2857<br />

1.2916<br />

C<br />

1.7336<br />

C<br />

C<br />

C<br />

F<br />

o<br />

F<br />

Cl<br />

108.27 113.20<br />

o<br />

Cl<br />

S<br />

S<br />

O<br />

O<br />

F<br />

Cl<br />

O<br />

O<br />

1.3010 1.7453 1.4050 1.4134<br />

B<br />

B<br />

S<br />

B<br />

F<br />

F<br />

Cl<br />

Cl<br />

O<br />

O<br />

O<br />

O<br />

Fig. (1). Schematic illustration <strong>of</strong> the family <strong>of</strong> Y-conjugated systems. The numbers are the bond lengths. After Ref. [43].<br />

ΑAA<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.0 0.5 1.0 1.5 2.0 2.5<br />

N 2<br />

A c<br />

ΑAB<br />

0.0<br />

0.2<br />

0.4<br />

0.6<br />

0.8<br />

1.0 0.8 0.6 0.4 0.2 0.0<br />

N A N B c<br />

Fig. (2). Correlation <strong>of</strong> Bader atomic population fluctuations (a) <strong>and</strong> atom-atom fluctuations (b) with the corresponding charge-flow polarizabilities for Y-<br />

conjugated systems. On (a) green points (dashed line) are central atoms, red points (solid lines) are the oxygen lig<strong>and</strong>s, <strong>and</strong> other type <strong>of</strong> lig<strong>and</strong> atoms are<br />

marked by blue points (dotted line). On (b) only the bonds with the central atom are considered. The linear regression lines were calculated by excluding a few<br />

outlier points, indicated by different colors. After Ref. [43].<br />

population <strong>and</strong> it can be considered as a domain localization index,<br />

giving the number <strong>of</strong> electrons localized in a<br />

[2]. A similar<br />

analysis can be performed [40] from the viewpoint <strong>of</strong> the domain<br />

averaged Fermi hole (DAFH) concept [8, 41, 42].<br />

The analog <strong>of</strong> the charge-conservation sum rule for the<br />

delocalization indices is a simple consequence <strong>of</strong> the normalization<br />

<strong>of</strong> the exchange hole: ˆN 2 a<br />

c<br />

= <br />

ˆN ba a<br />

ˆNb c<br />

<strong>and</strong> we can<br />

anticipate the following proportionality:<br />

ˆN a 2 c<br />

aa<br />

<strong>and</strong> ˆN a<br />

ˆNb c<br />

ab<br />

(a b) (18)<br />

The proportionality factor is in principle system-dependent <strong>and</strong><br />

it behaves like a kind <strong>of</strong> mean excitation energy, characteristic to<br />

the atom pair. The expected proportionality between (static) charge<br />

flow polarizabilities <strong>and</strong> delocalization indices (<strong>and</strong> on-site chargecharge<br />

polarizabilities) has been tested for a family <strong>of</strong> related Y-<br />

conjugated compounds [43], shown on Fig. (1).<br />

It is clear from Fig. (2) that both the one- <strong>and</strong> two-center<br />

indices are in fact proportional to the corresponding charge-flow<br />

polarizabilities. In the one-center case the atoms have been<br />

classified in different categories: central B, C, N, S, P atoms (with<br />

the exclusion <strong>of</strong> the carbon atoms <strong>of</strong> CS 3 <strong>and</strong> COCl 2 ), oxygen<br />

lig<strong>and</strong> atom <strong>and</strong> halogen (F,Cl) <strong>and</strong> S lig<strong>and</strong> atoms. For the twocenter<br />

case only formally bonded atom pairs are shown, since the<br />

too weak bonding/polarizability does not lead to interpretable<br />

correlations. Disregarded <strong>of</strong> the above mentioned outliers, very<br />

good linear regressions could be found, indicating that for these<br />

special bonds there is a unique mean excitation energy. In some<br />

cases the delocalization indices as well as the single-center<br />

localization indices are quite low, below 0.5. One can consider such<br />

values as indicators for the presence <strong>of</strong> strongly ionic bonding.<br />

The benzene molecule is an example where the Fermi-hole<br />

extends farther than the nearest-neighbour atoms <strong>and</strong> the charge<br />

density response remains large even in the para (1,4) position. The<br />

-electron system <strong>of</strong> the benzene is usually considered as a


3614 Current Organic Chemistry, 2011, Vol. 15, No. 20 János G. Ángyán<br />

delocalized system, because the Fermi-hole <strong>of</strong> the -electrons is not<br />

confined to atoms <strong>and</strong> pairs <strong>of</strong> atoms, <strong>and</strong> it extends over the whole<br />

molecular skeleton.<br />

We can compare the charge flow polarizabilities <strong>and</strong> the<br />

delocalization indices in the case <strong>of</strong> the benzene molecule. The<br />

well-known relationships in organic chemistry between orthometa-<br />

<strong>and</strong> para-positions in benzene are quite well reflected by both<br />

the charge-flow polarizabilities <strong>and</strong> the delocalization indices.<br />

Distributed polarizability calculations [38] with the time dependent<br />

coupled Hartree-Fock (TDCHF) method provide the following<br />

charge-flow polarizabilities: -0.800 (ortho), +0.103 (meta) <strong>and</strong> -<br />

0.316 (para), while the delocalization indices [44] are 1.386 (ortho),<br />

0.072 (meta) <strong>and</strong> 0.098 (para). It is remarkable that the charge-flow<br />

polarizability for the meta atom is positive, which means that (at<br />

this level <strong>of</strong> approximation) the electron flow between atoms in<br />

meta position is in the direction <strong>of</strong> the sign <strong>of</strong> electric potential<br />

difference, although for negatively charge electrons it is expected to<br />

be the opposite. Applying the proportionality rule between the<br />

charge density response <strong>and</strong> the xc-hole one concludes that in the<br />

region <strong>of</strong> the meta carbon atoms the xc-hole function changes its<br />

sign. We remind that TDCHF includes some approximate<br />

Coulomb-correlation effects, while the UCHF polarizabilities which<br />

are related to the Fermi-hole function, should be negative<br />

everywhere in the space. Therefore it seems that the Coulombcorrelation<br />

effect enhances considerably the difference in the<br />

behavior <strong>of</strong> the meta <strong>and</strong> para carbon atoms. A comparison <strong>of</strong> these<br />

sites show considerably less difference in terms <strong>of</strong> delocalization<br />

indices calculated from the single determinant.<br />

The question arises, whether it can be defined in general terms<br />

the criteria <strong>of</strong> having well-localized domains in a molecule In the<br />

light <strong>of</strong> the above discussion, such a criterion can be defined by the<br />

equality, I aa<br />

N a<br />

, which implies that the xc-hole must be, at least<br />

approximately, normalized over the domain a<br />

, i.e.<br />

drh xc<br />

(r, r ) 1 . Under such circumstances the local sum rule<br />

a<br />

dr drS(r, r ) 0 (19)<br />

a<br />

a<br />

holds true, providing a condition <strong>of</strong> the localizability over the<br />

domain a<br />

.<br />

Depending on the chemical nature <strong>of</strong> the atom in question,<br />

Bader's AIM domains can be either well-localized or delocalized<br />

character. For instance, domains corresponding to a mono- or<br />

polyatomic ion in a strongly ionic structure will be naturally<br />

localized, just like subsets <strong>of</strong> atoms that constitute a closed shell<br />

molecule in a weakly bound intermolecular complex. On the other<br />

h<strong>and</strong>, atoms connected by a covalent bond are not well-localized<br />

because their domains comprise only a part <strong>of</strong> the xc-hole,<br />

extending over two or more atomic basins. Genuinely welllocalized<br />

domains must be constructed according to appropriate<br />

criteria. The main purpose <strong>of</strong> numerous local functions<br />

characterizing localization is precisely to identify regions <strong>of</strong> highly<br />

localizable <strong>and</strong> less localizable character. For instance, ELF<br />

(electron localization function) attractor domains [45] are probably<br />

good c<strong>and</strong>idates for naturally well-localized regions <strong>of</strong> space <strong>and</strong> it<br />

can be expected that the inter-region fluctuations are relatively low.<br />

5. MOLECULAR ORBITAL LOCALIZATION CRITERIA<br />

In the following it will be shown that some <strong>of</strong> the most popular<br />

localization criteria can be, in fact, traced back to the requirement<br />

<strong>of</strong> minimum deformability (polarizability) <strong>of</strong> the one-electron<br />

orbitals building up the Slater-determinant <strong>of</strong> an independent<br />

particle wave function.<br />

In his 1960 paper, entitled “Construction <strong>of</strong> some molecular<br />

orbitals to be approximately invariant for changes from one<br />

molecule to another” Boys proposed to “impose a mathematical<br />

constraint <strong>of</strong> maximum insensitivity to alterations in the distant<br />

nuclear charges”. Such a criterion was supposed “to lead orbitals<br />

which are localized around the chemical valency links <strong>and</strong> the<br />

atomic lone pairs” [46]. This condition means that the shape <strong>of</strong><br />

these one-electron functions (orbitals) ensures their maximal<br />

stiffness <strong>and</strong> minimal polarizability <strong>and</strong>, by consequence their<br />

insensitivity to the chemical environment. In other, more familiar<br />

terms, such orbitals are supposed to be maximally transferable.<br />

The final form <strong>of</strong> the Boys-Foster localization criterion [47],<br />

established only a few years after the above-cited work, requires<br />

that , the spread<br />

= i<br />

| r 2 | i<br />

i<br />

| r | i<br />

2 (20)<br />

i<br />

be minimal after a series <strong>of</strong> unitary transformations <strong>of</strong> the orbitals.<br />

Of course, such transformations, preserving the orthonormality <strong>of</strong><br />

the orbitals do not not change the occupied manifold <strong>and</strong> leave the<br />

single-determinant wave function invariant. The condition <strong>of</strong><br />

minimizing the spread has been introduced in solid state physics<br />

literature many years after the formulation <strong>of</strong> Boys localization<br />

critera [11] leading to maximally localized Wannier functions in<br />

solids.<br />

The spread is the sum <strong>of</strong> the dipole moment fluctuations<br />

associated with the individual orbitals <strong>and</strong> therefore it can be<br />

considered as a measure <strong>of</strong> the stiffness <strong>of</strong> the orbitals. At the<br />

minimum <strong>of</strong> this functional the sum <strong>of</strong> the one-electron<br />

polarizabilities associated to the orbitals is supposed to be the<br />

smallest possible. Resta has shown that the trace <strong>of</strong> the localization<br />

function is a lower bound <strong>of</strong> the spread [14]. In fact, in case <strong>of</strong> a<br />

single determinant wave function, the trace <strong>of</strong> the localization<br />

tensor can be written as<br />

N = | i<br />

| r | j<br />

| 2 . (21)<br />

i<br />

Note that has been defined as the second cumulant moment<br />

per electron, therefore one has to multiply the lhs by N . Since is<br />

positive <strong>and</strong> invariant with respect to the unitary transformation <strong>of</strong><br />

the orbitals, the minimum <strong>of</strong> will correspond to the minimum <strong>of</strong><br />

squared sum <strong>of</strong> the <strong>of</strong>f-diagonal dipole moment matrix elements, or<br />

to the maximum <strong>of</strong> the squared sum <strong>of</strong> the diagonal dipole moment<br />

matrix elements. The scheme <strong>of</strong> the Boys localization can be thus<br />

summarized as,<br />

ji<br />

i<br />

min{ i<br />

| r 2 | i<br />

i<br />

| r | i<br />

2 }. (22)<br />

i<br />

<br />

invariant<br />

i<br />

<br />

maximal<br />

Although these auxiliary criteria are more <strong>of</strong>ten practical in<br />

computational implementations, it is important to keep in mind that<br />

the underlying physical criterion is the maximum transferability or<br />

minimum polarizability <strong>of</strong> the orbitals, expressed by the condition<br />

<strong>of</strong> minimal dipole fluctuation.<br />

One can proceed in an analogous way if the localizability is<br />

measured by the delocalization <strong>and</strong> localization indices over<br />

domains. Supposing a given partition <strong>of</strong> the molecular space, the


<strong>Linear</strong> <strong>Response</strong> <strong>and</strong> <strong>Measures</strong> <strong>of</strong> <strong>Electron</strong> <strong>Delocalization</strong> Current Organic Chemistry, 2011, Vol. 15, No. 20 3615<br />

sum <strong>of</strong> localization indices (cf. Eqs (16)-(18)) for all the regions<br />

can be written as<br />

i<br />

| ˆNa | i<br />

| i<br />

| ˆNa | j<br />

| 2 . (23)<br />

a<br />

i<br />

a<br />

While this quantity is invariant to unitary transformations <strong>of</strong> the<br />

orbitals, it is not so for the sum <strong>of</strong> single-orbital localization<br />

indices:<br />

D = i<br />

| ˆNa | i<br />

| i<br />

| ˆNa | i<br />

| 2 , (24)<br />

a<br />

i<br />

which should be minimized in order to get optimally localized<br />

orbitals. Following a similar reasoning as previously, the<br />

minimization <strong>of</strong> the fluctuation can be replaced by maximization <strong>of</strong><br />

the sum <strong>of</strong> square <strong>of</strong> the orbital populations. Summarizing these<br />

localization schemes based on the minimization <strong>of</strong> atomic<br />

populations over the orbitals:<br />

min{<br />

<br />

i<br />

| ˆNa | i<br />

<br />

i<br />

a <br />

invariant<br />

<br />

ij<br />

a<br />

i<br />

i<br />

| ˆNa | i<br />

2 }. (25)<br />

i<br />

a <br />

maximal<br />

This is the Pipek-Mezey localization criterion, which has been<br />

generalized to multideterminant wave functions <strong>and</strong> Bader atomic<br />

domain populations (instead <strong>of</strong> the originally used Mulliken<br />

populations) by Cioslowski [48].<br />

Popular in quantum chemistry before the appearance <strong>of</strong> more<br />

fundamental approaches in the years '90 <strong>and</strong> widely used to<br />

characterize bonding in molecules, localized orbitals are <strong>of</strong>ten<br />

considered as suspect <strong>and</strong> theoretically unfounded tools for the<br />

analysis <strong>of</strong> chemical bonding. Paradoxically, in solid state<br />

electronic structure studies, the equivalent maximally localized<br />

Wannier functions enjoy a certain popularity <strong>and</strong> play a crucial role<br />

in the interpretation <strong>of</strong> electric polarization phenomena in solids.<br />

The so-called "Wannier centers" [11], i.e. the centroids <strong>of</strong> localized<br />

orbitals, are routinely used to interpret the evolution <strong>of</strong> the<br />

electronic structure during ab initio molecular dynamics<br />

trajectories. Is there a deeper reason for the success <strong>of</strong> localized<br />

orbitals<br />

The answer is affirmative <strong>and</strong> it is based on a strong connection<br />

between localized orbitals <strong>and</strong> the Fermi-hole. This relationship can<br />

be understood by taking a simple model for a well-localized system<br />

described by a single-determinant wave function, expressed in<br />

terms <strong>of</strong> a set <strong>of</strong> localized spin-orbitals, { i<br />

(r)} [49]. If this<br />

system is well-localized, the orbitals are strictly localized, i.e. the<br />

product (differential overlap) i<br />

(r) j<br />

(r) <strong>of</strong> two different<br />

localized orbitals ( i j ) <strong>of</strong> -spin electrons is (vanishingly)<br />

small. In such a case the Fermi-hole, can be approximately<br />

represented in a “diagonal” form:<br />

h x<br />

(r, r ) 1<br />

occ<br />

| i<br />

(r)| 2 | i<br />

( r )| 2 . (26)<br />

<br />

(r) i<br />

The regions occupied by the strictly localized orbitals partition<br />

the molecular space is to approximately disjoint regions, i<br />

,<br />

characterized by the condition that | i<br />

(r)| 2 / <br />

(r) 1 . With the<br />

help <strong>of</strong> the function i<br />

(r) , defined as unity if r i<br />

<strong>and</strong> 0<br />

otherwise, the Fermi-hole function is approximately<br />

occ<br />

h x<br />

(r,r') i<br />

(r)| i<br />

(r') | 2 . (27)<br />

i<br />

The above equation expresses the rule observed by Luken [50-<br />

52], who remarked that the shape <strong>of</strong> the Fermi-hole remains stable<br />

within a domain <strong>of</strong> a given electron pair. Moreover, in these cases<br />

the shape <strong>of</strong> the Fermi-hole strongly resembles to the square <strong>of</strong> the<br />

localized orbital. In this way one can establish a qualitative or semiquantitative<br />

relationship between the localization from a "physical"<br />

linear response viewpoint, <strong>and</strong> the localized orbital picture,<br />

sometimes considered in the modern quantum chemical literature as<br />

an obsolete one. The relationship between localized orbitals <strong>and</strong><br />

domain averaged Fermi hole functions has been raised by Matito<br />

<strong>and</strong> Salvador on the one h<strong>and</strong> <strong>and</strong> Mayer on the other in the<br />

Faraday Discussion on Chemical Concepts from Quantum<br />

Mechanics [53].<br />

6. LOCALIZABILITY CHARACTERIZED BY SINGLE-<br />

VARIABLE FUNCTIONS OVER THE MOLECULAR SPACE<br />

In addition to a global characterization <strong>of</strong> localization, like<br />

Resta localization tensor or index, or coarse grain localization<br />

measures like the intra- <strong>and</strong> intra-domain (e.g. atomic <strong>and</strong> diatomic)<br />

localization <strong>and</strong> delocalization indices (sometimes referred to as<br />

covalent bond orders <strong>and</strong> valences), there is a definite need for<br />

pointwise measures <strong>of</strong> electron localization as well. The ultimate<br />

expectation towards such functions is that they provide a pictorial<br />

representation about the organization <strong>of</strong> electron pairs in the space<br />

<strong>and</strong> visualize the Lewis dot structure <strong>of</strong> molecules.<br />

One <strong>of</strong> the widely used function to characterize electron<br />

localization is the negative <strong>of</strong> the Laplacian <strong>of</strong> the electron density,<br />

L(r)= 2 (r) . The Laplacian reflects the shell structure <strong>of</strong> atoms<br />

<strong>and</strong>, as it has been first pointed out on empirical grounds by Bader<br />

<strong>and</strong> coworkers [54], there is a good correspondence between the<br />

topology <strong>of</strong> the negative Laplacian <strong>and</strong> the domains where the<br />

electron pairings are the strongest, providing thus a link between<br />

L(r) <strong>and</strong> Lewis or VSEPR models <strong>of</strong> chemical bonding [7]. The<br />

regions <strong>of</strong> maxima <strong>of</strong> L(r) correspond to local charge<br />

concentrations, while negative regions <strong>of</strong> L(r) can be associated<br />

with local charge depletion. This behavior <strong>of</strong> the 3-dimensional<br />

Laplacian function could be put into correspondence with the<br />

properties <strong>of</strong> the 6-dimensional Fermi-hole distribution. In<br />

particular there is a homeomorphism between the Laplacian <strong>of</strong> the<br />

charge density <strong>and</strong> the Laplacian <strong>of</strong> the conditional pair distribution<br />

function.<br />

As we were able to relate the xc-hole to the change density<br />

response function by a sum rule, the fluctuation-dissipation<br />

theorem, analogously, one can relate the Laplacian <strong>of</strong> the charge<br />

density to another sum rule, the generalized f-sum rule, which is<br />

reads as [55]<br />

+<br />

c(r, r )= 1 <br />

dm(r, r ;). (28)<br />

<br />

The real function c(r, r ) can be expressed as a ground state<br />

expectation value <strong>of</strong> the double commutator <strong>of</strong> the Hamiltonian<br />

with the charge density operator at two different positions,<br />

c(r 1<br />

,r )=0|[[Ĥ, ˆ(r 2 1)], ˆ(r 2<br />

)] | 0 = 1<br />

1 ( (r 1<br />

r 2<br />

)(r 1<br />

)). (29)<br />

After having taken the derivatives on the rhs, one <strong>of</strong> the terms is<br />

equal to the negative <strong>of</strong> Laplacian multiplied by the Dirac delta<br />

function, (r 1<br />

r 2<br />

) 2 1<br />

(r 1<br />

) . The usual dipole oscillator strength<br />

Thomas-Reiche-Kuhn sum rule can be obtained by double<br />

integrating the second moment <strong>of</strong> c(r 1<br />

,r 2<br />

) . In particular, the integral


3616 Current Organic Chemistry, 2011, Vol. 15, No. 20 János G. Ángyán<br />

<strong>of</strong> the second moment <strong>of</strong><br />

L(r)= 2 (r) turns out to be<br />

proportional to the number <strong>of</strong> electrons.<br />

Another family <strong>of</strong> localization functions is based on various<br />

forms <strong>of</strong> the local covariance <strong>of</strong> the electron pair distribution [6]. A<br />

first example <strong>of</strong> this category <strong>of</strong> pointwise functions has been the<br />

ELF proposed by Becke <strong>and</strong> Edgecombe [56] <strong>and</strong> later interpreted<br />

<strong>and</strong> reformulated by others, see e.g. [45, 57, 58]. The ELF is based<br />

on a second order expansion <strong>of</strong> the Fermi-hole density <strong>and</strong> use an<br />

appropriate transformation function, which has the role to<br />

renormalize its possible values between 0 <strong>and</strong> 1 <strong>and</strong> accentuate the<br />

graphical expression <strong>of</strong> the shell structure. Among the numerous<br />

possible formulations <strong>of</strong> electron localization functions one can cite<br />

e.g. the localized orbital locator (LOL) <strong>of</strong> Schmider <strong>and</strong> Becke [59],<br />

the electron localizability indicator (ELI) <strong>of</strong> Kohout [60]. An<br />

interesting generalization <strong>of</strong> this type <strong>of</strong> approach consists in<br />

extending the radius <strong>of</strong> spherical interation from an infinitesimal<br />

domain to a larger finite one, leading to the radial exchange density<br />

function <strong>of</strong> Geier [61].<br />

Still another possibility to construct local functions can be<br />

based on integrated or global properties <strong>of</strong> the xc-hole distribution,<br />

h xc<br />

(r, r ) belonging to the reference point, r . One <strong>of</strong> the<br />

characteristic features <strong>of</strong> the hole function normalized to -1 is the<br />

position <strong>of</strong> its centroid or its center <strong>of</strong> charge:<br />

'<br />

D <br />

(r)= drr <br />

h xc<br />

(r, r ), (30)<br />

which is a three-dimensional vector ( = x, y,z ) in the laboratory<br />

reference frame. The xc-hole distribution belonging to the reference<br />

point r , is centered in D(r) , which may serve as a natural<br />

expansion center for non-local hole models. Furthermore, according<br />

to the charge density reconstitution sum rule [62], the total<br />

electronic dipole moment μ <br />

can be obtained by the density<br />

weighted average <strong>of</strong> the xc-hole centroids:<br />

μ <br />

= dr(r)D <br />

(r). (31)<br />

This equation appears as the fuzzy generalization <strong>of</strong> the formula<br />

to obtain the polarization vector in terms <strong>of</strong> the Wannier centers, as<br />

it is used in solid state theory. Notice that the x-hole centroid<br />

function, D(r) has appeared as an auxiliary quantity in the Becke-<br />

Johnson model for dispersion forces [63], where these authors<br />

introduced the notion <strong>of</strong> "exchange-hole dipole moment" for the<br />

case <strong>of</strong> single-determinant wave functions. Associating a simple,<br />

classical picture to the negative unit point charge at r , representing<br />

the electron, <strong>and</strong> a positive unit charge in D(r) , representing the<br />

hole, the exchange-hole dipole moment function becomes:<br />

d(r) = D(r) - r, (32)<br />

which will be useful in the forthcoming discussions.<br />

As a tentative local scalar characterization <strong>of</strong> localization, one<br />

can start from the explicitly origin-independent form <strong>of</strong> the second<br />

cumulant moment expression (11) <strong>and</strong> define the following<br />

function,<br />

(r)= 1 2 (r) dr (r r )(r<br />

r )h<br />

(r, r )<br />

(33)<br />

xc<br />

measuring the trace <strong>of</strong> the second moment <strong>of</strong> the xc-hole, i.e. its<br />

spread in the space, weighted by the electron density at the<br />

reference point. This quantity has the advantage, that by its<br />

integration over the whole space one obtains the global Resta<br />

localization index, .<br />

Taking into account the fact that the hole-centroid, D(r)<br />

appears as a natural expansion centre for multipolar analysis <strong>of</strong> the<br />

hole, an "intrinsic" second moment distribution can be defined as<br />

(r)= 1 2 (r) dr (D(r) r )(D(r) r )h xc<br />

(r, r ). (34)<br />

Unfortunately this quantity does not integrate to the second<br />

cumulant moment <strong>of</strong> the total electronic system.<br />

Finally we can consider the difference <strong>of</strong> (r) <strong>and</strong> (r) <strong>and</strong><br />

construct the following function [64],<br />

(r)=2( (r)(r) )= (r)d(r) d(r), (35)<br />

which integrates to the density weighted squared exchange-hole<br />

dipole moment<br />

d 2 = dr(r)= dr(r)d 2 (r). (36)<br />

The quantity d 2 can be regarded as an approximation to the<br />

correlated second cumulant moment <strong>of</strong> the electronic system,<br />

calculated, somewhat paradoxically, from a single determinant<br />

wave function, i.e. from the exchange hole [15, 63, 65].<br />

It has been argued [64], that since (r) measures the squared<br />

distance <strong>of</strong> the electron to its exchange-hole centroid, it has a small<br />

value for the regions where the electron is close to the center <strong>of</strong> its<br />

own exchange-hole, therefore it indicates a high probability <strong>of</strong><br />

having strongly paired electrons. An electron localization function,<br />

which has been called Fermi-hole locality indicator (FHLI), can be<br />

constructed [64] by the help <strong>of</strong> a transformation function, as<br />

FHLI(r)= ( 1+(r) ) 1 . (37)<br />

The FHLI function is quite different from the ELF-like<br />

functions, since instead <strong>of</strong> using strictly local features <strong>of</strong> the<br />

exchange hole, it is constructed on the basis <strong>of</strong> a strongly non-local<br />

property, the distance <strong>of</strong> the electron to the centroid <strong>of</strong> its hole.<br />

(a) acetylene (b) ethylene (c) ethane<br />

Fig. (3). Ethylene, ethane <strong>and</strong> acetylene FHLI functions. (After Ref. [64]).<br />

A few illustrative examples <strong>of</strong> the FHLI function have been<br />

calculated by a development version <strong>of</strong> the MOLPRO program<br />

package [35]. The 6-311G** basis set has been used <strong>and</strong> the FHLI<br />

functions were represented on a regular grid in the Gaussian cube<br />

format <strong>and</strong> visualized by the VMD molecular graphics package<br />

[66].<br />

The function FHLI provides a quite reasonable picture <strong>of</strong> the<br />

electron localization, which resembles, at least in some respects, to<br />

conventional ELF images. In the series <strong>of</strong> the prototypes for<br />

carbon-carbon triple (Fig. 3) double <strong>and</strong> single bonds one can<br />

clearly distinguish the signature <strong>of</strong> different bonding types. At the<br />

chosen isodensity contour the triple <strong>and</strong> double bonds are more<br />

voluminous than the carbon-carbon -bond. However, the<br />

representation <strong>of</strong> the bonds between non-hydrogen atoms seems to<br />

differ markedly from the ELF pictures. While the X-H bonds


<strong>Linear</strong> <strong>Response</strong> <strong>and</strong> <strong>Measures</strong> <strong>of</strong> <strong>Electron</strong> <strong>Delocalization</strong> Current Organic Chemistry, 2011, Vol. 15, No. 20 3617<br />

(a) butatriene (CH 2=C=C=CH 2 ) (b) vinylacetylene (HC C CH=CH 2)<br />

Fig. (4). FHLI functions <strong>of</strong> butatriene <strong>and</strong> vinylacetylene.<br />

appear as a single domain, the electrons that contribute to a given<br />

X-Y bond leave their signatures in the form <strong>of</strong> a more-or-less<br />

spacious FHLI domains, located near the atomic centers which<br />

participate in the bonding. The -bond <strong>of</strong> ethane can be recognized<br />

from the two small domains localized on the C-C axis near the<br />

atoms. In the case <strong>of</strong> ethylene the four electrons <strong>of</strong> the double bond<br />

appear as symmetrically arranged lobes above <strong>and</strong> below the<br />

molecular plane, corresponding to a “banana-type” representation.<br />

The triple bond in acetylene can be recognized from the cylindric<br />

shapes located near the carbon atoms.<br />

In the case <strong>of</strong> isomeric C 4 H 4 structures <strong>of</strong> butatriene <strong>and</strong><br />

vinylacetylene, one can distinguish the triple <strong>and</strong> double bonds in<br />

the latter, while in the butatriane one clearly sees the allen-like<br />

arrangement <strong>of</strong> three double bonds. It is interesting to observe in<br />

butatriene that the two central atoms contribute by significantly less<br />

spacious in-plane lobes than the terminal atoms. In the case <strong>of</strong><br />

vinylacetylene, the deformation <strong>of</strong> the disks representing the triple<br />

bonds is obviously due to the perturbative effect <strong>of</strong> the molecular<br />

skeleton (Fig. 4).<br />

In the case <strong>of</strong> aminoacetylene (Fig. 5a) the most significant<br />

structure is the double disk <strong>of</strong> the triple bond. Around the NH 2<br />

group, the H atoms <strong>and</strong> the lone pair form an almost regular 3-fold<br />

symmetric arrangement. The methylenimine (Fig. 5b), which is isoelectronic<br />

with ethylene one can observe a kind <strong>of</strong> fusion <strong>of</strong> the<br />

nitrogen lone pair <strong>and</strong> the p z -like lobes <strong>of</strong> the double bond located<br />

on the same nitrogen.<br />

localized electrons are less, delocalized electron are more<br />

polarizable - can be put in more precise terms <strong>and</strong> establish a clear<br />

relationship with the more conventional descriptors <strong>of</strong> electron<br />

localization, based on the xc-hole or more specifically on the<br />

Fermi-hole <strong>and</strong> its properties. The view advocated in this account is<br />

inspired by recent developments in solid state electronic structure<br />

theory, based on a parallel between the modern theory <strong>of</strong><br />

polarization <strong>and</strong> the notion <strong>of</strong> electron localization. It has been<br />

shown that the Resta localization index, used until now almost<br />

exclusively as a global measure to characterize the itinerant or nonitinerant<br />

character <strong>of</strong> electrons in solids, can be <strong>of</strong> some utility for<br />

molecular systems as well. Moreover, using experimental dipole<br />

oscillator strength distributions, the localization index seems to be<br />

accessible on experimental grounds as well.<br />

The fluctuation or covariance <strong>of</strong> the domain populations in a<br />

molecule, widely used as coarse-grain measures to describe<br />

localization <strong>and</strong> delocalization between atoms (or other domains)<br />

has been related to the charge-flow polarizability, which can be<br />

derived directly from the charge density response function. The<br />

approximate proportionality between charge-flow polarizabilities<br />

<strong>and</strong> delocalization indices has been illustrated on a few selected<br />

examples.<br />

Orbital localization criteria can also be reformulated as a<br />

requirement <strong>of</strong> constructing the most transferable <strong>and</strong> therefore the<br />

less polarizable set <strong>of</strong> one-electron functions. The Boys localization<br />

criterion is based on the dipole fluctuation, while the Pipek-Mezey<br />

<strong>and</strong> related criteria can be derived from the requirement <strong>of</strong> minimal<br />

fluctuation <strong>of</strong> atomic (or domain) populations.<br />

Finally, it has been shown that the Fermi hole locality indicator,<br />

which is able to characterize the electron localization by measuring<br />

some non-local features <strong>of</strong> the Fermi-hole, integrates to an<br />

approximate value <strong>of</strong> the second cumulant moment, thereby related<br />

again to the linear response function.<br />

Hopefully, the above cited examples could illustrate the<br />

intimate link between the basic physical mechanisms that govern<br />

the response properties <strong>of</strong> an electronic system <strong>and</strong> the specific<br />

organization <strong>of</strong> fermions, which leads to the important chemical<br />

concept <strong>of</strong> electron localization.<br />

8. ACKNOWLEDGEMENTS<br />

The author is grateful to Dr. A. Savin (Paris) for reading the<br />

manuscript <strong>and</strong> for his comments. The financial support <strong>of</strong> the<br />

Agence Nationale de la Recherche (ANR) in the framework <strong>of</strong> the<br />

project WADEMECOM (ANR-07-BLAN-0272) is gratefully<br />

acknowledged.<br />

REFERENCES<br />

(a) aminoacetylene (HC C NH )<br />

2<br />

(b) methylenimine (CH2 =NH)<br />

Fig. (5). FHLI functions <strong>of</strong> methylenimine <strong>and</strong> aminoacetylene.<br />

Further examples have been presented in Ref. [64], which<br />

demonstrate that the FHLI is able to provide an useful pictorial<br />

representation <strong>of</strong> the electron pair structure <strong>of</strong> molecules.<br />

7. CONCLUSIONS<br />

The electron localization/delocalization problem has been<br />

presented from the viewpoint <strong>of</strong> the linear response properties <strong>of</strong><br />

the molecular charge density to an external electric field<br />

perturbation. It has been shown that a seemingly loose statement -<br />

[1] Merino, G.; Vela, A.; Heine, T. Description <strong>of</strong> electron delocalization via the<br />

analysis <strong>of</strong> molecular fields. Chem. Rev. 2005. 105, 3812-3841.<br />

[2] Poater, J.; Duran, M.; Solà, M.; Silvi, B. Theoretical evaluation <strong>of</strong> electron<br />

delocalization in aromatic molecules by means <strong>of</strong> Atoms in Molecules (AIM)<br />

<strong>and</strong> <strong>Electron</strong> Localization Function (ELF) topological approaches. Chem.<br />

Rev. 2005. 105, 3911-3947.<br />

[3] Gatti, C. Chemical bonding in crystals: new directions. Z. Kristallogr. 2005.<br />

220, 399-457.<br />

[4] Cortés-Guzmán, F.; Bader, R.F.W. Complementarity <strong>of</strong> QTAIM <strong>and</strong> MO<br />

theory in the study <strong>of</strong> bonding in donor-acceptor complexes. Coord. Chem.<br />

Rev. 2005. 249, 633-662.<br />

[5] Savin, A. The electron localization function (ELF) <strong>and</strong> its relatives:<br />

interpretations <strong>and</strong> difficulties. J. Mol. Struct. (Theochem) 2005. 727, 127-<br />

131.<br />

[6] Ayers, P.W. <strong>Electron</strong> localization functions <strong>and</strong> local measures <strong>of</strong> the<br />

covariance. J. Chem. Sci. 2005. 117, 441-454.<br />

[7] Bader, R.F.W.; Hernández-Trujillo, J.; Cortés-Guzmán, F. Chemical<br />

bonding: From Lewis to atoms in molecules. J. Comp. Chem. 2007. 28, 4-14.


3618 Current Organic Chemistry, 2011, Vol. 15, No. 20 János G. Ángyán<br />

[8] Ponec, R.; Cooper, D.L. Anatomy <strong>of</strong> bond formation. Bond length<br />

dependence <strong>of</strong> the extent <strong>of</strong> electron sharing in chemical bonds from the<br />

analysis <strong>of</strong> domain-averaged Fermi holes. Faraday Discuss. 2007. 135, 31-<br />

42.<br />

[9] Mayer, I. Bond order <strong>and</strong> valence indices: a personal account. J. Comp.<br />

Chem. 2007. 28, 204-221.<br />

[10] Scemama, A.; Caffarel, M.; Savin, A. Maximum probability domains from<br />

Quantum Monte Carlo calculations. J. Comp. Chem. 2007. 28, 442-454.<br />

[11] Marzari, N.; V<strong>and</strong>erbilt, D. Maximally localized generalized Wannier<br />

functions for composite energy b<strong>and</strong>s. Phys. Rev. B. 1997. 56, 12847-12865.<br />

[12] Resta, R.; Sorella, S. <strong>Electron</strong> localization in the insulating state. Phys. Rev.<br />

Lett. 1999. 82, 370-373.<br />

[13] Resta, R. Why are insulators insulating <strong>and</strong> metals conducting J. Phys.:<br />

Condens. Matter 2002. 14, R625-R656.<br />

[14] Resta, R. Kohn's theory <strong>of</strong> the insulating state: A quantum-chemistry<br />

viewpoint. J. Chem. Phys. 2006. 124, 104104.<br />

[15] Ángyán, J.G. On the exchange-hole model <strong>of</strong> London dispersion forces. J.<br />

Chem. Phys. 2007. 127, 024108.<br />

[16] Matta, C.F.; Hernández-Trujillo, J.; Bader, R.F.W. Proton spin-spin coupling<br />

<strong>and</strong> electron delocalization. J. Phys. Chem. A 2002. 106, 7369-7375.<br />

[17] Soncini, A.; Lazzeretti, P. Nuclear spin-spin coupling density functions <strong>and</strong><br />

the Fermi-hole. J. Chem. Phys. 2003. 119, 1343-1349.<br />

[18] Soncini, A.; Lazzeretti, P. Nuclear spin-spin coupling density in molecules.<br />

J. Chem. Phys. 2003. 118, 7165-7173.<br />

[19] Heine, T.; Corminboeuf, C.; Seifert, G. The magnetic shielding function <strong>of</strong><br />

molecules <strong>and</strong> -electron delocalization. Chem. Rev. 2005. 105, 3889-3910.<br />

[20] Hirschfelder, J.O.; Brown, W.B.; Epstein, S.T. Recent developments in<br />

perturbation theory. Adv.Quantum Chem. 1964. 1, 255-374.<br />

[21] Olney, T.N.; Cann, N.M.; Cooper, G.; Brion, C.E. Absolute scale<br />

determination for photoabsorption spectra <strong>and</strong> the calculation <strong>of</strong> molecular<br />

properties using dipole sum-rules. Chem. Phys. 1997. 223, 59-98.<br />

[22] Kumar, A.; Meath, W.J. Pseudo-spectral dipole oscillator strengths <strong>and</strong><br />

dipole-dipole <strong>and</strong> triple-dipole dispersion energy coefficients for HF, HCl,<br />

HBr, He, Ne, Ar, Kr <strong>and</strong> Xe. Mol. Phys. 1985. 54, 823-833.<br />

[23] Kumar, A.; Fairley, G.R.G.; Meath, W.J. Dipole properties, dispersion<br />

energy coefficients, <strong>and</strong> integrated oscillator strengths for SF 6. J. Chem.<br />

Phys. 1985. 83, 70-77.<br />

[24] Kumar, A.; Kumar, M.; Meath, W.J. Dipole oscillator strengths, dipole<br />

properties <strong>and</strong> dispersion energies for SiF 4. Mol. Phys. 2003. 101, 1535-<br />

1543.<br />

[25] Pazur, R.J.; Kumar, A.; Thuraisingham, R.A.; Meath, W.J. Dipole oscillator<br />

strength properties <strong>and</strong> dispersion energy coefficients for H 2S. Can. J. Chem.<br />

1988. 66, 615-619.<br />

[26] Kumar, A.; Meath, W.J. Dipole oscillator strength distributions, properties<br />

<strong>and</strong> dispersion energies for the dimethyl, diethyl <strong>and</strong> methyl-propyl ethers.<br />

Mol. Phys. 2008. 106, 1531-1544.<br />

[27] Kumar, M.; Kumar, A.; Meath, W.J. Dipole oscillator strength properties <strong>and</strong><br />

dispersion energies for Cl 2. Mol. Phys. 2002. 100, 3271-3279.<br />

[28] Zeiss, G.D.; Meath, W.J.; MacDonald, J.C.F.; Dawson, D.J. Dipole oscillator<br />

strength distributions, sums, <strong>and</strong> some related properties for Li, N, O, H 2, N 2,<br />

O 2, NH 3, H 2O, NO, <strong>and</strong> N 2O. Can. J. Phys. 1977. 55, 2080-2100.<br />

[29] Kumar, A.; Meath, W.J. Reliable isotropic <strong>and</strong> anisotropic dipole properties,<br />

<strong>and</strong> dipolar dispersion energy coefficients, for CO evaluated using<br />

constrained dipole oscillator strength techniques. Chem. Phys. 1994. 189,<br />

467-477.<br />

[30] Kumar, A.; Meath, W.J. Dipole oscillator strength properties <strong>and</strong> dispersion<br />

energies for acetylene <strong>and</strong> benzene. Mol. Phys. 1992. 75, 311-324.<br />

[31] Jhanwar, B.; Meath, W.J.; MacDonald, J. Dipole oscillator strength<br />

distributions <strong>and</strong> related properties for ethylene, propene <strong>and</strong> 1-butene. Can.<br />

J. Phys. 1983. 61, 1027-1034.<br />

[32] Koga, T.; Matsuyama, H. Physical significance <strong>of</strong> second electron-pair<br />

moments in position <strong>and</strong> momentum spaces. J. Chem. Phys. 2001. 115,<br />

3984-3991.<br />

[33] Bendazzoli, G.L.; Evangelisti, S.; Monari, A.; Paulus, B.; Vetere, V. Full<br />

configuration-interaction study <strong>of</strong> the metal-insulator transition in model<br />

systems. J. Phys.: Conf. Ser. 2008. 117, 012005.<br />

[34] Vetere, V.; Monari, A.; Bendazzoli, G.L.; Evangelisti, S.; Paulus, B. Full<br />

configuration interaction study <strong>of</strong> the metal-insulator transition in model<br />

systems: Li n linear chains (n = 2,4,6,8). J. Chem. Phys. 2008. 128, 024701.<br />

[35] Werner, H.J.; Knowles, P.J.; Lindh, R.; Manby, F.R.; Schütz, M.; Celani, P.;<br />

Korona, T.; Mitrushenkov, A.; Rauhut, G.; Adler, T.B.; Amos, R.D.;<br />

Bernhardsson, A.; Berning, A.; Cooper, D.L.; Deegan, M.J.O.; Dobbyn, A.J.;<br />

Eckert, F.; Goll, E.; Hampel, C.; Hetzer, G.; Hrenar, T.; Knizia, G.; Köppl,<br />

C.; Liu, Y.; Lloyd, A.W.; Mata, R.A.; May, A.J.; McNicholas, S.J.; Meyer,<br />

W.; Mura, M.E.; Nicklass, A.; Palmieri, P.; Pflüger, K.; Pitzer, R.; Reiher,<br />

M.; Schumann, U.; Stoll, H.; Stone, A.J.; Tarroni, R.; Thorsteinsson, T.;<br />

Wang, M.; Wolf, A. MOLPRO, version 2008.2, a package <strong>of</strong> ab initio<br />

programs. see http://www.molpro.net., 2008.<br />

[36] Stone, A.J. Distributed polarizabilities. Mol. Phys. 1985. 56, 1065.<br />

[37] Ángyán, J.G.; Jansen, G.; Loos, M.; Hättig, C.; Hess, B.A. Distributed<br />

polarizabilities using the topological theory <strong>of</strong> atoms in molecules. Chem.<br />

Phys. Lett. 1994. 219, 267-273.<br />

[38] Hättig, C.; Jansen, G.; Hess, B.A.; Ángyán, J.G. Topologically partitioned<br />

dynamic polarizabilities using the theory <strong>of</strong> atoms in molecules. Can. J.<br />

Chem. 1996. 74, 976-987.<br />

[39] in het Panhuis, M.; Popelier, P.L.A.; Munn, R.W.; Ángyán, J.G. Distributed<br />

polarizability <strong>of</strong> the water dimer: field-induced charge transfer along the<br />

hydrogen bond. J. Chem. Phys. 2001. 114, 7951-7961.<br />

[40] Ponec, R.; Cooper, D.L. Anatomy <strong>of</strong> bond formation. Bond length<br />

dependence <strong>of</strong> the extent <strong>of</strong> electron sharing in chemical bonds. J. Mol.<br />

Struct. (Theochem) 2005. 727, 133-138.<br />

[41] Ponec, R.; Cooper, D.L. Anatomy <strong>of</strong> bond formation. Domain-averaged<br />

Fermi holes as a tool for the study <strong>of</strong> the nature <strong>of</strong> the chemical bonding in<br />

Li 2, Li 4, <strong>and</strong> F 2. J. Phys. Chem. A 2007. 111, 11294 -11301.<br />

[42] Ponec, R.; Roithová. Domain-averaged Fermi holes - a new means <strong>of</strong><br />

visualization <strong>of</strong> chemical bonds. Bonding in hypervalent molecules. Theor.<br />

Chem. Acc. 2001. 105, 383-392.<br />

[43] Ángyán, J.G. Correlation <strong>of</strong> bond orders <strong>and</strong> s<strong>of</strong>tnesses. J. Mol. Struct.<br />

(Theochem) 2000. 501-502, 379-388.<br />

[44] Ángyán, J.G.; Loos, M.; Mayer, I. Covalent bond orders <strong>and</strong> atomic valence<br />

indices in the topological theory <strong>of</strong> atoms in molecules. J. Phys. Chem. 1994.<br />

98, 5244-5248.<br />

[45] Silvi, B.; Savin, A. Classification <strong>of</strong> chemical bonds based on topological<br />

analysis <strong>of</strong> electron localization functions. Nature 1994. 371, 683-686.<br />

[46] Boys, S.F. Construction <strong>of</strong> some molecular orbitals to be approximately<br />

invariant for changes from one molecule to another. Rev. Mod. Phys. 1960.<br />

32, 296-299.<br />

[47] Boys, S.F. Localized orbitals <strong>and</strong> localized adjustment functions. In P.O.<br />

Löwdin, ed., Quantum theory <strong>of</strong> atoms, molecules <strong>and</strong> solid state, A tribute<br />

to John C. Slater, Academic Press: New York, chap. Localized Orbitals <strong>and</strong><br />

Localized Adjustment Functions. 1966. pp. 253-262.<br />

[48] Cioslowski, J. Isopycnic orbital transformations <strong>and</strong> localization <strong>of</strong> natural<br />

orbitals. Int. J. Quantum Chem. Symp. 24 1990. 38, 15-28.<br />

[49] Tschinke, V.; Ziegler, T. On the shape <strong>of</strong> spherically averaged Fermi-hole<br />

correlation functions in density functional theory. 1. Atomic systems. Can. J.<br />

Chem. 1989. 67, 460-472.<br />

[50] Luken, W.L.; Beratan, D.N. Localized orbitals <strong>and</strong> the Fermi hole. Theor.<br />

Chim. Acta 1982. 61, 265-281.<br />

[51] Luken, W.L. Properties <strong>of</strong> the Fermi hole <strong>and</strong> electronic localization. In Z.<br />

Maksic, ed., The concept <strong>of</strong> the chemical bond, Springer Verlag: Berlin, vol.<br />

2. 1990. pp. 287-320.<br />

[52] Buijse, M.A.; Baerends, E.J. An approximate exchange-correlation hole<br />

density as a functional <strong>of</strong> the natural orbitals. Mol. Phys. 2002. 100, 401-421.<br />

[53] Matito, E.; Salvador, P. in General Discussion. Faraday Discuss. 2007. 135,<br />

125-149.<br />

[54] Bader, R.F.W.; Gillespie, R.; MacDougall, P.J. A physical basis for the<br />

VSEPR model for molecular geometry. J. Am. Chem. Soc. 1988. 110, 7329-<br />

7336.<br />

[55] van Leeuwen, R. Key concepts in time-dependent density functional theory.<br />

Int. J. Mod. Phys. B. 2001. 15, 1969-2023.<br />

[56] Becke, A.D.; Edgecombe, K.E. A simple measure <strong>of</strong> electron localization in<br />

atomic <strong>and</strong> molecular systems. J. Chem. Phys. 1990. 92, 5397-5403.<br />

[57] Dobson, J.F. Interpretation <strong>of</strong> the Fermi hole curvature. J. Chem. Phys. 1991.<br />

94, 4328-4333.<br />

[58] Savin, A.; Becke, A.D.; Flad, J.; Nesper, R.; Preuss, H.; von Schnering, H.G.<br />

A new look at electron localization. Angew. Chem. Int. Ed. 1991. 30, 409-<br />

412.<br />

[59] Schmider, H.L.; Becke, A.D. Two functions <strong>of</strong> the density matrix <strong>and</strong> their<br />

relation to the chemical bond. J. Chem. Phys. 2002. 116, 3184-3193.<br />

[60] Kohout, M. A measure <strong>of</strong> electron localizability. Int. J. Quantum Chem.<br />

2004. 97, 651-658.<br />

[61] Geier, J. Radial exchange density <strong>and</strong> electron delocalization in molecules. J.<br />

Phys. Chem. A 2008. 112, 5187-5197.<br />

[62] Gori-Giorgi, P.; Ángyán, J.G.; Savin, A. Charge density reconstitution from<br />

approximate exchange-correlation holes. Can. J. Chem. 2009. 87, 1444-<br />

1450.<br />

[63] Johnson, E.R.; Becke, A.D. A post-Hartree-Fock model <strong>of</strong> intermolecular<br />

interactions. J. Chem. Phys. 2005. 123, 024101.<br />

[64] Ángyán, J.G. <strong>Electron</strong> localization <strong>and</strong> the second moment <strong>of</strong> the exchange<br />

hole. Int. J. Quantum Chem. 2009. 109, 2340-2347.<br />

[65] Ayers, P.W. A perspective on the link between the exchange(-correlation)<br />

hole <strong>and</strong> dispersion forces. J. Math. Chem. 2008. 46, 86-96.<br />

[66] Humphrey, W.; Dalke, A.; Schulten, K. VMD - Visual Molecular Dynamics.<br />

J. Mol. Graph. 1996. 14, 33-38.<br />

Received: 03 March, 2010 Revised: 08 May, 2010 Accepted: 08 May, 2010

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!