Sunteți pe pagina 1din 20

Chem Soc Rev

View Article Online

Published on 17 February 2014. Downloaded by FAC DE QUIMICA on 12/03/2014 17:40:01.

REVIEW ARTICLE

Cite this: DOI: 10.1039/c3cs60456j

View Journal

Conceptual DFT: chemistry from the linear


response function
Paul Geerlings,* Stijn Fias, Zino Boisdenghien and Frank De Proft
Within the context of reactivity descriptors known in conceptual DFT, the linear response function (w(r,r0 ))
remained nearly unexploited. Although well known, in its time dependent form, in the solid state physics and
time-dependent DFT communities the study of the chemistry present in the kernel was, until recently, relatively
unexplored. The evaluation of the linear response function as such and its study in the time independent form
are highlighted in the present review. On the fundamental side, the focus is on the approaches of increasing
complexity to compute and represent w(r,r0 ), its visualisation going from plots of the unintegrated w(r,r0 ) to an
atom condensed matrix. The study on atoms reveals its physical significance, retrieving atomic shell structure,
while the results on molecules illustrate that a variety of chemical concepts are retrieved: inductive and
mesomeric eects, electron delocalisation, aromaticity and anti-aromaticity, s and p aromaticity,. . .. The

Received 11th December 2013


DOI: 10.1039/c3cs60456j

applications show that the chemistry of aliphatic (saturated and unsaturated) chains, saturated and aromatic/antiaromatic rings, organic, inorganic or metallic in nature, can be retrieved via the linear response function,
including the variation of the electronic structure of the reagents along a reaction path. The connection of the

www.rsc.org/csr

linear response function with the concept of nearsightedness and the alchemical derivatives is also highlighted.

1 Introduction
General Chemistry (ALGC), Vrije Universiteit Brussel (Free University Brussels-VUB),
Pleinlaan 2, 1050 Brussel, Belgium. E-mail: pgeerlin@vub.ac.be;
Tel: +32 2 629 33 14

One of the themes of this special issue of Chemical Society


Reviews is Computational Concepts for a better understanding

Paul Geerlings (1949) is full Professor at the Vrije Universiteit


Brussel (Free University of Brussels VUB) where he obtained his
PhD and Habilitation, heading a research group involved in conceptual and computational density functional theory with applications in organic, inorganic, bio- and material chemistry. He is the
author or co-author of more than 440 publications in international
journals or book chapters. He organized several conferences in the
field (DFT 2003, Chemical Reactivity 2005).
Stijn Fias (1982) received his PhD from the Ghent University in 2011
From left to right: Prof. Paul Geerlings, Zino Boisdenghien,
under the mentorship of Prof. Patrick Bultinck. His dissertation
Dr Stijn Fias and Prof. Frank De Proft
examined the much debated multi-dimensionality of aromaticity using delocalisation indices. He is now a postdoctoral fellow at the Vrije
Universiteit Brussel (Free University of Brussels VUB). His research interests include the development, implementation, and
interpretation of DFT-based reactivity descriptors, with special interest in the application on Polycyclic Aromatic Hydrocarbons and
Graphene Nanoribbons. He is the author or coauthor of more than 25 research publications.
Zino Boisdenghien (1988) obtained his masters degree in theoretical physics from the Vrije Universiteit Brussel (Free University of
Brussels VUB) where he is currently working on his PhD in theoretical chemistry in the ALGC group headed by P. Geerlings. His work
focusses on the different aspects of the linear response function in (Conceptual) Density Functional Theory.
Frank De Proft (1969) obtained his PhD in Chemistry from the Free University of Brussels VUB in 1995. From 1995 to 1999, he was a
postdoctoral fellow at the Fund for Scientific Research Flanders (Belgium) during which he was a postdoctoral fellow in the group of
Professor R. G. Parr at the University of North Carolina in Chapel Hill. He is the author or co-author of more than 230 research
publications, mainly on conceptual DFT. His present work involves the development and/or interpretative use of DFT-based chemical
concepts to study the properties and reactivity of organic, inorganic, biochemical and solid state compounds.

This journal is The Royal Society of Chemistry 2014

Chem. Soc. Rev.

View Article Online

Published on 17 February 2014. Downloaded by FAC DE QUIMICA on 12/03/2014 17:40:01.

Review Article

of chemical structure and reactivity. This is a particularly well


chosen topic as it has been known for a long time that to
compute is not to understand.1 Parrs dictum is nowadays still
very topical because both on the wavefunctional theory side,
with its golden CCSD(T) standard,2 and on the DFT side,
with highly complicated functionals containing a considerable
number of parameters, amazingly accurate computational results
can be obtained for molecular properties, reaction energetics
(including even activation energies),. . ..3
It gives us an amazing series of accurate computational
results but to get the trends out of it, most chemists often go
back to traditional chemical concepts (resonance, electronegativity,4. . .) which, although often not sharply defined, have
helped chemists for decades. It is worth noting that in the early
days of quantum chemistry concepts and computations stood
ckel theory;58 also
closer together, a striking example being Hu
one of the reasons why this simple and elegant approach is still
taught in many introductory quantum chemistry courses.
Density Functional Theory has known an explosion in the
past two decades and replaced wavefunction theories as the
workhorse for most computations on medium and certainly
large systems free from peculiar intricacies. Besides this evolution, and parallel to it, the so called conceptual DFT916 (one
should better call it Chemical or Chemical Reactivity DFT) was
developed from the 1980s on. Based on Parrs illuminating
ideas, this approach can be casted in the study of the E = E[N,v]
functional, where E is the energy of the system, N is the number
of electrons and v the external potential felt by the electrons.
The basic idea is that in every reaction a molecule (or atom or
solid state system) is perturbed either in its number of electrons, its external potential, or both. The response of the
system, together with the magnitude of the perturbation, then
offers an answer to the question of how strongly the system is
perturbed, at the onset of the reaction, allowing comparative
studies for similar reactions if the so called non crossing rule is
obeyed. Crucial in this approach is the evaluation/study of the
response functions (dn+mE/qNndvm), the (mixed) functional derivatives of E with respect to N and v. At the present moment the
two derivatives with n + m = 1 (the electron density itself and the
electronic chemical potential)17 have been thoroughly investigated (vide infra) as well as most (2) of those with n + m = 2 (the
chemical hardness18 and the Fukui function19) and even some
with n + m = 3 (for a review, see Geerlings and De Proft20): the
dual descriptor,21,22 hyperhardness,23. . . More surprisingly the
second order functional derivative of E with respect to v at
constant N, (d2E/dv(r)dv(r 0 ))N, received little attention within
this context. This so-called linear response function w(r,r 0 )
is well known in theoretical physics and in the chemical
physics literature, essentially in its frequency dependent form
w(r,r 0 ;o).24,25 It appears in response theory, called the bread
and butter of theoretical physics.26 On the other hand it
appears in Time Dependent Density Functional Theory, where
the poles of the frequency dependent linear response function
correspond to excitation energies.2729 From this short excursion in Physics and Chemical Physics it is clear that the (static,
o = 0) linear response function has not been exploited yet.

Chem. Soc. Rev.

Chem Soc Rev

In Section 2 we will briefly mention which steps have been


taken by various authors and then situate the linear response
function in the complete series of response functions,
dn+mE/qnNdmv(r), highlighting the fundamental significance of
(dr(r)/dv(r 0 ))N, telling us how the density at a given point
in space changes when an external perturbation at an other
position r 0 is exerted.
After this introductory paragraph, three main lines will be
developed, partly in parallel, partly separately:
1. How to calculate the linear response function.
2. How to represent this awkward six dimensional kernel.
3. How to extract chemistry from it: can we retrieve classical
chemical concepts from the linear response function, and if so,
can we use the linear response function as a carrier of information on organic, inorganic and metallic systems.
The review will end with some miscellanea linking the linear
response function to some peculiar concepts of growing importance in chemistry/physics: alchemical derivatives for exploring
chemical space and the link between the linear response
function and the concept of nearsightedness.

2 The linear response function in the


context of conceptual DFT
2.1

General context

As stated above, conceptual DFT is about response functions of


the energy vs. perturbations in the number of electrons and/or
the external potential. These response functions are used either
as such or in the context of principles of which the electronegativity equalisation principle3033 and the Hard and Soft
Acids and Bases principle, both at global and local levels,3436
are the best known examples.
In Scheme 1 the dierent response functions are listed. The
two main ingredients of DFT are recognised in first order
response functions. (qE/qN)v is equal to the electronic chemical
potential m, the Lagrangian multiplier introduced when applying the variational principle to the E = E[r] functional with the
demand that the electron density should integrate to the total
number of electrons at all times. Using perturbation theory the
other first order derivative, (dE/dv(r))N, can easily be seen to be
equal to the electron density r(r) itself, the fundamental
quantity in DFT. Starting from m and r(r) the second order
derivatives appear: a global one (i.e., independent of r, as is m),
the chemical hardness (q2E/qN2)v; a local one (i.e. r dependent,
just as r(r)), the electronic Fukui function f (r) = (qr(r)/qN)v (note
that through Maxwells equations this can also be written and
computed as (dm/dv(r))N37,38); and finally a kernel (i.e. r and r 0
dependent), the linear response function w(r,r0 ) = (d2E/dv(r)dv(r0 ))N,
or to put it in another way (dr(r)/dv(r 0 ))N. The latter equation
clearly shows the utmost relevance of w(r,r0 ) for the chemist as it
tells us how a system changes its electron density at a given point r
when the external potential is changed at another point r 0 .
Some of the third order derivatives already received attention in the literature: the hyperhardness (q3E/qN3)v is expected
to be small.23 A lot of information is present in the so-called

This journal is The Royal Society of Chemistry 2014

View Article Online

Published on 17 February 2014. Downloaded by FAC DE QUIMICA on 12/03/2014 17:40:01.

Chem Soc Rev

Scheme 1

Energy derivatives and response functions in the canonical ensemble, dn+mE/qNndvm, (m + n r 3).

dual descriptor ((qf (r)/qN)v, (q2r(r)/qN2)v or alternatively (dZ/dv)),39


enabling a one shot picture of the electrophilic and nucleophilic regions in a molecule.21,40 The two remaining ones are
nearly unexplored in view of the complexity of exploring the
linear response function itself of which they are the N- and
v derivative respectively (with the latter expression being known
as the non-linear response function).
Please note that in the diagram, when going from left to right in
a given row, one passes from global to local quantities and to
kernels or even a second order kernel. As said before, where
nowadays quite accurate procedures are available for the descriptors mentioned above, the linear response function (and by extension its N- and v-derivatives) remained relatively unexplored in the
conceptual DFT context. It has been one of the main research
themes of the ALGC group in the last five years. Below we give a
short sketch of work published before or in parallel.
2.2

Review Article

Early work

Fundamental properties and formal solutions of the linear


response function or kernel can, as noted above, be found in
dierent papers. Senets study41,42 on the exact functional
relations between the linear and non-linear response functions
and the ground state electron density should be mentioned.
The elaboration on the properties and expressions for these
functions within a KohnSham formalism was presented by
Cohen et al.,43 Senet41,42 and Ayers.44,45 A summary of the most
important mathematical properties of the linear response
kernel can be found in a recent contribution by Liu et al.,46
one of the key issues being that the integration of w(r,r 0 ) over r
or r 0 gives zero.
There are, as also stated above, few numerical results on the
linear response function in the context of conceptual DFT, and
those available are typically obtained in a highly approximate
way. After all, the mutual atomatom polarizability, prs = qqr/qas
ckel theory (with qr the p charge of
of atoms r and s in early Hu
ckel MO parameter of atom s), is a highly
atom r and as the Hu
simplified form of an atom condensed linear response function
(as proposed by Coulson and Longuet-Higgins47 and the classical
treatise by Streitwieser48). See also the work at atoms in Molecular
resolution and its generalization to different resolutions by

This journal is The Royal Society of Chemistry 2014

Nalewajski and Korchowiec.16,49,50 Baekelandt et al.51 and Wang


et al.52 computed atom condensed linear response matrices
within the context of an electronegativity equalisation method
(EEM) based on the above mentioned principle of electronegativity equalisation. Some numerical data on an atom condensed charge response kernel using a modified definition of
atomic partial charges in a MO Coupled Perturbed HF approach
are also given in papers by Morita and Kato.53,54 In the work by
Langenaeker and Liu55 the explicit calculation of the response Dr(r)
upon a point charge perturbation in the case of atoms was evaluated
(cf. Section 3.2.2, see also the related work by Cedillo et al.56). We
also mention the work by Cioslowski and Martinov57 who calculated
approximate atomic softness matrices which can be interpreted
as the negatives of approximate linear response matrices
ngya
n should also be mentioned,
(cf. Section 5.1). The work by A
who correlated covalent bond orders and softness kernels58 and
connected linear response and electron delocalisation.59 Finally,
some recent studies by Korchowiec should be mentioned adopting
a molecular mechanics/forcefield ansatz to study the evolution of
selected elements of the condensed linear response function in a
molecular dynamics study as bond (formation) detectors.60,61
To summarize we can state that, despite the steps taken and
mentioned above, until a few years ago no direct, practical,
generally applicable and nearly exact approach was available for
the evaluation of w(r,r 0 ).

3 Evaluation and representation


of v(r,r 0 )
3.1

Numerical evaluation

The first approach to evaluate w(r,r 0 ) we undertook was a purely


numerical approach based on our expertise with the numerical
evaluation of the Fukui function.62 This approach was inspired
by numerous eorts by several authors, including ourselves, to
circumvent the tricky (qr(r)/qN)v evaluation involving the derivative with respect to the number of electrons. We therefore
opted to start from the alternative expression for f (r):


dm
f r
(1)
dvr

Chem. Soc. Rev.

View Article Online

Review Article

Chem Soc Rev

Published on 17 February 2014. Downloaded by FAC DE QUIMICA on 12/03/2014 17:40:01.

where now a functional derivative with respect to v(r) is at stake.


The methodology presented in (62) starts from a general problem,
namely the evaluation of dQ/dv(r), Q = Q[v] being a functional
of v(r). Suppose we can change the external potential using any
of P perturbations {wi(r)}Pi=1. Then from the definition of the
functional derivative we have up to first order


dQv
Qv wi   Qv dr
(2)
wi r
dvr
Expanding the functional derivative in a basis set {bj (r)}Kj=1,


K
X
dQv

qj bj r
(3)
dvr N j1
where qj are expansion coecients. Combining eqn (2) and
(3) we get

K
X
Qv wi   Qv
qj drbj rwi r
(4)
j1

which is a set of linear equations which can be written in matrix


form as
d = Bq

(5)

di = Q[v + wi]  Q[v]

(6)

Bij = drwi(r)bj (r)

(7)

where

and

In practice P is chosen to be larger than K and eqn (5) is


solved via least squares fitting, where P is varied until the result
converges. The perturbations themselves are point charge
perturbations (zi)
wi r

zi
jr  Ri j

(8)

and the expansion functions are uncontracted s- and p-type


Gaussians on each center.
As a first example we depict in Fig. 1 the Fukui function f +(r)
for H2CO showing the carbon atom to use the preferential site
for nucleophilic attack with the typical Dunitz angle63 for the
approaching nucleophile in agreement with experiment and
with finite dierence (qr/qN)v results.

Fig. 2 The dual descriptor f(2)(r) for CO: isosurface (0.02 a.u.) with positive
regions in red and negative regions in yellow. (Reprinted with permission
from Journal of Chemical Theory and Computation, (Copyright 2008),
American Chemical Society (ref. 62).)

In Fig. 2 we plot as a second example the functional


derivative of the hardness with respect to v(r) yielding the dual
descriptor


dZ
2
f r
(9)
dvr N
in the case of CO, clearly showing the one shot picture
representative power of f (2)(r) showing both electrophilic and
nucleophilic regions around a molecule, where the negative
values of the descriptor (prone to an electrophilic attack) are
mainly situated at the oxygen terminus of the CO-bond and
positive values (prone to nucleophilic attack) are located near
the carbon atom (region of the C lone pair in a Valence Bond
picture).39
This approach can now be extended to the second functional
derivative,38 where we immediately concentrate on the linear
response function w(r,r 0 ) being (d2E/dv(r)dv(r 0 ))N. Following the
ansatz of eqn (2) we can write

E[v + wi]  2E[v] + E[v  wi] = dr dr 0 w(r,r 0 )wi(r)wi(r 0 )


(10)
expanding w(r,r 0 ) now as
wr; r0

K
X

qkl bk rbl r0

(11)

k;l

we can again write down a matrix equation


d = Bq

(12)

where B is now a P  K2 matrix (with P the number of


perturbations) composed of the integrals over the various basis
functions and the external perturbations:

Bj,(k1)K+l = drbk(r)wj (r) dr 0 bl(r 0 )wj (r 0 )


(13)

Fig. 1 The Fukui functon f+(r) for H2CO in the plane perpendicular to the
molecular plane and containing the C and O nuclei. (Reprinted with
permission from Journal of Chemical Theory and Computation, (Copyright 2008), American Chemical Society (ref. 62).)

Chem. Soc. Rev.

with j = 1,. . .,P and k,l = 1,. . .,K and where q is a K2 dimensional
column matrix with elements q(k1)K+l = qkl (k,l = 1,. . .,K). At the
time when the first results were obtained, we opted for a representation of the six dimensional kernel by an atomatom condensed
linear response matrix with elements

wAB
dr dr0 wr; r0
(14)
VA VB

This journal is The Royal Society of Chemistry 2014

View Article Online

Chem Soc Rev


Table 1

Published on 17 February 2014. Downloaded by FAC DE QUIMICA on 12/03/2014 17:40:01.

H1
C
O
H2

Review Article

Linear response matrix elements wAB (in a.u.) for H2CO

H1

H2

1.21
1.19
0.42
0.40

4.82
2.45
1.19

3.28
0.42

1.21

which can be circumvented through the basis set approach by


computing (using numerical integration).

XX
qkl drbk r dr0 bl r0
(15)
wAB

We now consider how r(r) changes under an external


potential perturbation dv(r). Applying a self-consistent perturbational ansatz (which can be carried out at the HS or KS level)
we write the following perturbation expansion:
(1)
ji = j(0)
i + ji +   

represents the first order correction to the unperwhere j(1)


i
turbed orbital j(0)
i . The first order correction to the density,
r(1)(r) is then
r1 r 4

N=2
X

ji rj1 r

(18)

k2A l2B

(for more detailed discussions on the problem of the


representation/visualisation of the linear response kernel
vide infra).
The results, depicted in Table 1 for H2CO, are converged,
typically necessitating about 50 000 external perturbations,
each requiring a single point calculation. Note that the
dimensions of the non-condensed linear response function
are charge  volume2  potential1 (in analogy with charge 
volume1 for the first order functional derivative which is the
electron density). Upon condensation the dimension becomes
charge  volume1  potential1 in analogy with charge for
the first order functional derivative. All calculations were
performed in atomic units.
Even if the number of calculations can be reduced by
imposing symmetry, the vast computational eort most probably will prevent this approach to be used systematically on
middle sized and a fortiori large molecules. The results (both
condensed and uncondensed) are however useful as a benchmark because pushing both K and P for a given molecule
results in an in principle near exact value, to be obtained in a
single computational project, which may serve as a benchmark
for other more approximate methods.
We postpone the discussion of the numerical results until a
comparison within a larger series of molecules can be carried
out in the next paragraphs.

(17)

The evaluation of r(1)(r) necessitates the expansion coecients


(0)
Cai when writing j(1)
i (r) in terms of ja (r):
X
1
Cai j0
(19)
ji r
a r
a

where it is known that this summation only extends over the


(0)
64
unoccupied orbitals j(0)
a ,jb ,. . ..
Let us take the HF ansatz as our case study. Starting from the
HF equation
Fji = eiji

(20)

F=h+G

(21)

with

(respectively the one and two electron operators) we write the


perturbation expansions
F = F (0) + F (1) +   

(22)

(1)
ei = e(0)
i + ei +   

(23)

and obtain a first order perturbation equation


)j(1)
+ (h(1) + G(1))j(0)
= e(1)
j(0)
. (24)
(h(0) + G(0)  e(0)
i
i
i
i
i
Neglecting the influence of the external potential perturbation dv (= h(1)) on F(1) (through the perturbed orbitals) one gets
(1)
(1) (0)
(1) (0)
(h(0) + G(0)  e(0)
i )ji = h ji + ei ji

3.2

(25)

A simple perturbational approach

3.2.1 Methodology. Below we briefly describe a simple


perturbational approach, the result of which is known in the
literature at various places and sometimes with dierent names
(for a highly instructive account, see Cook64). This simple
derivation of the terminology which will be discussed in this
review will serve as a guide to the more advanced approach in
Section 3.3.
It starts from the hypothesis that a system is described by a
single Slater determinant (HartreeFock64 (HF) or KohnSham65
(KS)). Let us moreover suppose, for the sake of simplicity, that
the system is of the closed shell type and that all orbitals are real,
as well as the perturbation in the external potential. The density
can then be written as
rr 2

N=2
X

Introducing now the expansion eqn (19) and left-multiplying


with j(0)
we obtain
b


X
0
0
Cai drjb h0 G0  ei j0
a
a

0
0
1
0 0
 drjb h1 ji ei drjb ji

(26)

was chosen from among the unoccupied orbitals, the


As j(0)
b
last term is zero, so that introducing a Dirac bracket notation
for the dv term we get

X 
0
Cai e0
(27)
dab hbjdvjii
a  ei
a

or
ji2 r

This journal is The Royal Society of Chemistry 2014

(16)

(0)
Cbi(e(0)
b  ei ) = hb|dv|ii

(28)

Chem. Soc. Rev.

View Article Online

Review Article

Chem Soc Rev

or converting the index b to a


Cai 

hajdvrjii
0

ea  ei

(29)

Published on 17 February 2014. Downloaded by FAC DE QUIMICA on 12/03/2014 17:40:01.

The KS case can be treated in an analogous way, this time


neglecting the influence of the variation in the external potential
dvext on dvJ and dvxc, the Hartree and exchangecorrelation
potentials,
(30)
One then ends up with a first order (coupled) KS perturbed
equation


1
0
0
1
0
0 1
1 0
 r2 vext v0

v
ji dvext ji ei ji ei ji
xc
J
2
(31)
yielding

 


1
0
0
0
1
0
0

e
ji dvext ji e1 ji
 r2 vext vJ v0
xc
i
2
(32)
which has exactly the same structure as eqn (25) noticing that
h(1) = dvext, leading to identical expansion coefficients Cai as in
eqn (29).
Inserting these coecients in r(1)(r) we get

0 0
0 0
0
X 0
0 ja r dvr ji r
r1 r 4
ji rj0
(33)
a r dr
0
0
ea  ei
i;a
indicating that the functional derivative of r(r) with respect to
v(r 0 ) can be written as
X j rj rj r0 j r0
drr
i
a
a
i
4
;
0
dvr
e

e
a
i
i;a

(34)

where the superscripts (0) have been dropped.


As said, this simplified expression, known in the physics
community as the Independent Particle Approximation (IPA),66
can be found at various places in the literature (see e.g. Ayers67)
but as also stated in the introduction, its exploitation to study
the chemistry present in the kernel w(r,r 0 ) has to the best of
our knowledge hardly or not been described. We will compare
in Section 3.3: the expression (34) with the one without the
approximations introduced in the first order perturbation
expression. Let us finally mention that as we supposed that dvext
had no influence on vJ and vxc, eqn (34) can in fact be seen as the
exact expression for (dr(r)/dvKS(r 0 ))N, i.e. the functional derivative
of the electronic density with respect to the KS potential.
3.2.2 Atoms revisited. We now adopt this simple procedure
to investigate the characteristics of the kernel for the
simplest systems, isolated atoms, and in particular closed shell
atoms68 (the extension to the open shell case is presently under
investigation69).
In Fig. 3 we depict the results of the kernel evaluation using
a KS ansatz with 6-311+G* and an aug-cc-pVTZ basis at the PBE
level of theory for some light atoms, the cases He and Be being
the only ones which were previously studied (by Savin et al.70).

Chem. Soc. Rev.

Fig. 3 The linear response kernel for light atoms: contourplots for its
radial distribution for He and Be. (Reprinted with permission from Journal of
Chemical Theory and Computation, (Copyright 2013), American Chemical
Society (ref. 68).)

In particular we plot a radial distribution of the linear response


kernel r 2w(r,r 0 )r 0 2 for He and Be in case of a spherical potential
perturbation. The variables r and r 0 denote the distance to
the nucleus (located in the origin) for r and r 0 which are the
coordinates of the points where we investigate the change in
electron density dr(r) for a perturbation dv(r 0 ). These plots are
directly comparable and highly similar to those in Savins
paper. In the case of helium, two regions can be discerned,
a negative and a positive one, which get duplicated when
passing to Beryllium reflecting the shell structure of the atoms.
Turning to a simplified version of the representation, in
order to facilitate the interpretation, we plot in Fig. 4 a one
dimensional section of the radial distribution along the r-axis,
r 2w(r,0) (i.e. fixed r 0 , taking the value 0).
P
Za
In the absence of external fields, v(r) is given by  a
jr  R a j
(the summation extends over all nuclei a of a given system with
position Ra and charge Za, in the atomic case only one term
survives). Suppose now we have a positive perturbation dv(r).
This yields a less negative v(r) leading to electron depletion in
the vicinity of the nucleus. Writing

Dr(r) = dr 0 w(r,r 0 )dv(r 0 )


(35)
and realising dv(r 0 ) by a positive perturbation placed at the
nucleus,
dv(r 0 ) = xd(r 0  0),

(36)

with x A R , one obtains


Dr(r) = xw(r,0)

(37)

Fig. 4 A one dimensional section of the radial distribution of the linear


response kernel, r 2w(r,0), along the r axis for light atoms (He and Be).
(Reprinted with permission from Journal of Chemical Theory and
Computation, (Copyright 2013), American Chemical Society (ref. 68).)

This journal is The Royal Society of Chemistry 2014

View Article Online

Published on 17 February 2014. Downloaded by FAC DE QUIMICA on 12/03/2014 17:40:01.

Chem Soc Rev

Review Article

Fig. 6 The linear response kernel w(r,0) for Ar with a perturbation in the
origin and r varying in the (x,y)-plane. (Reprinted with permission from
Journal of Chemical Theory and Computation, (Copyright 2013), American
Chemical Society (ref. 68).)

A final representation is a 3D plot of w(r,0) where r is a vector


in the (x,y)-plane and the perturbation is placed at the origin
(Fig. 6). In the case for Argon, the shell structure with alternating
positive and negative values comes out nicely.
As an application we make use of the property that the
elements of the polarizability tensor aij, (i,j = x,y,z) can be
written in terms of an integral over w(r,r 0 ), e.g.71

axy =  dr dr 0 xw(r,r 0 )y
(38)

Fig. 5 Contourplots of r 2w(r,r 0 )r 0 2 and one dimensional sections along the


r-axis (r 2w(r,0)) for the first four noble gases (He, Ne, Ar, Kr). (Reprinted with
permission from Journal of Chemical Theory and Computation, (Copyright 2013), American Chemical Society (ref. 68).)

This expression was used as a test for the calculated values


of the linear response kernel: a comparison with high level
calculations is given in Fig. 7 for the series He, Be, F, Ne, Na+,
Mg2+, Mg, Cl, Ar, K+, Ca2+, Kr. It is clear that the calculated
w values capture the trends governing a throughout the
periodic table quite well. The correlation coecient (R2) for
the linear relationship amounts to 0.998. The deviations in the

This indicates that negative regions in the linear response


kernel imply a drop in the electron density corresponding to a
negative region in the one dimensional plots in Fig. 4 close to
the nucleus (located in the origin) accompanied at larger
distances by positive regions in view of the conservation of
electrons. This is consistent with the 2D plots in Fig. 3 where we
recognized for He an alternation of positive and negative
regions close to the r and r 0 axes with more oscillations as
Z increases, reflecting the atomic shell structure. It is also in
line with the fact that along the diagonal (r = r 0 ), the linear
response function is always negative: increasing v(r) leads to a
depletion of r(r) when r = r 0 . Note that this approach bears
similarity to the one proposed by Langenaeker and Liu.55
In this way the non-condensed kernel has been interpreted
through 1 and 2D plots. We finally remark that the shell
structure nicely pops up when passing through the first four
noble gases, both in the 1D and 2D plots (Fig. 5).

Fig. 7 Comparison of the calculated values of polarizability tensor components (eqn (38)) to high level calculated values. (Reprinted with permission from Journal of Chemical Theory and Computation, (Copyright 2013),
American Chemical Society (ref. 68).)

This journal is The Royal Society of Chemistry 2014

Chem. Soc. Rev.

View Article Online

Review Article

Chem Soc Rev

Table 2 Linear response matrix elements wAB (in a.u.) for H2CO, calculated
via eqn (34)

Published on 17 February 2014. Downloaded by FAC DE QUIMICA on 12/03/2014 17:40:01.

H1
C
O
H2

H1

H2

1.22
0.68
0.34
0.19

4.35
2.99
0.68

3.67
0.34

1.22

magnitude should be traced back to the approximate nature of


the w values.
3.2.3 Molecules
3.2.3.1 H2CO revisited. The approach described in Section 3.2.1
can easily be extended to molecules where condensation to an
atomatom matrix72,73 will in many cases be a natural simplification for the representation of w (see however Section 4.2.3 for a
more detailed representation).
In Table 2 we represent the atomatom matrix for H2CO.62
Comparison with the matrix resulting from a numerical evaluation yields a correlation coefficient R2 E 0.96 between the matrix
elements of the two methods with a slope of nearly 1 and a very
small intercept.
Similar results were obtained for other simple molecules: high
correlation, with a very small intercept, the only dierence being
the slope which varies between 1 and 2 this can be expected to be
avoided when passing to the softness kernel s(r,r0 ) (vide infra), as is
the case in the replacement of the Fukui function f (r) by the local
softness s(r) for intermolecular comparison.11,25
3.2.3.2 A first idea on Chemistry from the linear response
function. In Table 3 we depict the atomatom linear response
matrix for ethanal.
As already seen in the case of H2CO, the diagonal elements
are all negative, which is easy to grasp by analogy with the
arguments for the r = r 0 diagonal in the non-condensed case for
atoms (Fig. 3 and 5). These results are comparable to the
negative diagonal values in Morita and Katos atom condensed
Charge Response Kernel.53,54 If a positive perturbation is placed
on the nucleus of an atom, electron depletion in its neighbourhood occurs leading to a negative wAA element. These negative
elements (in absolute value increasing with the size of the atom)
should be compensated by positive o-diagonal elements in view
of one of the basic properties of the linear response function,46

drw(r,r 0 ) = 0,
(39)
which should be obeyed in its condensed form as well.

The link between the linear response function and the


polarizability of the electron density (cf. Section 3.2.2) is clearly
manifested in the values of the wC1C1, wOO and wOC1 elements
which, when compared to the corresponding values for ethanal
(3.42, 2.18 and 2.73) are considerably larger in absolute
value. In particular the wOC1 value (2.73) is much larger in
ethanal than in ethanol (0.87) reflecting the much higher
polarizability of a CQO vs. a CO bond. Another trend is the
decreasing value of the wAB elements when atoms A and B are
more distant.
These results gave us confidence to explore the transmission
of electronic eects (inductive and mesomeric) through (in the
first place) organic molecules, which will be discussed in
Section 4.1.1.
3.3 The coupled perturbed KohnSham (CPKS) approach and
the introduction of (non integrated) molecular plots
3.3.1 Theory. A third road towards the linear response
function is the refinement of the approach presented in Section
3.1, i.e. taking into account G(1) (HF) or dvJ and dvxc (KS)74 in the
perturbation equations. In what follows we first introduce the
Coulombic part of G(1) (HF) and the dvJ term in KS which are
completely equivalent.
Starting again from the HF case, eqn (27) now becomes


X 
0
0
0
dab hbjdvjii  drjb rJ 1 ji r (40)
Cai e0
a  ei
a

where again dv = h(1). The integral containing J (1) reads


!

1
X j0
j r2 jj r2
0
0
j i r1
2 dr1 jb r1 2 dr2
r12
j
4

C1

H1

C2

H2

H3

H4

C1 4.2080
H1 0.6900 1.2365
O
2.7289 0.3338 3.7827
0.5067 0.1507 0.3522 3.3676
C2
H2 0.0966 0.0024 0.1707 0.7813 1.1423
H3 0.0966 0.0024 0.1707 0.7813 0.0372 1.1413
H4 0.0892 0.0572 0.0264 0.7950 0.0532 0.0532 1.0738

Chem. Soc. Rev.

XX
j

dr1 dr2 ji r1 jb r1

1 0
1
j r2 jj r2
r12 j

(41)

Ccj ibjjc

Eqn (40) then yields



X 
X
0
dab hbjdvjii  4
Cai e0
Ccj ibj jc
a  ei
a

(42)

j;c

or


X
0
0
Cbi eb  ei
hbjdvjii  4
Ccj ibj jc

(43)

j;c

or
X

Table 3 Linear response matrix elements wAB (in a.u.) for ethanal, calculated via eqn (34)

Ccj

h

e0
c  ej

i
dbc dij 4ibj jc hbjdvjii 0

(44)

j;c

which, upon dropping the second term in the double summation, again yields
(0)
Cbi(e(0)
b  ei ) = hb|dv|ii

(45)

i.e. eqn (28).


Introducing the matrix M with elements
Mib,jc = (ec  ej)dijdbc + 4(ib| jc)

(46)

This journal is The Royal Society of Chemistry 2014

View Article Online

Chem Soc Rev

Review Article

Eqn (44) can be written as


MC = dV

(47)

C = M1dV

(48)

or

vxc, namely its r dependence) one can write in analogy with


the evaluation of the J (1) term in HF (eqn (41)):

0
0
0
0
drjb dvxc ji drjb rvxc r dr  vxc rji r

dvxc r
0
0
drr0 ji r
drjb r dr0
(55)
drr0

d2 Exc
0
0
drr0 ji r
drjb r dr0
drrdrr0

Published on 17 February 2014. Downloaded by FAC DE QUIMICA on 12/03/2014 17:40:01.

i.e.
Cbi 

X

M1


ib;jc

dVjc

(49)

j;c

where we used the relation vxc = (dExc/dr(r))N. Writing


dr(r 0 ) as
X
0 0
0
drr0 r1 r0 4
Caj j0
(56)
a r jj r

Introducing this expression in r(1) gives (cf. eqn (18))


r1 r 4

j;a

ji rji r

XX
i

4

we obtain
0
0
Cbi ji rjb r

X X
i;b

1

(50)

hbjdvxc jii 4

0
0
Caj drdr0 jb rji r

j;a

(57)

0
0
j rjb rdVjc
ib;jc i

d2 Exc
0
j0 r0 jj r0

drrdrr0 a

j;c

with
dVjc = h j|dv|ci

(51)

Substituting this term in eqn (46) we obtain the general


expression for the M matrix elements in the KS case
Mib,jc = (ec  ej)dijdbc + 4(ib|jc) + 4(ib|fxc(r,r 0 )|jc)

yielding

(58)

where we defined fxc(r,r 0 ) as


wr; r0

X X

drr
4
M1 ib;jc ji rjb rjj r0 jc r0
dvr0
i;b j;c
(52)

where the superscripts (0) have been dropped.


It is clear that eqn (52) in combination with eqn (44)
can also be used in a KS context when neglecting the first
order corrections of dvext upon dvxc, yielding the so-called
Random Phase Approximation (RPA) to the linear response
function.
Note that when in eqn (46) the last term is dropped, eqn (52)
reduces to
wr; r0  4

XX
i;b

4

X
i;b

j;c

1
j rjb rjj r0 jc r0 dij dbc
ec  ej i

1
j rjb rji r0 jb r0
eb  ei i

fxc r; r0

(53)
4

Caj bijaj

j;a

i.e. eqn (34) is retrieved, yielding (dr(r)/dvKS(r 0 ))N.


When in HF the correction to the exchange operator is taken
into account, it is easily seen that in eqn (46) an additional term
appears:
(54)

whereas in the KS case the influence of dvext on vxc should be


included. In general terms (i.e. without specifying the nature of

This journal is The Royal Society of Chemistry 2014

(59)

which can be inverted as in eqn (48) oering via eqn (52) a


possibility to evaluate w(r,r 0 ) using a coupled perturbed HF or
KS ansatz (see ref. 53 for an early approach). Note that in KS
the term discussed previously can, in analogy to eqn (57) be
written as

d2 EJ

drr
hbjdvJ jii b

i
0
drrdrr

drr

i
j r  r0 j

X
1
0
0
0
j0 r0 jj r0
Caj drdr0 jb rji r
4
0j a
j
r

r
j;a

Mib,jc = (ec  ej)dijdbc + 4(ib| jc)  2(ic| jb)

d2 Exc
drrdrr0

Ccj ibjjc

(60)

j;c

retrieving eqn (41).


So we are now in a position to study w(r,r 0 ) both at the HF
and KS level, neglecting or not the influence of external
potential perturbations on the Coulomb and/or exchange
correlation operators/potential.
3.3.2 The dierence between v and vKS. We finally mention
that the dierence between wKS and w can be written in a

Chem. Soc. Rev.

View Article Online

Review Article

Chem Soc Rev

compact way, found in Solid State Physics texts as a Dyson


equation24,26 (see also ref. 45).

wr; r0 wKS r; r0 dr00 dr000 wKS r; r00



1
00 000

fxc r ; r wr000 ; r0
jr00  r000 j

Published on 17 February 2014. Downloaded by FAC DE QUIMICA on 12/03/2014 17:40:01.

(61)

This can also be written terms of the perturbation in the KS


potential dvKS considering an equivalent non-interacting
system

dr(r) = dr 0 wKS(r,r 0 )dvKS(r 0 )


(63)
As we now have

with (cf. eqn (55) and (60))




1
0
0
fxc r; r drr0
dvJ dvxc dr
j r  r0 j

(64)

(65)

we obtain

dr0 wr; r0 dvext r0 dr0 wKS r; r0 dvext r0

dr

00

w  wKS = wKS * K * w

(70)

yielding

which is proven in the notation adopted here as follows:


From the definition of the response function w(r,r0 ) (Scheme 1),
it follows that

dr(r) = dr 0 w(r,r 0 )dvext(r 0 )


(62)

dvKS = dvext + dvJ + dvxc

Simplifying this expression by combining the Coulomb and


the exchangecorrelation contributions into K and introducing

the shorthand * to denote integrals like dr00 wKS(r,r00 )K(r00 ,r 0 ) as


wKS * K, we get the Dyson equation

(71)
or
(72)
or
w1 = wKS1  K

(73)

In principle w can be obtained from w0 if the functional Exc[r]


is known. For an extension to the time dependent case we refer
to Casida and Huix-Rotllant,75 Van Leeuwen,76 Ghosh,77. . . in
which the time dependent perturbation expressed in the frequency domain is dvext(r,o). In the limit of o - 0 the results in
the present paper are retrieved.
3.3.3 Examples. In Fig. 8 we give an example, to the best of
our knowledge one of the first, if not the first, plots of w(r,r 0 ) at
this level for benzene, separated into its s and p components
(vide infra) where a comparison is made between the noninteracting picture (dvJ = dvxc = 0, eqn (34)) and the full CPKS.



1
0 00
00
fxc r ; r drr
jr0  r00 j
(66)

or

dr0 wr; r0  wKS r; r0 dvext r0





1
0 00

f
r
;
r

wr00 ; r000 dvext r000


dr0 dr00 dr000 wKS r; r0 0
xc
jr  r00 j
(67)
Upon interchanging r 0 0 0 and r 0 in the last integral, one
obtains

dr0 wr; r0  wKS r; r0 dvext r0





1
000 00
r
;
r

wr00 ; r0 dvext r0

f
dr0 dr00 dr000 wKS r; r000 000
xc
jr  r00 j
(68)
yielding an integral equation involving the two response
functions:

wr; r0  wKS r; r0 dr00 dr000 wKS r; r000





Chem. Soc. Rev.


1
000 00

f
r
;
r

wr00 ; r0 dvext r0
xc
jr000  r00 j
(69)

Fig. 8 The unintegrated linear response function for benzene, separated


in its s (a, b) and p (c, d) compoments. The sigma linear response plots
show the response in s-electron density in the molecular plane, when the
perturbation is placed at the nucleus of the top carbon atom (a, b) and the
linear response of the p-electron density in a plane 0.5 atomic units above
the molecular plane, when the perturbation is placed symmetrically at the
same height as the plane, 0.5 atomic units above and below the nucleus of
the top carbon atom (c, d). The plots show the linear response using the
independent particle approximation (a, c) and using the full exchange and
correlation (b, d; using LDA).78 (Reprinted with permission from Journal of
Physical Chemistry, (Copyright 2013), American Chemical Society (ref. 79).)

This journal is The Royal Society of Chemistry 2014

View Article Online

Published on 17 February 2014. Downloaded by FAC DE QUIMICA on 12/03/2014 17:40:01.

Chem Soc Rev

Review Article

It is clear that qualitatively the two plots are similar: the


oscillating behaviour near the atoms shows up in a more
pronounced way in the interacting (s) picture, the o, m, p
distributions (vide infra) are highly similar to the p distribution.
(The analytical evaluation of the dv term has been performed at
the LDA level.)
Note that a similar strategy can be followed74 in evaluating/
plotting the Fukui function where one starts from the
relationship




dm
deF
f r

(74)
dvr N
dvr N
where eF denotes the orbital energy of the frontier MO ff.
Writing

eF ff

 r2 vKS

ff
(75)
2
or

def 2 ff

 r2 vKS

dff hff jdvKS jff i


2

4 Chemistry from v(r,r 0 )

def = hff|dvext|ff i

(77)

def
jff rj2
dvr

(78)

or

i.e. the HOMOLUMO approximation to the Fukui functions.


Introducing the changes in r brought by dvext in the contributions of dvJ and dvxc one obtains in analogy to eqn (52) juncto
eqn (58)
X
drdr0 Cbj ff rff r
def jff j2 4
j;b


1
0

r;
r

fj r0 fb r0

f
xc
jr  r0 j
using eqn (49) for Cbj as
X

M1 jb;kc hkjdvext jci
Cbj 

(see also Michalak et al.81) and therefore avoids the cumbersome derivation with respect to N.
In Fig. 9 f +and the LUMO for H2CO are depicted, where it is
seen that the two profiles qualitatively agree but that e.g. the
Dunitz angle is better reproduced in Fig. 9(a).

(76)

It is easily seen that the first term equals zero and that the
direct dvext contribution to the second term is

Fig. 9 The Fukui function f+and the LUMO for H2CO. (Reprinted with permission from Journal of Physical Chemistry, (Copyright 2013), American Chemical
Society (ref. 79).)

(79)

(80)

k;c

one obtains for


X X 

 1
def
0

jff rj2 4

f
ff

r;
r

xc

jb M jb;kc
0
dvr
jr  r j
k;c j;b

4.1

Concepts

In the present paragraph, we try out if chemical concepts can be


retrieved by the linear response function; the work equation for
these relatively large systems is eqn (34) unless stated otherwise.
4.1.1 Inductive and mesomeric eects82. On the basis of
the results for H2CO we further explore the capability of the
linear response kernel to reproduce inductive and mesomeric
eects. We therefore evaluated the transmission of a dv perturbation through a carbon chain by comparing saturated and
unsaturated (conjugated) carbon chains: hexan-1-ol and hexa1,3,5-trien-1-ol (Fig. 10(a)). Placing the perturbation at the
oxygen yields in the saturated chain a monotonous decrease
in wOX (X = C1, C2,. . .) in parallel with the inductive eect, which
as also seen in wOX, dies out after two or three bonds. In other
words, the linear response function nicely, and without any
preconception, reproduces the inductive eect. Note that an
exponential fall-o was found in the computed values (R2 for an
exponential curve = 0.982) (vide infra, Section 5.1) An analogous
picture emerges for hexan-1-amine hexan-1-amine (Fig. 10(b)).
Passing to the unsaturated system the figures show an alternating behaviour with maxima at C2, C4 and C6 and minima at
C1 (the OH, NH2 bearing carbon), C3, C5 corresponding to
mesomeric active and passive atoms respectively as seen below:

 fk rfc r
(81)
The second term gives an explicit expression for the relaxation term denoted before by Yang, Parr and Pucci80 as
X @f r
k
fk
(82)
f r jff rj2 4
@N
vr
k

This journal is The Royal Society of Chemistry 2014

The mesomeric active atoms 2, 4, 6 carry a negative charge in


one of the four canonical forms, the 1, 3, 5 atoms dont. The
NH2/OH ratio of B0.5/0.4 for C4 is in line with the ratio of the
soR Hammetts Structure Activity parameters of 0.5/0.44.83
The charge evolution, depicted in the linear response function,

Chem. Soc. Rev.

View Article Online

Published on 17 February 2014. Downloaded by FAC DE QUIMICA on 12/03/2014 17:40:01.

Review Article

Fig. 10 Linear response elements (a) wOX (with X = C1, . . ., C6) between
oxygen and the various Carbon atoms and (b) wNX between nitrogen and the
various carbon atoms. (Reprinted with permission from Journal of Physical
Chemistry Letters, (Copyright 2010), American Chemical Society (ref. 82).)

therefore dies out for the mesomeric passive atoms, just as in the
alkane chains due to the presence of an inductive eect only,
whereas for the mesomeric active atoms, the eect remains
consistently large even after 6 bonds, albeit tempered to some
extent by the inductive eect working in parallel. Similar results
with an inverted activity pattern were obtained for mesomeric
acceptor substituents (CHO and CRN) as opposed to the
mesomeric donor substituents OH and NH2.
4.1.2 Electron delocalisation84. The previous results suggest that in cyclic systems the linear response could be used as
a measure of electron delocalisation, as later on also emphangya
n.59 In Fig. 11 we depict the linear response
sized by A
matrix elements wC1Ci (i = 2, 3,. . .6) for cyclohexane, benzene,
cyclohexanol and phenol in order again to distinguish between
cyclic alkenes and cyclic conjugated alkenes leading to aromatic

Chem Soc Rev

systems. C1 denotes the number of the atom on which the


substituent is placed.
Fig. 11 clearly illustrates that in cyclohexane and cyclohexanol w
decreases along the ring until a minimum is reached at the para
position and then increases at the ortho position. This exponential
decrease is typical for the inductive eect (vide supra). In the
aromatic systems, a zig-zag curve is obtained as in Section 4.1.1
with maxima at C2, C4, C6 which are mesomerically active (carrying a
negative charge in one of the canonical forms) and minima at C3
and C5, the mesomerically inactive atoms. Similar conclusions have
been drawn from mesomeric acceptor substituents (CHO). Going
back to Fig. 11, more precisely benzene, a pattern is found, evocative
of results obtained in another context. Bader et al.85 have shown
that electron delocalisation is linked to the spatial distribution of
the Fermi hole. The delocalisation function or index proposed on
this basis in the framework of the AIM theory shows that the
delocalisation between para carbons in benzene is higher than
`86 is retrieved
between meta atoms. This (1,4) effect as termed by Sola
in the linear response function, hinting that the linear response
function might be a valuable indicator of aromaticity.
4.1.3 Aromaticity87,88. As stated above, the (1,4) eect recognised in the linear response function plots of aromatic systems is
reminiscent of the Para Delocalisation Index. Derived from AIM,
` et al. as a quantitative measure of the
it was put forward by Sola
number of electrons delocalised between two atoms (in casu ipso
and para atoms of an aromatic ring), and was investigated as a
potential index of aromaticity. In general, the delocalisation index
between two atoms A and B is defined as:

dA; B 2
Gxc r; r0 drdr0
(83)
A B

where Gxc represents the exchange correlation density and the


double integration is performed over the atomic basins of atoms A
` and coworkers came to an armative conclusion for
and B. Sola
d(1,4) as an indicator of aromaticity by an extensive survey on
polycyclic aromatic rings representing a broad range of aromatic
character. In Fig. 12 we depict the 16 non-equivalent six-membered
`. We indeed concentrated on six-membered
rings studied by Sola
rings for reasons of simplicity, evaluating the Para Linear
Response (PLR) index as an average 1/3(w14 + w25 + w36).
Fig. 13 shows that all six rings considered display the
zig-zag pattern with an outspoken (1,4) peak. In Fig. 14 the
strength of the (1,4) interaction is plotted versus the PDI,
yielding an excellent correlation with R2 = 0.96. In fact it
can be shown that using the resolution of the identity and
ld approximation and using
certain approximations (the Unso
non-overlapping basins for atoms A and B when integrating
w(r,r 0 ), optimal for 1,4 interaction) an expression of the type
1
wdA;B  dA; B:
D

Fig. 11 The atom condensed linear response elements wC1Ci (with i = 2,


3,. . .6) between the carbon atoms C1 and Ci of the cyclohexane, benzene,
hexanol and phenol rings. All values are in a.u. (Reprinted with permission
from Chemical Physics Letters, (Copyright 2010), Elsevier (ref. 84).)

Chem. Soc. Rev.

(84)

holds. Here D is an average of all ei  ea, being negative so that


the linear response function and the para delocalisation index
positively correlate.
4.1.4 Antiaromaticity88. The linear response function being
related to aromaticity, it was tempting to see if anti-aromaticity

This journal is The Royal Society of Chemistry 2014

View Article Online

Published on 17 February 2014. Downloaded by FAC DE QUIMICA on 12/03/2014 17:40:01.

Chem Soc Rev

Review Article

Fig. 14 The linear correlation plot between PLR and PDI (extracted from
ref. 86) values (in a.u.) for all six-membered rings of Fig. 12. (Reprinted with
permission from the RSC, (Copyright 2012) (ref. 87).)

Table 4 Linear response of C4H4 and C8H8, together with the s and p
contributions (vide infra, Section 4.1.5). The shorter (double) bonds are
located between atoms 1 and 2 in both molecules

w11

Fig. 12 Structures of the planar polycyclic hydrocarbons studied. (Reprinted


with permission from the RSC, (Copyright 2012) (ref. 87).)

is also regained via the response function. Prototypes of antiaromaticity studied were cyclobutadiene (D2h) and cyclooctatetraene (D4h). Table 4 shows that the wA,B values between atoms
on opposite sides of the rings, w13 and w15 for cyclobutadiene

w12

w13

w14

C4H4
wAB
3.182
s+p
wAB
1.357
s
wAB
1.825
p

1.947
0.201
1.746

0.124
0.140
0.016

0.582
0.512
0.070

C8H8
wAB
2.866
s+p
1.188
wAB
s
wAB
1.677
p

1.290
0.172
1.118

0.139
0.086
0.053

0.486
0.010
0.476

w15

w16

w17

w18

0.038
0.003
0.035

0.148
0.013
0.136

0.139
0.086
0.053

0.080
0.299
0.218

and cyclooctatetraene respectively, are smaller than the response


with the neighbours of these atoms, w12/w14 and w14/w16 respectively.
Table 4 shows that the pattern of these molecules, well
known to be anti-aromatic, is nearly opposite to that of their
aromatic counterparts. A zig-zag shape is displayed in the
response, but now with a local minimum instead of a maximum

Fig. 13 The atom condensed linear response elements wC1Ci (with i = 2, 3,. . .6) between the carbon atoms C1 and Ci of the six membered rings in the
molecules in Fig. 12. All values are in a.u.

This journal is The Royal Society of Chemistry 2014

Chem. Soc. Rev.

View Article Online

Published on 17 February 2014. Downloaded by FAC DE QUIMICA on 12/03/2014 17:40:01.

Review Article

Chem Soc Rev

at the atom on the opposite side of the ring. The following


rules of thumb can thereby be formulated to identify a
molecule as aromatic or anti-aromatic using the linear response
function:88
 aromatic molecules have a typical zig-zag shape, with
a local maximum between atoms on the opposite side of the
aromatic ring.
 anti-aromatic molecules also have a zig-zag shape, but
with a local minimum between atoms on opposite sides of the
aromatic ring. They have their highest value between neighbouring atoms on a double bond.
Both rules can be understood in terms of mesomerically
active or inactive atoms as sketched in Section 4.1.1.
4.1.5 r and p aromaticity88. Digging further in aromaticity,
s and p aromaticity can be recovered by combining eqn (34)
and (14) to
wA;B  4

XX
i

4

XX
i

1
ea  ei

fi rfa rdr

fa r0 fi r0 dr

(85)

1
SA SB ;
ea  ei ia ai

where SAia and SBai are the overlap matrix elements for atoms A
and B respectively. Concentrating now on occupied MO fi, one
can write an orbital contribution to wA,B as
wA;B
4
i

X
a

1
SA SB ;
ea  ei ia ai

(86)

where in the case of planar molecules, the wis can be


recollected in s and p parts of the linear response. In Table 5
we (again) depict the results for benzene, now split up into s
and p contributions. The s contribution shows a monotonous
decrease along the ring, with a minimum at the para position,
as was the case for cyclohexane. The typical increase in the para
position is only due to the p contribution.
Further examples of s aromaticity will be analysed in
Section 4.2.3 (all metal rings).

Table 5 Linear response of benzene, together with the s and p contributions (Reprinted with permission from the RSC, (Copyright 2013)
(ref. 88).)

wAB
s+p
wAB
s
wAB
p

w11

w12

w13

w14

w15

w16

2.668
1.128
1.540

0.537
0.226
0.311

0.281
0.074
0.206

0.503
0.020
0.483

0.281
0.074
0.206

0.537
0.226
0.311

4.2

Substrates

In the following paragraphs some applications of the aforementioned concepts will be described substratewise.
4.2.1 Organic rings and cycloaddition reactions88. Saturated
and unsaturated (in casu aromatic and anti aromatic) organic
rings have been treated in previous paragraphs; in this section we
go one step further and investigate the linear response function
along a reaction profile passing through an aromatic transition
state: the DielsAlder reaction between butadiene and ethene.
Fig. 15 nicely shows that the maximum of the PLR is found exactly
at the transition state with a value of 0.497 a.u. which corresponds
to 99% of the PLR value of benzene.
It is reassuring that the linear response function can be used,
not only at the minima of the potential energy surface, but also for
the transition state and along the reaction coordinate. As a second
example we consider the [p2s + p2s + p2s] symmetry allowed89
trimerisation of acetylene to benzene (although with a relatively
high energy barrier90). We dissected the PLR into s and p
components. The shape of the total PLR curve has a maximum
around the transition state, then decreases and increases again to
the final product (Fig. 16). The s and p contributions to the PLR
reveal that the aromaticity of the transition state is almost entirely
s in nature and that the p contribution only becomes significant
when the s framework is returned to a more localised state. It is
pleasing to note that these findings are in complete agreement
with the results from ring current maps and Nucleus Independent
Chemical Shift (NICS) analysis: the transition state of this trimerisation has no p character at all. Only after the relocalisation of the
s electrons to form the CC s-bonds, does the p ring current start
to appear.91,92

Fig. 15 The para linear response (PLR) versus the reaction coordinate for
the DielsAlder reaction (IRP in amu1/2 Bohr) The PLR is given in percentages of the PLR value of benzene. (Reprinted with permission from the RSC,
(Copyright 2013) (ref. 88).)

Fig. 16 The para linear response (PLR) versus the reaction coordinate for
the acetylene trimerisation (IRP in amu1/2 Bohr) The PLR is given in
percentages of the PLR value of benzene. (Reprinted with permission from
the RSC, (Copyright 2013) (ref. 88).)

Chem. Soc. Rev.

This journal is The Royal Society of Chemistry 2014

View Article Online

Published on 17 February 2014. Downloaded by FAC DE QUIMICA on 12/03/2014 17:40:01.

Chem Soc Rev

As a whole, the success of the PLR in these two reactions


` and coworkers on the performance
parallels the result by Sola
of the para-delocalization index as an indicator of aromaticity.
As an afterthought, this could have been expected on the basis
of eqn (84).93,94
4.2.2 Inorganic rings87. The same methodology as for studying
the aromaticity of organic rings (use of equation eqn (34) in
condensed form) was used to study some inorganic analogues
of benzene, the most prominent example being borazine,
B3N3H6, extended to the oxygen analogue boroxine, B3O3H3,
and its phosphorus analogue s-triphosphatriborin, B3P3H6:

Concentrating on borazine, the atom condensed linear response


function presents a remarkable behaviour. In Fig. 17 it is seen
that the typical resonance pattern, observed in aromatic systems, is recovered if one of the nitrogen atoms is chosen as the

Fig. 17 Atom-condensed linear response values (in a.u.) for (a) borazine,
(b) boroxine and (c) s-triphosphatriborin. The values between a reference
ring atom and all other ring atoms are given. (Reprinted with permission
from the RSC, (Copyright 2012) (ref. 87).)

This journal is The Royal Society of Chemistry 2014

Review Article

Fig. 18 The four Kekule


structures used in the valence bond study of
ref. 95. X stands for NH, O and PH in the cases of borazine, boroxine and
s-triphosphatriborin, respectively.

reference atom; whereas a purely inductive behaviour, exponential decay of electron delocalisation with internuclear
distance, is observed if one of the boron atoms is chosen as
the reference atom. This corresponds with a dual aromatic
character of borazine: the nitrogen atoms have a lone pair at
their disposal which can delocalise over the ring, resulting in a
LRF pattern resembling the aromatic situation. This is in line
with the valence bond (VB) analysis by Engelberts et al.95
indicating that the weight of this resonance structure is 90%
(cf. the VB structures in Fig. 18). The boron atoms do not have
such a lone pair, resulting in an LRF pattern which merely
reflects a pure inductive eect. The overall eect is that borazine
is non-aromatic as it is now generally classified (but with a still
ongoing debate).96
The related case of boroxine now presents a situation in VB
terms amounting to 96% contribution of structure III. The
resonance energy lowers (from 61.6 kJ mol1 in borazine
to 27.6 kJ mol1 in boroxine) in agreement with the lower
(1,4)-element of the LRF (1 being the nitrogen or oxygen atom
respectively) which decreases from 0.30 a.u. for borazine to
0.17 a.u. for boroxine. The phosphorus analogue has a (1,4)element of 0.87 a.u. and is thus expected to have a higher
resonance energy, which indeed turns out to be the case
(132.6 kJ mol1 vs. 161.4 kJ mol1 for benzene). The contribution of the dierent canonical forms is also more benzene-like.
All in all the qualitative trend in atom condensed linear

response functions parallels the contributions of possible Kekule


structures to the resonance energy in the context of VB theory.
4.2.3 All metal rings79. As a final series of ring systems we
studied the aromaticity of the all-metal Al42 ring, which as a
subunit in MAl4 clusters (M = Li, Na,. . .) has received great
interest in recent years with an ongoing debate on its classification as s, p or s + p aromatic.
In Fig. 19 we represent the unintegrated linear response kernel,
again obtained starting from eqn (34), at that time the first
molecular LRF plot to the best of our knowledge. We compare it
immediately with the archetype of an organic aromatic molecule,
benzene (Fig. 19(c) and (d)). The plots show the response in the s
electron density in the molecular plane where a perturbation in the
external potential, dv(r) = d(r  r0 ) (cf. eqn (36)), is imposed at
the nucleus of the top aluminium atom (a) and the response in the
p-electron density in a plane 0.5 a.u. above the molecular plane
when the perturbation is placed symmetrically, 0.5 atomic units
above and below the nucleus of the top aluminium atom.
In the s plot of Al42 we observe positive and negative oscillations
close to the nucleus (cf. the shell structure of the atom, Section 3.2.2).

Chem. Soc. Rev.

View Article Online

Review Article

Chem Soc Rev


Table 6

2

Published on 17 February 2014. Downloaded by FAC DE QUIMICA on 12/03/2014 17:40:01.

Al4
Al3Ge
Al2Ge2
AlGe3+
Ge42+

The linear response and d(1,3) of Al42 to Ge42+

OLR

d(1,3)

8.117
6.422
5.627
5.666
6.558

5.526
4.346
3.795
3.614
4.169

2.591
1.271
1.832
2.052
2.389

0.897
0.847
0.755
0.822
0.814

opposite atoms in mixed four-membered rings. It should be


noted anyway that without exception, w13 and w24 are greater than
between adjacent atoms, despite the larger internuclear distance.
This is in accordance with the studies on aromatic organic and
inorganic rings discussed above. The data in Table 6 yields an
aromatic order Al42 > Al3Ge Z Al2Ge2 r AlGe3+ o Ge42+, with
in all cases a s aromatic character, the s-OLR being about 65% of
the total value. The sequence itself follows the one proposed by
` et al.101,102 and corresponds to d(1,3) and NICSzz(1) values
Sola
except for a minor deviation.

Fig. 19 The unintegrated linear response (as obtained from eqn (34)) of
Al42 (a, b) compared with benzene (c, d) (cf. Fig. 8(a) and (c)). The linear
response plots of Al42 show the response of s-electron density in the
molecular plane, when the perturbation is imposed at the nucleus of the
top aluminium atom (a) and the linear response of the p-electron density in
a plane 0.5 atomic units above the molecular plane, when the perturbation
is placed symmetrically at the same height as the plane, 0.5 atomic units
above and below the nucleus of the top aluminium atom (b). The linear
response plots of benzene show the response in s-electron density in the
molecular plane, when the perturbation is placed at the nucleus of the top
carbon atom (a) and the linear response of the p-electron density in a
plane 0.5 atomic units above the molecular plane, when the perturbation is
placed symmetrically at the same height as the plane, 0.5 atomic units
above and below the nucleus of the top carbon atom (b). (Reprinted
with permission from Journal of Physical Chemistry, (Copyright 2013),
American Chemical Society (ref. 79).)

The delocalised nature of the linear response shows up, but it is


more pronounced in the s density, indicating that the system is
mainly s-aromatic. This also shows up in the atom-condensed
w13 value, where the s contribution (5.53 a.u.) is twice as large
as the p contribution (2.59 a.u.). The plots for benzene show
that the s response decreases in magnitude when moving away
from the origin of the perturbation. The s LRF is, as compared
to the 3 position in Al42, less extended in space around the
para carbon in benzene, showing a more localised s framework
in benzene. The p linear response for benzene shows a preferential connection between C1 and its ortho and para positions. Compared to benzene, the p linear response is much less
outspoken in Al42. All in all, the LRF clearly shows that Al42 is
mainly s aromatic, supporting the claims put forward in the
literature.97100 We finally mention that by systematically replacing the aluminium atoms by germanium, the series Al42,
Al3Ge, Al2Ge2, AlGe3+, Ge42+ is obtained which was proposed
` et al.101,102 to be double s and p aromatic. In Table 6 we
by Sola
depict the Opposite Linear Response OLR = 1/2 (w13 + w24) in
order to circumvent problems of different possible choices for

Chem. Soc. Rev.

5 Miscellaneous
5.1 The linear response function and the concept of
nearsightedness
In recent years attention has been paid by a few authors to the
concept of nearsightedness of electronic matter (NEM),103,104
introduced by Kohn, and its connection to conceptual DFT.
This intricate principle quintessentially states that for systems
with many electrons at constant electronic chemical potential,
the change in the electron density at a given point r0, |Dr(r0)|,
induced by an external potential perturbation at position r 0 ,
Dv(r 0 ), with r 0 outside a sphere with radius R centered at r0 will
always be smaller than a maximum value Dpr0 ; R, no matter
how large the perturbation is. To put it otherwise, the particle
density at r0 cannot see any perturbation beyond R with an
accuracy beyond Dr(r0,R). Kohn and Prodan proved that for one
dimensional gapless model systems, the decay of Dp as a
function of R (i.e. upon increasing |r  r0|) follows power laws
(in the case of real 3D systems corresponding to metals). For
gapped systems, i.e. systems with a hardness Z dierent from
zero, an exponential decay can be expected. This suggests that
in the molecular world, where Z is typically dierent from zero,
this situation can be expected to be predominant. Numerical
data on real atoms and molecules for this fundamental result
were lacking, although one feels that this basic physical theorem
must be connected with the issue of transferability of atoms and
functional groups from one molecule to another, the basis of a
rational build up of organic chemistry. In 2008 Bader wrote a
vigorous paper105 in which he pointed out that the nearsightedness
of matter has been used frequently but in a dierent way by
chemists (say for example in an atoms-in-molecules context) to
discuss/exploit the concept of transferability.
Let us now return to the NEM principle in the context of
conceptual DFT. When looking at the fundamental quantities in
Kohns argument one recognizes Dr(r0) and Dv(r0 ) suggesting that
the linear response function connects both. However, one of Kohns

This journal is The Royal Society of Chemistry 2014

View Article Online

Chem Soc Rev

Review Article

assumptions is the condition of constant electronic chemical


potential m (and not of constant N) as is done when evaluating
w(r,r 0 ). We therefore introduce along with w(r,r 0 ) the quantity
(dr(r)/dv(r 0 ))m, which turns out to be minus the softness kernel
s(r,r 0 ) appearing in a natural way as the counterpart to the
linear response function when working in an open system with
grand potential O[m,v] as the basic functional instead of E[N,v].
Both functionals are related through the equation
Published on 17 February 2014. Downloaded by FAC DE QUIMICA on 12/03/2014 17:40:01.

O = E  mN.

(87)

Transferring the rules of partial dierentiation with dierent


constraints to the case of functional dierentiation (with dierent
constraints) one obtains



   
drr
drr
dr
dm

(88)
dvr0 N
dvr0 m
dm v dv N
or
w(r,r 0 ) = s(r,r 0 ) + s(r) f (r 0 ),

(89)

introducing the local softness s(r), or


wr; r0 sr; r0

srsr0
;
S

(90)

introducing the total softness, and retrieving the famous


ParrBerkowitz relation.106
rdenas et al. followed Prodan and Kohn stating that in an
Ca
open-system representation (at constant m) reactivity indicators
that depend on more than one point are nearsighted.107 This
should be compared with our identification of w(r,r 0 ) as nearsighted in saturated systems (Section 4.1.1). It suggests that in
these cases the second term to w(r,r 0 ) in eqn (90) does not
counteract the nearsightedness of s(r,r 0 ), in agreement with the
fact that for larger aliphatic chains the Fukui function or local
softness for a given atom will be approximately constant.
In the case of polyenes the totally dierent behaviour
of w should be ascribed to a cancellation of the nearsighted
character of s(r,r 0 ) by the variation of f (r) and s(r) along the
unsaturated chain due to conjugation where charge transfer is
at work.
5.2 A direct application of the linear response function when
dv(r) results from a nuclear charge variation: exploring
chemical space through alchemical derivatives
We finally give a brief account of the application of the linear
response function when the external potential perturbation
is known. Suppose for example that we are interested in the
so-called alchemical derivatives,108,109 i.e. derivatives of the
atomic or molecular energy with respect to one or more nuclear
charges. In analogy with m in Scheme 1, it can be termed
the electronic alchemical potential. The first derivative is easily
written,



@Eel
dEel @vr
1
;
(91)
dr
drrr
jr  Ra j
@Za N
dvr @Za
which is nothing else than the electronic part of the Molecular
Electrostatic Potential110 (MEP) at position of nucleus a.

This journal is The Royal Society of Chemistry 2014

Passing to a mixed second order derivative one gets




 2


@ Eel
d2 Eel
@vr @vr0
drdr0
@Za @Zb N
dvrdvr0 N @Za @Zb
(92)

1
1
0
0
;
drdr wr; r
jr  Ra j jr  Rb j
in which the linear response function appears. In the same vein
as before, the second derivative
 2 



@ Eel
1
@rr
dr
(93)
@Za2 N
jr  Ra j @Za N
can be identified as the electronic part of the alchemical
hardness.
It was recently shown that in a Coupled Perturbed HF ansatz
up to second order these derivatives can be evaluated111 and used
for eciently exploring chemical (compound) space.112 It is recomforting to see that this fundamental problem with which chemists
are confronted nowadays (surfing through Chemical (compound)
Space113 in search of compounds with optimal properties) can be
done in an ecient way in which again the linear response
function plays a predominant role. Chemistry is again at stake!

6 Conclusions
Within the diversity of reactivity descriptors oered to the
chemist by conceptual DFT, the linear response function
remained nearly unexploited in this context. Although well
known in say the solid state physics community, mostly in its
frequency dependent form, as well as in time-dependent DFT,
its evaluation as such and its study in a time independent
context are highlighted in the present review. The focus on the
fundamental side is on the approaches of increasing complexity
to compute w(r,r 0 ) and several representations going from plots
of the unintegrated w(r,r 0 ) to an atom condensed matrix. The
study on atoms reveals its physical significance revealing atomic
shell structure and alternating positive and negative regions.
The results on molecules illustrate that a variety of chemical
concepts are retrieved: inductive and mesomeric eects,
electron delocalisation, aromaticity and anti-aromaticity, s and
p aromaticity,. . .. The applications show that the chemistry of
aliphatic (saturated and unsaturated) chains, saturated and
aromatic/anti-aromatic rings, organic, inorganic or metallic in
nature, can be retrieved via the linear response function,
including the variation of the electronic structure of the
reagents along a reaction path.
Nevertheless work needs to be done: the eciency of the
computation should be improved and combined with ecient
representations to make routine application and interpretation
aordable. Finally the intriguing relation of the linear response
function to the principle of the nearsightedness of electronic
matter should be studied in a more quantitative way, the
softness kernel finally being candidate of a thorough understanding of one of the key concepts in (organic) chemistry:
transferability. The chemical future of the linear response
function looks bright!

Chem. Soc. Rev.

View Article Online

Review Article

Published on 17 February 2014. Downloaded by FAC DE QUIMICA on 12/03/2014 17:40:01.

Acknowledgements
The authors would like to thank many of their collaborators in
the ALGC group for discussions and contributions to dierent
aspects of the linear response function: Dr Nick Sablon, Dr Tim
Fievez, Dr Fabiana Da Pieve, . . .. They had the privilege to host
two outstanding scientists in the field of DFT, Prof. Paul Ayers
(in 2007 and on various later occasions) and Prof. Weitao Yang
(in 2010 on the occasion of his Solvay Chair in Chemistry at the
international Solvay Institutes of Chemistry and Physics). Both
contributed substantially to the development of the computational techniques described in this review. With Prof. Paul Ayers it
has also been a pleasure to discuss the intricacies of the concept
of nearsightedness. Dr Aron Cohens contributions in the field
were equally important and appreciated. Dr Miquel TorrentSucarrat is gratefully acknowledged for his programming activity
` for
on atom condensed versions of functions, Prof. Miquel Sola
the collaboration on electron delocalisation and aromaticity, as is
Prof. Christian Van Alsenoy for his assistance with the computation and analysis of the linear response function on atoms. The
authors would also like to thank Prof. Gabriel Merino for suggesting the all-metal systems for investigation. Dr Robert Balawender
opened the gate to exploring the chemical space through alchemical derivatives. Finally PG and FDP which to thank the FWO and
the VUB for continuous support to the ALGC group over the past
20 years, specially mentioning the Strategic Research Program
awarded by the VUB to the ALGC group which started on January
1, 2013. SF also wishes to thank the FWO for financially supporting his postdoctoral research at the ALGC group.

References
1 R. G. Parr, Density Functional Theory in Chemistry, in
Density Functional Methods in Physics, ed. R. Dreizler and
J. Providencia, Plenum Press, 1985.
2 G. D. Purvis and R. J. Bartlett, J. Chem. Phys., 1982, 76,
19101918.
3 Y. Zhao and D. G. Truhlar, Acc. Chem. Res., 2008, 41,
157167.
4 L. Pauling, The Nature of the Chemical Bond and the
Structure of Molecules and Crystals: An Introduction to
Modern Structural Chemistry, Cornell University Press,
Ithaca, New York, 1960.
ckel, Z. Phys., 1930, 60, 423456.
5 E. Hu
ckel, Z. Phys., 1931, 70, 204286.
6 E. Hu
ckel, Z. Phys., 1931, 72, 310337.
7 E. Hu
ckel, Z. Phys., 1932, 76, 628648.
8 E. Hu
9 R. G. Parr and W. Yang, Annu. Rev. Phys. Chem., 1995, 46,
701728.
10 H. Chermette, J. Comput. Chem., 1999, 20, 129154.
11 P. Geerlings, F. De Proft and W. Langenaeker, Chem. Rev.,
2003, 103, 17931874.
12 F. De Proft and P. Geerlings, Chem. Rev., 2001, 101,
14511464.
13 P. W. Ayers, J. S. M. Anderson and L. J. Bartolotti,
Int. J. Quantum Chem., 2005, 101, 520534.

Chem. Soc. Rev.

Chem Soc Rev

zquez, J. Mex. Chem. Soc., 2008, 52, 310.


14 J. L. Ga
15 S. B. Liu, Acta Phys.-Chim. Sin., 2009, 25, 590600.
16 R. F. Nalewajski, J. Korchowiec and A. Michalak, in Density
Functional Theory IV, Topics in Current Chemistry, ed.
R. Nalewajski, Springer, Berlin, Heidelberg, 1996, vol. 183,
pp. 25141.
17 R. Parr, R. A. Donnelly, M. Levy and W. E. Palke, J. Chem.
Phys., 1978, 68, 38013807.
18 R. G. Parr and R. G. Pearson, J. Am. Chem. Soc., 1983, 105,
75127516.
19 R. G. Parr and W. Yang, J. Am. Chem. Soc., 1984, 106,
40494050.
20 P. Geerlings and F. De Proft, Phys. Chem. Chem. Phys., 2008,
10, 30283042.
, J. Phys. Chem. A,
21 C. Morell, A. Grand and A. Toro-Labbe
2005, 109, 205212.
, Chem. Phys. Lett.,
22 C. Morell, A. Grand and A. Toro-Labbe
2006, 425, 342346.
23 P. Fuentealba and R. G. Parr, J. Chem. Phys., 1991, 94,
55595564.
24 E. K. U. Gross and W. Kohn, Phys. Rev. Lett., 1985, 55,
28502852.
25 R. G. Parr and W. Yang, Density-Functional Theory of Atoms
and Molecules, Oxford University Press, London and
New York, 1989.
26 R. M. Martin, Electronic Structure: Basic Theory and Practical
Methods, Cambridge University Press, Cambridge, New York,
2004, Appendix D.
27 E. Gross and W. Kohn, Time-Dependent Density-Functional
Theory, Academic Press, 1990, vol. 21, pp. 255291.
28 M. E. Casida, in Recent Advances in Density Functional
Methods, Part I, ed. D. P. Chong, World Scientific Pub Co
Inc., Singapore, 1995, p. 155.
29 A. D. Laurent and D. Jacquemin, Int. J. Quantum Chem.,
2013, 113, 20192039.
30 R. T. Sanderson, Science, 1952, 116, 4142.
31 W. J. Mortier, S. K. Ghosh and S. Shankar, J. Am. Chem.
Soc., 1986, 108, 43154320.
32 K. A. Van Genechten, W. J. Mortier and P. Geerlings,
J. Chem. Phys., 1987, 86, 50635071.
33 P. Bultinck, W. Langenaeker, P. Lahorte, F. De Proft,
P. Geerlings, M. Waroquier and J. P. Tollenaere, J. Phys.
Chem. A, 2002, 106, 78877894.
34 J. L. Gazquez and F. Mendez, J. Phys. Chem., 1994, 98,
45914593.
35 P. Geerlings and F. De Proft, Int. J. Quantum Chem., 2000,
80, 227235.
36 M. Torrent-Sucarrat, F. De Proft, P. W. Ayers and
P. Geerlings, Phys. Chem. Chem. Phys., 2010, 12,
10721080.
37 P. W. Ayers, F. De Proft, A. Borgoo and P. Geerlings,
J. Chem. Phys., 2007, 126, 224107.
38 N. Sablon, F. De Proft, P. W. Ayers and P. Geerlings,
J. Chem. Phys., 2007, 126, 224108.
39 T. Fievez, N. Sablon, F. De Proft, P. W. Ayers and
P. Geerlings, J. Chem. Theory Comput., 2008, 4, 10651072.

This journal is The Royal Society of Chemistry 2014

View Article Online

Published on 17 February 2014. Downloaded by FAC DE QUIMICA on 12/03/2014 17:40:01.

Chem Soc Rev

40 P. W. Ayers, C. Morell, F. De Proft and P. Geerlings,


Chem.Eur. J., 2007, 13, 82408247.
41 P. Senet, J. Chem. Phys., 1996, 105, 64716489.
42 P. Senet, J. Chem. Phys., 1997, 107, 25162524.
43 M. H. Cohen, M. V. Ganduglia-Pirovano and J. Kudrnovsky,
J. Chem. Phys., 1995, 103, 35433551.
44 P. W. Ayers and R. G. Parr, J. Am. Chem. Soc., 2001, 123,
20072017.
45 P. W. Ayers, Theor. Chem. Acc., 2001, 106, 271279.
46 S. Liu, T. Li and P. W. Ayers, J. Chem. Phys., 2009,
131, 114106.
47 C. A. Coulson and H. C. Longuet-Higgins, Proc. R. Soc.
London, 1947, 194, 3960.
48 A. Streitwieser, Molecular Orbital Theory for Organic
Chemists, Wiley, 1961.
49 J. Korchowiec, H. Gerwens and K. Jug, Chem. Phys. Lett.,
1994, 222, 5864.
50 R. F. Nalewajski and J. Korchowiec, Charge Sensitivity
Approach to Electronic Structure and Chemical Reactivity,
World Scientific Pub Co Inc., 1997.
51 B. G. Baekelandt, W. J. Mortier, J. L. Lievens and R. A.
Schoonheydt, J. Am. Chem. Soc., 1991, 113, 67306734.
52 C.-S. Wang, D.-X. Zhao and Z.-Z. Yang, Chem. Phys. Lett.,
2000, 330, 132138.
53 A. Morita and S. Kato, J. Am. Chem. Soc., 1997, 119, 40214032.
54 A. Morita and S. Kato, J. Phys. Chem. A, 2002, 106, 39093916.
55 W. Langenaeker and S. Liu, THEOCHEM, 2001, 535,
279286.
n, A. Aizman, J. Andre
s
56 A. Cedillo, R. Contreras, M. Galva
and V. S. Safont, J. Phys. Chem. A, 2007, 111, 24422447.
57 J. Cioslowski and M. Martinov, J. Chem. Phys., 1994, 101,
366370.
ngya
n, THEOCHEM, 2000, 501502, 379388.
58 J. A
ngya
n, Curr. Org. Chem., 2011, 15, 36093618.
59 J. G. A
60 A. Stachowicz, M. Rogalski and J. Korchowiec, J. Mol.
Model., 2013, 19, 41634172.
61 A. Stachowicz and J. Korchowiec, J. Comput. Chem., 2013,
34, 22612269.
62 N. Sablon, F. De Proft, P. W. Ayers and P. Geerlings,
J. Chem. Theory Comput., 2010, 6, 36713680.
63 H. B. Burgi, J. D. Dunitz, J. M. Lehn and G. Wip,
Tetrahedron, 1974, 30, 15631572.
64 D. B. Cook, Handbook of Computational Quantum Chemistry,
Oxford University Press, 1998.
65 W. Kohn and L. J. Sham, Phys. Rev., 1965, 140, A1133A1138.
66 C. Jamorski, M. Casida and D. R. Salahub, J. Chem. Phys.,
1996, 13, 51345147.
67 P. W. Ayers, Faraday Discuss., 2007, 135, 161190.
68 Z. Boisdenghien, C. Van Alsenoy, F. De Proft and
P. Geerlings, J. Chem. Theory Comput., 2013, 9, 10071015.
69 Z. Boisdenghien, S. Fias, C. Van Alsenoy, F. De Proft and
P. Geerlings, submitted.
70 A. Savin, F. Colonna and M. Allavena, J. Chem. Phys., 2001,
115, 68276833.
71 R. G. Parr and W. Yang, Density-Functional Theory of Atoms and
Molecules, Oxford University Press, New York, 1989, p. 212.

This journal is The Royal Society of Chemistry 2014

Review Article

`,
72 M. Torrent-Sucarrat, P. Salvador, P. Geerlings and M. Sola
J. Comput. Chem., 2007, 28, 574583.
` and P. Geerlings,
73 M. Torrent-Sucarrat, P. Salvador, M. Sola
J. Comput. Chem., 2008, 29, 10641072.
74 W. Yang, A. J. Cohen, F. De Proft and P. Geerlings, J. Chem.
Phys., 2012, 136, 144110.
75 M. Casida and M. Huix-Rotllant, Annu. Rev. Phys. Chem.,
2012, 63, 287323.
76 R. Van Leeuwen, Int. J. Mod. Phys. B, 2001, 15, 19692023.
77 S. K. Ghosh, Time-Dependent Density Functional Theory
of Many-Electron Systems, in Chemical Reactivity Theory: A
Density Functional View, ed. P. K. Chattaraj, CRC Press, 2009.
78 S. Fias, Z. Boisdenghien, F. De Proft and P. Geerlings,
in preparation.
79 S. Fias, Z. Boisdenghien, T. Stuyver, M. Audired,
G. Merino, P. Geerlings and F. De Proft, J. Phys. Chem. A,
2013, 117, 35563560.
80 W. Yang, R. G. Parr and R. Pucci, J. Chem. Phys., 1984, 81,
28622863.
81 A. Michalak, F. De Proft, P. Geerlings and R. F. Nalewajski,
J. Phys. Chem. A, 1999, 103, 762771.
82 N. Sablon, F. De Proft and P. Geerlings, J. Phys. Chem. Lett.,
2010, 1, 12281234.
83 M. Smith and J. March, Marchs Advanced Organic
Chemistry: Reactions, Mechanisms, and Structure, Wiley,
6th edn, 2007.
84 N. Sablon, F. De Proft and P. Geerlings, Chem. Phys. Lett.,
2010, 498, 192197.
85 X. Fradera, M. A. Austen and R. F. W. Bader, J. Phys. Chem.
A, 1999, 103, 304314.
`, Chem.Eur. J.,
86 J. Poater, X. Fradera, M. Duran and M. Sola
2003, 9, 400406.
` and P. Geerlings, Phys.
87 N. Sablon, F. De Proft, M. Sola
Chem. Chem. Phys., 2012, 14, 39603967.
88 S. Fias, P. Geerlings, P. Ayers and F. De Proft, Phys. Chem.
Chem. Phys., 2013, 15, 28822889.
89 T. L. Gilchrist and R. C. Storr, Organic Reactions and Orbital
Symmetry, Cambridge University Press, 1979.
s, Chem. Phys. Lett., 2005,
90 J. C. Santos, V. Polo and J. Andre
406, 393397.
91 I. Morao and F. P. Cosso, J. Org. Chem., 1999, 64, 18681874.
92 R. W. A. Havenith, P. Fowler, L. Jenneskens and E. Steiner,
J. Phys. Chem. A, 2003, 107, 18671871.
`, THEOCHEM,
93 E. Matito, J. Poater, M. Duran and M. Sola
2005, 727, 165171.
`, J. Comput.
94 F. Feixas, E. Matito, J. Poater and M. Sola
Chem., 2008, 29, 15431554.
95 J. Engelberts, R. W. A. Havenith, J. H. van Lenthe,
L. W. Jenneskens and P. W. Fowler, Inorg. Chem., 2005,
44, 52665272.
96 W. Shen, M. Li, Y. Li and S. Wang, Inorg. Chim. Acta, 2007,
360, 619624.
97 E. Steiner and P. W. Fowler, J. Phys. Chem. A, 2001, 105,
95539562.
98 P. W. Fowler, R. W. A. Havenith and E. Steiner, Chem. Phys.
Lett., 2001, 342, 8590.

Chem. Soc. Rev.

View Article Online

Published on 17 February 2014. Downloaded by FAC DE QUIMICA on 12/03/2014 17:40:01.

Review Article

99 P. W. Fowler, R. W. A. Havenith and E. Steiner, Chem. Phys.


Lett., 2002, 359, 530536.
100 R. Islas, T. Heine and G. Merino, J. Chem. Theory Comput.,
2007, 3, 775781.
nez-Halla, E. Matito, J. Poater and M. Sola
`,
101 F. Feixas, J. Jime
J. Chem. Theory Comput., 2010, 6, 11181130.
`, Theor.
102 F. Feixas, E. Matito, M. Duran, J. Poater and M. Sola
Chem. Acc., 2011, 128, 419431.
103 W. Kohn, Phys. Rev. Lett., 1996, 76, 31683171.
104 E. Prodan and W. Kohn, Proc. Natl. Acad. Sci. U. S. A., 2005,
102, 1163511638.
105 R. F. W. Bader, J. Phys. Chem. A, 2008, 112, 1371713728.
106 M. Berkowitz and R. G. Parr, J. Chem. Phys., 1988, 88,
25542557.

Chem. Soc. Rev.

Chem Soc Rev

rdenas, N. Rabi, P. W. Ayers, C. Morell, P. Jaramillo


107 C. Ca
and P. Fuentealba, J. Phys. Chem. A, 2009, 113, 86608667.
108 O. A. von Lilienfeld, R. D. Lins and U. Rothlisberger, Phys.
Rev. Lett., 2005, 95, 153002.
109 O. A. von Lilienfeld, Int. J. Quantum Chem., 2013, 113,
16761689.
110 R. Bonaccorsi, E. Scrocco and J. Tomasi, J. Chem. Phys.,
1970, 52, 52705284.
111 M. Lesiu, R. Balawender and J. Zachara, J. Chem. Phys.,
2012, 136, 034104.
112 R. Balawender, M. A. Welearegay, M. Lesiuk, F. De Proft
and P. Geerlings, J. Chem. Theory Comput., 2014, 9,
53275340.
113 P. Kirkpatrick and E. Ellis, Nature, 2004, 432, 823865.

This journal is The Royal Society of Chemistry 2014

S-ar putea să vă placă și