Sei sulla pagina 1di 14

Coordination Chemistry Reviews 257 (2013) 196209

Contents lists available at SciVerse ScienceDirect

Coordination Chemistry Reviews


journal homepage: www.elsevier.com/locate/ccr

Review

Consistent descriptions of metalligand bonds and spin-crossover in inorganic


chemistry
Kasper P. Kepp
Technical University of Denmark, DTU Chemistry, DK 2800 Kongens Lyngby, Denmark

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197
2. General methodological considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 198
2.1. Electron correlation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 198
2.2. Complex density functionals as a path toward accuracy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 198
2.3. Energies and basis sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199
2.4. Identifying the congurations that are lowest in energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199
3. Systematic effects in electronic structure calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200
3.1. HF exchange . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200
3.2. Self-interaction and non-locality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201
3.3. Dispersion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201
3.4. Relativistic effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201
3.5. Solvent effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202
3.6. Zero-point energies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203
3.7. Entropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203
4. Insight gained from benchmarking of metalligand bonds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204
4.1. Correlation effects in metalligand bonds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204
4.2. Bond strengths of M L diatomics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204
5. Metalligand bonds and spin crossover . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
5.1. General aspects of spin crossover . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
5.2. HF exchange and spin states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
6. Larger-system effects not described in this review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
7. Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
List of abbreviations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208

a r t i c l e i n f o a b s t r a c t

Article history: Density functional theory (DFT) is today the unchallenged tool for routinely obtaining molecular infor-
Received 30 November 2011 mation on chemical stability, reactivity, and electronic structure across the Periodic Table. The chemical
Received in revised form 12 April 2012 bond is the fundamental unit of molecular structure and reactivity, and thus, large-scale DFT studies of
Accepted 15 April 2012
inorganic systems in catalysis and bioinorganic chemistry rely directly on the ability to balance corre-
Available online 21 April 2012
lation effects in the involved bonds across the s-, p-, and d-blocks. This review concerns recent efforts
to describe such bonds accurately and consistently across the s-, p-, and d-blocks. Physical effects and
Keywords:
ingredients in functionals, their systematic errors, and approaches to deal with them are discussed, in
DFT
Systematic effects
order to identify broadly applicable methods for inorganic chemistry.
Metalligand bonds 2012 Elsevier B.V. All rights reserved.
Electron correlation
Spin
Catalysis
Bioinorganic chemistry

E-mail address: kpj@kemi.dtu.dk

0010-8545/$ see front matter 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.ccr.2012.04.020
K.P. Kepp / Coordination Chemistry Reviews 257 (2013) 196209 197

1. Introduction One can also distinguish a priori and a posteriori functionals: the
a priori functionals fulll physical bounds on the energy and density
The last two decades have been characterized by the massive or its derivatives without violating fundamental laws [3739], risk-
diffusion of density functional theory (DFT) [13] into all elds of ing poor accuracy for specic, perhaps important classes of systems
chemistry, a development that already in 1998 led to Walter Kohn in exchange for a sound platform for further improvement. Rep-
sharing the Nobel Prize in chemistry with John Pople for their pio- resentative functionals include PBE [40], TPSS [41], and to some
neering efforts in DFT and general quantum chemistry, respectively extent their hybrid counterparts, PBE0 [42] and TPSSh (contain-
[4]. ing only one parameter, the amount of HF exchange) [43]. The a
The current status 14 years later is that DFT is the unchal- posteriori functionals use several (>1) tted parameters and prag-
lenged method for describing electronic structure in medium- and matically and efciently reproduce vital observables, guaranteeing
large-size molecular systems [5,6]. Its unique combination of com- good accuracy for systems and properties within the data sets used
putational speed and accuracy affects most areas of molecular for developing the functionals, but possibly restricted from expand-
science from molecular physics and chemistry to biochemistry, ing this accuracy beyond their parameterization range [44]. Some
nanochemistry, and materials science. As an illustration, state-of- representative functionals are B3LYP [33,4547] of the hybrid type
the-art DFT is an invaluable complementary and supporting tool for and M06 [48], a meta hybrid.
spectroscopic methods that have greatly contributed to shaping the Fig. 1 shows the occurrence of topical words in papers from
eld of bioinorganic chemistry [7,8]. Thomsons Web of Science 19902011 as a very rough estimate
DFT can describe molecular systems with an accuracy that of the overall trends in use of various theoretical methods (the
often competes with experimental data, e.g. for local geometries absolute numbers may be affected by search strategy). To the
of metal centers in proteins [9,10]. However, due to the formula- left, various general methods are compared: while semi-empirical
tion of DFT, there is no rigorous hierarchy of accuracy of functionals methods such as PM3 [49,50] are in decline due to DFT outcom-
corresponding to the many-electron space of conguration state peting them for widely studied medium-size systems (10100
functions in wave function theory (WFT), although DFT is varia- atoms), MP2, with its N5 size-restriction, maintains a steady use
tional for each functional separately. Thus, no universal functional of 1000/year due to its accuracy and stringent interpretation. DFT
is available in algorithm to accurately describe any given electron passed MP2 in the nineties and has grown super-linearly since then
density [1,3]. Scientists today rely on a relatively large arsenal of to reach 6000/year. Coupled-cluster methods (CCSD keyword)
functionals with various strengths and weaknesses, depending on display modestly increasing use, but a declining market share due
molecular systems and electronic properties studied, as reviewed to the massive increase in the use of DFT the last ten years.
before [6,1123]. The center and right parts of Fig. 1 show trends of a few
Functionals can be categorized according to their design and individual functionals (they were all searched as method syn-
mathematical complexity [14]. In terms of complexity, one can onym + dft, reducing total hits against dft alone to eliminate
distinguish (i) local uniform functionals (local spin density approx- bias from other acronym uses). B3LYP has for many years been
imations [24]) with the energy being a functional of the electron almost synonymous with DFT and remains by far the most applied
density [2527]; (ii) functionals that depend also on the gradient functional, although during the last 23 years, new functionals are
of the density (generalized gradient approximation, GGA) [2830]; quickly conquering market shares. PBE was a rst contender, which
(iii) functionals that also depend on more or higher derivatives, now seems to be overrun by the Minnesota functionals (M0x series)
e.g. the Laplacian of the density [31] or the kinetic energy density [51], but the trends are not completely established yet.
(meta functionals) [32]; (iv) functionals that substitute part of the The chemistry community currently faces a transition from an
exchange functional with non-local exchange from a determinant, almost uniform standard during the golden age of B3LYP to a new,
typically from a HartreeFock (HF) procedure (hybrid functionals) refreshingly vibrant, but highly fragmented and confused market
[33]; and (v) functionals that substitute also part of the correla- place of density functionals with numerous persuasive vendors. No
tion energy for WFT correlation, e.g. from MP2 (double hybrids) new golden standard will arise soon, as judging from Fig. 1, and
[34]. Further levels of complexity also occur, e.g. range-corrected this could damage the credibility of theoretical chemists to their
hybrids [35,36]. experimental colleagues and may divert productivity away from

Fig. 1. Papers published 19902011 containing keywords corresponding to various electronic-structure methods; from Thomson Web of Science. Left: total counts of general
methods: DFT, MP2, CCSD, and PM3. Center: DFT methods. Right: DFT methods without B3LYP.
198 K.P. Kepp / Coordination Chemistry Reviews 257 (2013) 196209

real problems of science. Thus, it is necessary to consciously eval- energy in some cases, e.g. in high-spin complexes with small ligand
uate new approaches, if not in the hope for a universal functional, eld splitting, compared to systems where electrons are paired in
then at least for a state-of-the-art for DFT, just as Pople dened a MOs subject to strong splitting.
model chemistry many years ago [52]. A third distinction is that of Coulomb and Fermi correlations,
On the path toward new standards, many critical assessments which are also critical to inorganic chemistry, as they emphasize
of DFT have been performed by comparison to experimental prob- the role of electron spin: Coulomb correlation is the part of the total
lems [6]. The situation was reviewed in 2007 [14], concluding that correlation that is not due to electron spin, whereas Fermi correla-
there is no consensus to deduce from these benchmarks. The all- tion arises from the Pauli repulsion between electrons, which are
dominating B3LYP in its various modications [53] performs only S = fermions. In HF theory, although often said to be uncorrelated
slightly better than average when viewing these benchmarks in (in Lwdins sense [57]), some Fermi correlation is described by the
total [14], and worse than average in a recent benchmark even for Slater determinant (Eq. (1)), with N electrons in N orbitals [58]:
main-group systems [54]. Still, the restricted nature of benchmarks,  
 1 (1) 1 (2) 1 (N) 
the number and types of functionals included, the systems stud-  
ied, the properties chosen, and the specics of calculation, render 1  2 (1) 2 (2) 2 (N) 

 = (1)
N!  
comparisons very hard. How does one normalize such studies to .. .. ..
. . . 
the same basis sets, correct for zero-point energies (ZPEs), thermal  (N) 
N (1) N (2) N
effects, dispersion, or relativistic effects, identify use of erroneous
electronic congurations, and remedy electronic bias of any spe- The energy lowering obtained in this way is called exchange,
cic choice of benchmark systems, e.g. iron complexes with nearly because the determinant exchanges (permutes) electrons among
half-full d-shells? all orbitals to ensure an anti-symmetric wave function as required
Whereas there are excellent reviews of DFT applications in inor- by the Pauli Principle [58]. Changing the number of unpaired elec-
ganic chemistry [6,1214,16,22,23], this mini-reviews aim is to trons in a process, as often observed in inorganic chemistry, e.g.
provide a detailed discussion of specic systematic effects often high-spin or odd-electron systems, is particularly challenging to
ignored in mainstream DFT modeling. Emphasis is on metalligand describe theoretically, as discussed below.
bonds, the core entity of coordination chemistry. As shown, neglect
of systematic effects may cause erroneous conclusions about chem-
2.2. Complex density functionals as a path toward accuracy
ical energies and the adequacy of functionals, and a good practice
of theoretical inorganic chemistry is suggested.
One signicant difference between WFT and the KohnSham
(KS) procedure [2] used in most DFT is that WFT keeps the
Hamiltonian H simple while expanding the one-electron and many-
2. General methodological considerations
electron spaces, as seen in Eq. (2), whereas DFT keeps to the
one-electron space while expanding the complexity of the Hamil-
2.1. Electron correlation
tonian (i.e. functional, fxc which can depend in various ways on the
electron density), and integrating only over space d3 r, as seen in Eq.
Electron correlation is the correlated movement of electrons
(3) [59].
beyond simple mean-eld behavior, and its realistic treatment is
required for computational accuracy, also in inorganic systems, 
N
where electron correlation is typically highly complex.  = ai i , 
E =  |H| (2)
In terms of spatial correlation of electrons, one can distinguish i
three components of the electron correlation: radial, angular, and 
leftright correlation [55]. Radial correlation reects correlated elec-
Exc [ ,  ] = d3 rfxc ( ,  ,  ,  , . . .) (3)
tron movement along the radial coordinate, i.e. toward and away
from the nuclei, and is allowed by using diffuse basis functions
with small exponents and thus, long tails. Angular correlation Whereas for WFT, the path toward universally accurate energies
is correspondingly achieved via higher-angular momentum basis is well-dened but slowly converging in many-electron space, in
functions. Leftright correlation is the third component of the elec- DFT, the path is less straight-forward but tends to converge quickly
tron correlation that reduces the probability of having two bonding due to ad hoc inclusion of the most relevant effects in the func-
electrons in the same part of the bond. This distinction is important tional. The difference arises because physical interactions between
when choosing proper basis functions. electrons (i.e. the electronelectron cusp) affect much more the
Another distinction of relevance to inorganic chemistry is energy than the precise electronic distribution beyond some accu-
between dynamic and static (non-dynamic) correlation, aris- racy. Thus, in DFT, the correlation energy converges much faster
ing from close-range electron encounters or from quantum- with the one-electron basis set than in WFT, because more of the
mechanical interference of energetically close-lying electronic important correlation is obtained directly from the functional: the
congurations, respectively. Substantial changes in static corre- He-atom is a good example, where s-functions are enough to obtain
lation often occur upon bond cleavage and in transition metal the correct correlation energy with the exact density functional
systems with close-lying d- or f-orbitals [56]. [60].
Consider a metalligand bond that is broken homolytically, e.g. It is thus more apparent why current DFT research focuses on
the organometallic Co C bond of cobalamins: leftright correlation expanding the complexity of functionals: they progress from the
may weaken the bond as it destabilizes the bond by increasing the local density functionals that only depend on the electron densi-
probability of having electrons on opposite sides of the molecule, ties of spin-up ( ) and spin-down ( ) electrons (Eq. (4)), toward
e.g. by moving electron density from a metalligand -MO to a *- GGA functionals that depend on the gradients (  ,  ) (Eq. (5)),
MO. Also, static correlation affects the metalligand bond cleavage or meta functionals that depend on higher terms, e.g. the kinetic
as it is often larger in the dissociated states, or on the path toward energy densities  and  (Eq. (6)) [61].
dissociation, where electronic congurations have more similar 
LOCAL
energies. Furthermore, close-lying electronic congurations with Exc [ ,  ] = d3 r(r)fxc
LOCAL
( ,  ) (4)
variable d-orbital occupations will display more static correlation
K.P. Kepp / Coordination Chemistry Reviews 257 (2013) 196209 199


2.4. Identifying the congurations that are lowest in energy
GGA
Exc [ ,  ] = d3 r(r)fxc
GGA
( ,  ,  ,  ) (5)
However, even before that, one must correctly identify the elec-
 tronic conguration of lowest energy mimicking the ground state
meta
Exc [ ,  ] = d3 r(r)fxc
meta
( ,  ,  ,  ,  ,  , . . .) (6) density. Most standard DFT approaches rest on the KS approach
where the KS-determinant is an effective approximation to the
true (mixed) state. Therefore, KS molecular orbitals (MOs) are not
As discussed in this review, while expanding the complexity of strictly representative of congurations as in WFT, but more of
these functionals further, one should assess the impact of other approximate states. Furthermore, as in UHF, spin contamination
effects usually added as corrections to the calculations, since they will be observed in most non-closed-shell systems, with S not
often affect accuracy more than incorporation and parameteriza- being a good quantum number in unrestricted calculations where
tion of new terms in functionals. Experimental structural, energetic, the Fermi correlation polarizes the - and -spaces differently
and spectroscopic data are critical for validating the accuracy of [71]. Both hybrids and non-hybrids can display spin contamina-
functionals for given tasks, and should be practiced whenever tion, which will tend to be larger in hybrids due to the HF exchange
experimental data of any kind are available; if such data are unavail- that enhances spin polarization.
able, correlated ab initio methods may provide insight into the Despite this, any conguration of the S, MS quantum numbers
accuracy of the energies and properties obtained by a given func- is approximately accessible by local convergence or spin-down-
tional. coupling: one may obtain both closed-shell singlets (S, MS = 0, 0)
that are pure, i.e. similar to the true single-congurational state,
and impure open-shell singlets (S, MS 1, 0) that are only one of
2.3. Energies and basis sets the several congurations of the true multi-congurational state.
The latter is usually obtained from the input MOs of the pure triplet
As in WFT, molecular geometries are less sensitive to the method conguration (S, MS = 1, 1) by changing the spin of the highest singly
and can be modeled quite accurately by most GGA and hybrid func- occupied molecular -electron in the calculation input. To honor
tionals in wide use, with mean absolute errors (MAE) of 0.02 A the KS effective determinant principle, it is absolutely necessary
for bonds and 12 for angles [1315,6265]. However, geometries to obtain the correct (i.e. lowest in energy) of these congurations.
of systems with very weak bonds or signicant dispersion interac- It is quite common to see excited congurations resembling lower
tions [66] may be subject to larger functional-dependent errors and spin polarization and higher energy: a typical example is O2 binding
require special attention. Classical bioinorganic examples are the to heme, where the closed-shell conguration is excited, whereas
soft Co Nax bond in cobalamin enzymes [67] or the soft Cu S bonds the correct conguration is spin-polarized reecting some Fe(III)
in blue copper proteins [68], whereas strong -donating ligands character [72].
usually exhibit smaller geometric errors but equally large effects In clusters with coupling between several metal sites, the
on energy [67]. maximally spin-polarized congurations are often referred to as
When modeling accurately energies of bonds and transition broken symmetry congurations that best resemble the true
metal systems, to allow for the necessary electronic freedom, triple- anti-ferromagnetic state [73]. A conguration with spin densities
zeta basis sets with polarization and diffuse functions are required close to 1 and 1 in a coupled iron(III) cluster is thus a single-
[52,69]. While moving to quadruple-zeta is important in post-HF ab conguration-approximation to an excited multi-congurational
initio WFT, it has little effect on DFT-computed bond dissociation state (namely, an impure S, MS = 1, 0 state). If the congurations
enthalpies (BDE) of most metalligand bonds [54]. However, dif- of lowest energies are not obtained by careful procedures both in
fuse functions can be important due to the differential correlation benchmarks and in routine studies, the states and processes inves-
effects in the compact (bound) state not present in the dissociated tigated will not be accurately described. This is true particularly in
state [62], and use of unbalanced basis sets is error-prone as elec- systems with large static correlation such as transition metal sys-
trons favor to reside in the part of space covered mostly by basis tems and potential energy surfaces. Careful, manual sampling of
functions. the electronic congurations is necessary.
Polarization functions on hydrogens are required if the electron The problem of static correlation also haunts the use of coupled-
density changes around hydrogen atoms, as e.g. in hydrogen-bond- cluster methods or indeed, any post-HF ab initio method based
breaking or H-abstraction reactions, and more diffuse functions are on a single HF reference determinant [74]: coupled-cluster with-
required if electron-rich or excited electronic congurations are out explicit triples excitation operator (i.e. CCSD(T) or less) from
involved. Electron afnities, excitation energies, and some softly such a reference is unreliable in systems with static correlation
bound systems should routinely be studied with diffuse functions [75,76], e.g. transition metal systems and bond breaking reactions
that enhance the radial correlation of the electrons to allow unbi- [77,78]. Time-consuming CCSDT can recover signicantly more
ased reproduction of both states of the real process. When this is of the static correlation, as can active-space CCSDT or hybrids
done and other effects such as solvation are taken into account, rel- thereof [79]. Sometimes, using a DFT or CASSCF/CASPT2 [80] refer-
ative reduction potentials of electronically similar systems can be ence may drastically improve otherwise erroneous CCSD(T) results
obtained with almost chemical accuracy, as evidenced by studies [81]. CASSCF/CASPT2 wave functions are not biased toward a
of rst-coordination sphere single-mutants of ironsulfur proteins single HF reference but are fully optimized within their active
[70]. space and are thus important tools for statically correlated tran-
Thus, a state-of-the-art basis set is necessary but not suf- sition metal systems and for properties such as excitation energies
cient to provide electrons with the freedom to distribute optimally [82]. The known one-determinant bias of single-reference post-
in space in the involved states of the process: the second criti- HF methods has not prevented some researchers from blindly
cal requirement is how the functional describes short-, medium-, using UHF- or even ROHF-based CCSD(T) energies with strong
and long-range Fermi- and Coulomb correlation and thus effec- bias toward some HF conguration for systems with substan-
tuates this distribution, and the third critical requirement is tial static correlation. The failure of the single-determinant HF
proper account of corrections for other physical effects not directly reference for describing pseudo-degenerate states is also the rea-
included in the functional (entropy, ZPE, relativistic effects, disper- son why perturbation theory (e.g. MPn) is not valid in these
sion, etc.), which is the main focus of this review. situations.
200 K.P. Kepp / Coordination Chemistry Reviews 257 (2013) 196209

Still, ab initio methods such as CASPT2 and coupled-cluster 3.1. HF exchange


methods can offer critically important insight into statically cor-
related inorganic systems and constitute the only validation of It was early thought that HF exchange from the Slater determi-
density functionals when experimental data are poor or absent [83]. nant (Eq. (1)) might improve DFT [11], and the hybrid functional
Much work has been done using CASPT2 in inorganic chemistry B3LYP showed that a fraction of HF exchange can improve descrip-
[84], including problems with substantial static correlation such as tions of many properties [33]. Due to the permutation of all
charge-transfer states [85,86] and biologically relevant tetrapyrrole indistinguishable (i.e. same-spin) electrons caused by the deter-
systems [8791]. Also, substantial work has been done with e.g. minant, HF exchange has an important ability to mimic the real
equation-of-motion coupled-cluster methods, e.g. Fe-, Co-, and Ni- long-range (non-local) component of electron correlation that is
carbenes with nearly degenerate ground states [92,93], considering very hard to mimic by other simple means. Some hybrid func-
various reference wave functions to assess one-conguration bias. tionals, notably B3LYP with 20% HF exchange, perform well for
Both types of studies may accurately account for the lowest-lying structures and many types of chemical reactions involving s- and
electronic states in statically correlated inorganic systems and thus p-block elements [1,5]. This success has made B3LYP the most used
provide important data in the future benchmarking of inorganic functional in chemistry, including inorganic chemistry, as men-
DFT energies. tioned Section 1 [14,21,22].
HF provides a decent description of some atomic and ionic con-
gurations where errors tend to cancel, although the absence of
3. Systematic effects in electronic structure calculations dynamic correlation renders it very inaccurate for electron-rich
systems, as observed e.g. in the increasing errors in 3dn1 4s1 3dn
In addition to basis set, choice of functional, model system, and excitation energies going from Sc+ to Cu+ [11]. HF dramatically
efcient conguration optimization, several other aspects need to underbinds in many cases [95,96]. In the d- and f-blocks, electronic
be addressed. Table 1 outlines these additional effects that should congurations are closer in energy than the s- and p-blocks and
ideally be addressed in DFT model chemistry. As will be discussed, thus enhance static correlation, which is also absent in HF [97].
many of these effects are not always appreciated although they The optimal amount of HF exchange depends on the electron cou-
are often systematic and predictable and can be generalized to pling via the adiabatic connection formula [98]: near-degenerate
favor either the state with the more compact or the looser electron ground states are typically more accurately described by less than
density. 20% HF exchange [42], because a single HF Slater determinant is
On one hand, there are real, physical effects often with system- then biased. Non-hybrid exchange functionals are not biased by the
atic behavior such as dispersion interactions, relativistic effects, HF exchange of a single, in such cases misleading, conguration,
entropy, ZPE, and solvent effects, which must be included as a and thus tend to better describe systems with substantial static
correction to the electronic energy. Of these, dispersion and rela- correlation often found in the d-block [62].
tivistic effects are electronic in nature, whereas entropy and ZPE are When the electronic structure changes qualitatively, Fermi cor-
mainly related to the vibrational states of the system, although the relation constitutes the majority of the differential correlation
electronic congurations also affect them via the electronic state energy, as witnessed from the impact of HF exchange on BDEs [99],
function and the vibrational coupling. Thus, their effect on observ- and 20% HF exchange is excessive [54,99]. As observed already by
ables such as bond strengths is indirect (and their effects marked by Reiher [100] and Paulsen et al. [101], relative multiplet energies
in Table 1), as they are added after convergence of the electronic dependent almost linearly on the amount of HF exchange, which
structure, but are generally systematic as the electronic effects. Sol- thus favors high-spin (HS) over low-spin (LS) [102,103]. This also
vent effects deal with the systems surroundings. On the other hand, leads to weaker bonds with lower BDEs of 30 kJ/mol per 10%
there are some features of the functionals themselves, notably HF of HF exchange for metalligand bonds of diatomics in vacuum,
exchange and self interaction, which are also often of a systematic where dispersion, solvent effects, and entropy errors are negligible
nature. [62,99].

Table 1
Some systematic effects in density functional theory.

Systematic effect Typical effect on  Typical effect on bonds Examples Representative references

Ingredients in functionals
HF exchange Looser Longer/weaker HF and hybrid functionals [11,62,100,104,127,160]
underbind and favor HS, up to
30 kJ/mol per 10% HF exchange
Self-interaction error (SIE) Looser Weaker in Articial low dissociation limits of [108,113115]
odd-electron systems X2 + ; favors delocalization of
SOMOs, errors can be >50 kJ/mol
Physical effects
Dispersion More compact Shorter/stronger Dispersion increases BDE, favors [126,128]
LS, often >10 kJ/mol
Relativistic effects s, p-contraction, valence Stronger, compact Favor stronger -bonds; but favor [62,136,139]
expansion -bonds; weaker atomic states in middle d-block
bonds in middle d- and 40 kJ/mol in some cases
p-blocks
Solvent effects Depends on polarity and Signicant in charged, E.g. 0.020.03 A shorter M L [94,144,145,147,151]
charge of molecule polar, weakly bonded, bonds, favors LS. Polar reactions
and unsaturated are highly sensitive to solvent
systems (often 100 kJ/mol or more)
Zero-point energies (ZPEs) Indirectly loosening Longer/weaker ZPE favors HS, typically by [62,99,160]
515 kJ/mol
Entropy Indirectly loosening Longer/weaker Entropy lowers BDE and favors HS, [126,163]
typically by 1030 kJ/mol per
metal site per bond/metal
K.P. Kepp / Coordination Chemistry Reviews 257 (2013) 196209 201

Down-toning the HF exchange component of B3LYP improves DFT-D [120,121], dDsC [122] or the non-empirical corrections by
accuracy in d-block chemistry, rst done via the B3LYP* functional Tkatchenko and Schefer [123]. These methods develop rapidly: as
(15% HF exchange) [104,105] and later via a BP86 hybrid with an example, the newest of the DFT-D corrections, DFT-D3 [124],
10% HF exchange giving better agreement with spectroscopic data differs substantially from the previous DFT-D1 and -D2 methods,
[106]. Newer hybrid functionals display various HF exchange com- in particular by using new physical cutoff radii for the dispersion
ponents in the range from 10% (TPSSh) and up to approximately interaction and specic atom-pair coefcients, rendering it less
50% (B2PLYP, M06-2X). The impact of HF exchange on results for empirical [124,125]. DFT-D2 improved DFT-D1, e.g. by providing
late transition metal systems has also been discussed in detail in a more realistic geometric-mean combination rules for estimat-
review that also encouraged the use of meta GGA functionals [107]. ing the dispersion interaction from individual atomic coefcients
[121].
3.2. Self-interaction and non-locality Dispersion in a purely attractive form favors closer interaction
between electrons in different MOs and produces more compact
The HF exchange, so far necessary although problematic in DFT, electron densities, stronger binding, and shorter bonds. LS states
is accompanied by other effects that are often systematic, notably are favored by shorter metalligand bonds (see Section 5), and the
the self-interaction energy (SIE) of electrons [3,108,109]. Whereas stabilization of LS by such dispersion corrections can be large, as
this energy is strictly zero in WFT, where K and J integrals cancel realized in recent work on axially substituted iron(III)porphines
completely, in DFT, due to the above-mentioned separate treatment [126]. Dispersion also increases BDEs of bulky molecules such as the
of exchange and correlation functionals, it is not zero except by Co C of cobalamins, and the good performance of GGA functionals
special design. Constraining the SIE to zero damages the accuracy for this bond [127] may be due to neglect of dispersion [128]. How-
of current functionals dramatically [110,111], showing that the SIE ever, the D3 correction is substantially smaller than D2 in these
is coupled to other effects, as will be discussed below. cases, and the models used for AdoCbl neglected many strongly
SIE increases with the distance between orbital centers and is interacting side chains of the corrin ring, suggesting a cancelation
thus important in processes that separate single electrons over of two errors in the study, although the general conclusion remains
delocalized MOs, e.g. dissociation of odd-electron systems [112], valid. Dispersion is also necessary to describe relative energies of
near-degenerate ground states as seen in the d-block, some open- isomers and conformations with different steric crowding, as seen
shell transition states, or radicals in delocalized -systems. in e.g. alkanes [129].
SIE is often illustrated by dissociation of X2 + : at shorter dis- A natural step forward is to include dispersion corrections
tances, the - and *-MOs are separated and the electron resides directly in benchmark studies once these are sufciently trusted
in the -MO. SIE corresponds to a localized electron due to the [54,130], which, as other systematic effects, will change the overall
signicant energy gap to *, thus always carrying its SIE with it. conclusions from many previous benchmarks [14]. Overall perfor-
However, as X2 + dissociates, the unpaired electron is delocalized by mance is markedly better than for old benchmarks of DFT without
static correlation. In WFT, this is reected by multi-congurational dispersion [131], conrming the utility of such methods. They
character, whereas in DFT, it is reected by polarization of the should therefore be routinely applied as good practice, to test its
singly occupied molecular orbital (SOMO) in response to the func- impact on computed results.
tionals attempt to mimic the static correlation. Upon dissociation,
the splitting is reduced and the ground state becomes a quantum- 3.4. Relativistic effects
superposition of two ionic states with energy E(X+ ) + E(X). The exact
energy is a linear function of any fractional electronic occupa- Relativistic effects generally stabilize and contract the most rel-
tion of the MOs. This situation is correctly described by HF/WFT, ativistic core s- and p-electrons, and in response all outer s-shells
which is SIE-free, due to 100% HF exchange. However in standard due to their overlap and reduced repulsion [132]. This often leads to
DFT, this cancelation is incomplete, and the energy curves down- strengthening of late d-block metalligand -bonds, e.g. Au L vs.
wards between integer electron systems, due to the lowering of Ag L [133]. Importantly, these systematic relativistic contractions
SIE obtained from partial electron delocalization. The problem is are often >0.02 A already for Cu L (Zn L are less contracted) [132]
always the differential SIE of the process, which in this case equals and are important already for the electronic properties of transition
the difference between the symmetric (half-integer electron) and metal systems from the rst row of the d-block [81,134].
the ionic (full electron) dissociation limits and can be massive Because of the resulting, enhanced shielding of the nuclei, the
[113,114]. diffuse, higher angular momentum MOs (d- and f-type orbitals
Normally, functionals will enhance resonance and delocaliza- and anti-bonding MOs in the valence shell) are typically expanded
tion of electrons between close-lying congurations as a means and destabilized [135], leading to polarization of electrons and
to lower the SIE. Therefore, SIE is sometimes said to mimic static weakened bonds, especially of higher bond order. Furthermore,
correlation [115,116]. Thus, SIE is, as spin contamination, an arti- spinorbit coupling can contribute substantially to weakening
cial by-product of the functionals attempt to optimally correlate of bonds by favoring the dissociated states that often exhibit
the electrons. In coordination complexes, the error will be smaller higher spin and angular quantum numbers [138]. The effects are
because the MOs made from the d-orbitals are usually localized, but comparable in magnitude and often roughly proportional to Z2
cases have been reported where differential SIE signicantly affects [132].
exchange coupling constants [117] and reaction energies [118,119] Sometimes, effective core potentials can account for most of
of transition metal systems. the relativistic effects [81]. They confer a reasonably accurate con-
tracted atomic core and reduced effective nuclear charge, thus
3.3. Dispersion mimicking the above-mentioned effects. For isodesmic reactions
where the quantum numbers are qualitatively unchanged, the
Dispersion, a second-order perturbation of instantaneously effective core potential often sufces. However, in processes where
induced dipoles in closely interacting electron densities, is gen- valence electronic structures change qualitatively, differential rel-
erally not accounted for in standard functionals, even meta ativistic effects are important, both due to corevalence correlation
functionals, although it could possibly be accounted for in higher- effects not imitated by a rigid core and due to intrinsic effects of the
order meta functionals as local density gradient products. The valence shell [81]. To remedy this, one can compute the scalar rela-
absence of dispersion can be remedied e.g. by corrections such as tivistic effects directly on all explicit electrons and include them in
202 K.P. Kepp / Coordination Chemistry Reviews 257 (2013) 196209

Fig. 2. Scalar-relativistic effects in 80 rst-row d-block M L diatomic bond enthalpies (gure based on data from Ref. [62]). Notice the behavior of Cr and Cu due to half-
and full d-shell congurations 3d5 4s1 and 3d10 4s1 sacricing the 4s2 inert pair stabilization for lower electron repulsion.

benchmarks of rst-row [62,99] and especially second- and third- 525 kJ/mol toward methods that compensate this, e.g. hybrids
row elements [136]. with large percentages (>15%) of HF exchange.
In the rst-row of the d-block, scalar-relativistic effects aris-
ing especially from changes in electronic structure (i.e. quantum 3.5. Solvent effects
numbers) contribute signicantly to reaction energies. As seen in
Fig. 2, on average, scalar-relativistic effects weaken M L bonds Solvent effects play a key role in two forms: via uniform
by 7 kJ/mol with large variations [62], and the effects are largest electrostatic screening in the condensed phase, obtainable from
in the middle of the d-block, where more unpaired electrons, i.e. macroscopic continuum models, and via heterogeneous screening
highest spin, are generated in the atoms. Another cause of the rel- and explicit solvation, obtainable from microscopic solvent struc-
ativistic destabilization of the bonds is the 4s2 inert pair of the ture.
atomic states. This pair is absent in the half- and full-d-shell Cr The effect of electrostatic screening is largest in molecules with
and Cu (3d5 4s1 and 3d10 4s1 ) where normal shell-wide lowering of very heterogeneous charge distributions, non-zero total charges,
electron repulsion dominates over the relativistic 4s stabilization, soft bonds or loose conformations exposed to solvent, and/or
causing overall M L -bond stabilization as commonly seen for explicit hydrogen bonding. In such cases, geometries can also
heavy metal halides [137]. These effects (e.g. +13 kJ/mol for Cu H) be affected, and hydrogen-bonded systems have a dispersion-
are comparable to full relativistic considerations (+8 kJ/mol) [138]. component that also needs to be addressed [143]. For ironsulfur
However, in all diatomics except Cu L and Cr L, scalar relativistic [144] and metallothionein [145] clusters, which are 2 or even 3
effects favor the dissociated congurations. charged, optimizing geometries in a dielectric continuum model
These non-negligible relativistic effects on M L bonds in the such as Cosmo [146] leads to a substantial (0.020.03 A) shortening
rst row of the d-block are consistent with second-row p-block of the M S bonds by screening the repulsion between negatively
systems, e.g. 20 kJ/mol scalar-relativistic destabilization already in charged sulfur atoms. This profoundly inuences the electronic
SF6 , increasing with higher substitution (corresponding to coor- structure of these biological metal clusters. The rst 5 units of
dinative saturation in the d-block) [139]. Both scalar-relativistic from homogeneous screening captures most of the effect because
and spinorbit effects, which favor higher spin and angular of the inverse relation of on the Coulomb interaction (Eq. (7)),
momentum in dissociated states, systematically favor bond weak- rendering discussions of = 10 or 20 more academic [145].
ening. Bauschlicher showed this in the case of scalar-relativistic
Q1 Q2
weakening of Ga Cln bonds, increasing quickly with n and reach- E= (7)
4r12
ing 25 kJ/mol of bond weakening with n = 3 [140]. Martin and
co-workers identied systematic bond weakening from scalar- Heterogeneous solvent effects and the accuracy of cluster mod-
relativistic effects in the order of 15 kJ/mol for the rst- and els have been discussed in detail by Siegbahn and Himo [147,148]
second-row of the p-block, but increasing rapidly with size [141]. and by Ryde and co-workers [149]. As has been observed by
A discussion of the reduction in p-block bond energies by performing full DFT optimization of small proteins [150], the
relativistic effects has been presented by Wang et al. [142]: interest- heterogeneous and homogeneous effects on geometry can be
ingly, this destabilization of bonds is not simply correlated to bond quite similar, i.e. a large part of explicit solvation is obtainable
elongation as in non-relativistic situations. Thus, bond contraction from the homogeneous component in dielectric continuum mod-
will often be observed, as in many -bonds described in Pyykkos els. However, this is not always true: if the system is subject
review [132], while destabilization of the bond energy occurs [142]. to specic, non-uniform interactions from hydrogen-bonding or
This is also mentioned by Pyykko, giving the example of AuCl [132]. metal-coordinating solvents such as water, acetonitrile, or tetrahy-
The consequence of scalar-relativistic bond weakening is that drofuran, explicit solvation can affect structure and energies.
studies that neglect scalar relativistic corrections in non-isodesmic Energies of polar reactions often exhibit solvation effects of the
reactions (enthalpies of formation, bond breaking, spin crossover, order of 100 kJ/mol, since the hydration energy of monovalent
etc.) of third- and fourth period can be biased by typically ions is typically several hundred kJ/mol [151]. This is e.g. seen
K.P. Kepp / Coordination Chemistry Reviews 257 (2013) 196209 203

Table 2
Effects of homogeneous solvation, Cosmo radii of Fe, dispersion, and explicit THF solvation on the stability of high-spin vs. intermediate-spin in iron(II)porphine (in kJ/mol).

Energy gaps (kJ/mol) Vacuum Cosmo, = 25, rFe = 2.0 A Cosmo, = 25, rFe = 1.5 A THF2

EHSIS (BP86 no dispersion) 70.3 76.3 66.9 41.9


EHSIS (BP86-D3) 70.8 76.6 67.4 41.3
EHSIS (B3LYP-D3) 23.0 32.7 23.2

in polar bond cleavage reactions [152], in neutral-molecule-ion at all. For example, in cobalamin coenzymes, the effect of ZPE on
substitutions [153], and in SN 2 reactions with charged reactants the Co C bond was 20 kJ/mol, however canceling with several
or products such as in the cobalt-containing methionine synthase other effects, leaving the BP86 result close to the real Co C BDE
[154], to mention a few examples. [127]. This is probably due to cancelation of errors from absence
As an illustrative example, coordinatively unsaturated planar of both dispersion and HF exchange with SIE, as suggested from
porphyrins are affected by explicit solvation not captured by dielec- Table 1 and observed partly by adding dispersion to B3LYP [128].
tric models: from Table 2, THF coordination to iron(II)porphine Most researchers include ZPE as state-of-the-art, although it is still
moves HS 2535 kJ/mol closer to intermediate spin (IS), although sometimes ignored, causing the systematic errors mentioned.
the HS state is still an excited state with this method. The effect
originates from oxygen lone-pairs in THF that destabilize IS, which
has one more electron in solvent-exposed occupied d-orbitals with
3.7. Entropy
a z-component (either dz 2 or dxz /dyz ) compared to HS.
In this case, dispersion does not change the relative energy
Observed ground states in condensed phase at thermal equi-
because the porphine is rigid, but it would if axial ligands were
librium are the result of free energies, not electronic energies or
present, via the differential ring-ligand dispersion interactions in
enthalpies. Therefore, any conclusions regarding the chemical equi-
longer-bond HS and shorter-bond LS [126]. Notice also the sig-
librium or the thermally averaged ground state need to take
nicant effect of changing the Cosmo radius of iron, allowing the
entropy into account [126].
solvation probe to perturb the two states differently. The experi-
In chemical reactions where bonds break, which include the
mental ground state is a triplet state [155,156], but the energy to
majority of interesting catalytic processes, entropy is highly impor-
the quintet state is not known, although CASPT2 computes them
tant due to the additional translational and rotational degrees of
to be very close in energy [157]. A microscopic solvation effects of
freedom. For example, when identifying partially ligand-bound
25 kJ/mol could change conclusions, as is also seen experimentally
intermediates in (bioinorganic) catalysis [161], entropy is critical,
when changing solvent from e.g. CH2 Cl2 to THF [158], and coordi-
as evidenced e.g. in the identication of catalytic intermediates of
natively unsaturated systems in condensed phase thus need critical
di-iron enzymes with differing O2 -binding geometries [162]. The
validation.
differential entropy increases more or less linearly with the num-
ber of metal centers in clusters, up to 60 kJ/mol for isomers of
3.6. Zero-point energies di-iron enzymes [162].
Neglect of entropy is particularly grave because
Zero-point energy, the quantum-mechanical vibrational energy entropyenthalpy compensation, as in many other aspects of
at the bottom of the potential energy curve, does not affect the physics and chemistry, is a major force in molecular interac-
computed electron density or bond directly, because it is added as tions: a strong bond favored by enthalpy will usually display
a correction after the convergence of the electronic structure, as unfavorable entropy because of the smaller contribution of more
is entropy. Still, ZPE can affect calculated heats of formation and widely separated vibrational energy levels to the state func-
bond strengths signicantly to the same nal effect as a loosen- tion, whereas the vibrational levels are accessible in the weaker
ing of the density or bond. Thus, ZPE and entropy are marked with bond. This entropyenthalpy compensation is signicant even
in Table 1, as their effect on observables is after nuclear correc- in reactions where bonds do not break, as is well illustrated by
tions. Thus, their neglect can cause error cancelation with effects spin-crossover complexes (vide infra) for which the spin transition
that directly inuence the electron density. Reasonably, functional- is an equilibrium with temperature usually favoring HS [163].
dependent scale factors should be used in such studies [159]. Under typical conditions, entropy contributes 1030 kJ/mol to
ZPE tends to favor the HS states that generally have weaker M L the stability of HS in mononuclear coordination complexes [163].
bonds and therefore smaller ZPEs than the LS states. Neglect of The reason for this entropy effect in spin-crossover is that the
ZPE in spin crossover, even when entropies are absent, is therefore occupation of d-orbitals pointing toward the ligands (the eg set
a source of systematic errors, typically 515 kJ/mol bias against in Oh symmetry) weakens metalligand bonds in the HS state and
the HS state [105,160]. With a GGA functional that itself is biased renders its entropy larger. Many studies neglect this large compen-
toward compact electron densities (LS, strong bonds), neglect of sation effect when hypothesizing on the nature of ground states.
ZPE will make errors even larger. For hybrid functionals such as Neglecting entropy will typically, because its neglect indirectly
B3LYP, which intrinsically favor HS, neglecting ZPE will cause two favors electronic structure methods giving compensatory, looser
systematic errors to partly cancel, even more so if entropy is also electron densities, tend to cancel with errors in hybrids having too
neglected (vide infra). much HF exchange [126].
When computing reaction energies, dispersion effects that favor Calculations of entropy contributions are associated with sub-
the compact electronic structure (LS or bound state) work oppo- stantial uncertainties, in particular when low-energy modes are
site to ZPE corrections and HF exchange, and neglect of dispersion involved, which are affected by the environment of the molecule
in larger complexes, in particular with bulky groups, will lead to [100]. Still, entropies within 10 kJ/mol of experimental data are
an articial bias toward looser states, something that has been attainable for mononuclear complexes using standard harmonic
common until the advent of dispersion-corrected DFT methods vibrations and subsequent thermodynamic analysis [105]. Thus,
[120,121]. the entropy effect is signicant and computable, although the
In the calculation of M L BDEs, ZPE is even more important as uncertainty should be acknowledged, and its neglect will lead
the dissociated state does not have ZPE from the dissociated bond to wrong conclusions of the performance of electronic structure
204 K.P. Kepp / Coordination Chemistry Reviews 257 (2013) 196209

methods if the property used for benchmarking is related to the saturated complexes, hybrids such as PBE0, B3LYP, or TPSSh give
free energy, not energy or enthalpy [164]. slightly more accurate M L bond lengths than GGA functionals
across the three rows of the d-block, although these differences
are not critical to the reaction energies [6264,165].
4. Insight gained from benchmarking of metalligand As other explicit correlation effects, HF exchange favors delo-
bonds calization of electrons into MOs of higher angular, spin, and radial
quantum numbers, including anti-bonding MOs that lowers bond
4.1. Correlation effects in metalligand bonds orders. This polarization is witnessed in M L dipole moments,
which are substantially (typically 0.5 a.u.) and systematically
The M L diatomics (M = metal of the d-block, L = s-block larger when computed with B3LYP than with BP86 without HF
or p-block atom) have been the focus of recent benchmarks exchange (see Fig. 3) [62]. The only exceptions (TiC, TiN and ZnO)
[62,99,119,136] for several reasons, other than being the funda- could be due to differences in BP86 and B3LYP electronic struc-
mental entity of many inorganic chemical reactions: (i) there are tures. The increased dipole moments caused by HF exchange is
a large number of experimental bond enthalpies available; (ii) another manifestation of the polarization of the electron density
these data are from vacuum, eliminating random errors associ- also responsible for weakening of bonds. Given the importance
ated with use of solvent models; (iii) entropies are not required of HF exchange in most properties, testing the variation of pos-
to be estimated, eliminating a further cause of error, including a sible outcomes using two functionals such as B3LYP and BP86 (or
large random error component of at least 5 kJ/mol; (iv) disper- another non-hybrid GGA functional) is a reasonable way to rou-
sion effects in diatomics in vacuum are of the order of 1 kJ/mol tinely perform a reality check of results, which should in this
even with larger metals, reducing this source of error; and (v) the authors opinion always be practiced.
large number of data for the entire d-block eliminates commonly Electronic polarization also increases spin contamination in
observed bias toward certain electronic systems, as required in unrestricted computations (except in pure congurations of
universal benchmarking. None of the other commonly used bench- MS = S) and is sometimes called mixing-in of higher multiplet
mark data sets fulll all these criteria, and therefore, the M L determinants although it is the composition of the MOs that
diatomic bonds are an important part of the benchmark tool box in change. It illustrates the KS effective conguration principle in
this authors opinion. Elimination of bias to electronic structures is action, and spin contamination is thus a natural by-product of this
incidentally also a motivation of the articial molecule data sets optimization under the inuence of the functional [166].
by Grimme and co-workers [188].
An all-purpose functional claimed to work well across chemistry
should balance long- and short-range Fermi- and Coulomb corre- 4.2. Bond strengths of M L diatomics
lation for both loose (atomic) and compact (molecular) states. In
Section 3, it was shown how various effects that are often partially For the d-block, BDEs for M L with L = H, F, Cl, Br, I, O, N, C, S,
ignored affect energies. In the quest for a state-of-the-art of inor- and Se are in many cases known experimentally in vacuum and
ganic DFT modeling, M L BDEs are critically challenging because constitute a benchmark database with the advantages described
they depend on two very different limits of interaction: bond break- above. Furthermore, BDEs are useful proxies for reaction barriers
ing leads to redistribution of electrons in new, more atom-centered of M L bond-breaking reactions in coordination chemistry.
MOs (AOs) under substantial changes in correlation energy. Mov- Common non-hybrid GGA functionals or hybrids with HF
ing from atoms to diatomics to saturated systems, orbitals change exchange of 20% or more (B3LYP, PBE0) exhibit mean errors in BDEs
from being more atomic to become more molecular, residing in of 4050 kJ/mol. These errors are largest for M L bonds with L = O,
pairs in the bonding regions. Functionals may apparently perform S, N, C, i.e. bonds with higher bond order and more complex elec-
well for coordinatively saturated systems only because of cancela- tronic structures. Early d-block M L bonds (Sc, Ti and V) exhibit
tion of errors, e.g. dispersion, thermal or relativistic effects that are larger errors in BDEs [62]. However, mean errors are largely due
more prominent in larger systems. to systematic errors of over-binding in non-hybrid GGA function-
For the rst row of the d-block, both GGA functionals and als and under-binding in B3LYP and PBE0 [62]. This suggested that
hybrids are known to give accurate bond lengths, typically systematic errors may be reduced by reducing HF exchange, as was
with uncertainties of 0.02 A [62,63]. For larger, coordinatively previously observed in spin splitting [100].

Fig. 3. Effect of functional on metalligand bond polarity: differences in Mulliken dipole moment (a.u.) of diatomics computed with B3LYP and BP86 (positive means that
the B3LYP dipole moment is larger). The main reason for the systematic difference is HF exchange (see text).
Figure produced using data from Ref. [62].
K.P. Kepp / Coordination Chemistry Reviews 257 (2013) 196209 205

where systematic errors will cancel to an extent that goes well


beyond the cancelation achieved with 10% HF exchange [168].
Such energies were used to conclude that tetrapyrrole systems
have co-evolved with their respective, native metal ions (i.e. LS/HS
Fe(II/III) with porphyrin vs. LS Co(I/II/III) with corrin), based on
relative thermodynamics of metal exchange reactions that quan-
tify how tetrapyrrole cavity sizes and electrostatics are optimal
for each metal ion and multiplicity [169]. Using specially designed
isodesmic reference reactions such as Eq. (8) greatly increases accu-
racy by cancelation of otherwise obscure errors, and should, even
when practicing the inclusion of all effects in Table 1, be applied
when possible.

5. Metalligand bonds and spin crossover

5.1. General aspects of spin crossover

Fig. 4. DFT-computed bond dissociation energies plotted against experimental data. When bonds break in inorganic chemistry, it is common that
Figure reproduced with permission from Ref. [99]. Copyright 2008: The American the overall spin changes. Together with bond dissociation of M L
Chemical Society. bonds, spin crossover (SCO) is a challenging process to describe,
due to substantial changes in electron correlation, in particular
The meta hybrid TPSSh was developed from the non-empirical Fermi correlation. SCO is an important process found in widely
TPSS, designed to fulll fundamental constraints, with 10% HF different areas of chemistry [170,171]. It occurs during many
exchange [41,43]. Thus, it was a plausible candidate for balancing M L bond dissociations, e.g. ligand binding to metalloproteins
the systematic errors mentioned above. TPSSh improves perfor- [72,172], and is of great interest in molecular magnetism, catal-
mance over GGA functionals for rst-row transition metal systems ysis, and molecular storage [173176]. Thus, there have been
[119]. It removes most of the systematic error of BDEs, as evi- many attempts to provide accurate theoretical predictions of SCO
dent from a linear regression of experimental vs. computed BDEs [101103,160,175,177181].
(Fig. 4) [99]. For the second- and third-row d-block, the coefcient SCO occurs at thermal equilibrium and is subject to a very
is 0.99 [136], i.e. the systematic error component is eliminated high degree of entropyenthalpy compensation: typical values
in the whole d-block, not obtained by any other method so far are TSSCO 1030 kJ/mol and HSCO 1030 kJ/mol [163]. Thus,
investigated. Uniform accuracy across the d-block is valuable upon using only the enthalpies or electronic energies to judge whether
evaluation of compounds e.g. for catalyst design. This universal- a functional predicts the observed ground state (which includes
ity indicates the potential future success of a priori functionals in entropy) introduces a systematic error of this magnitude and such
inorganic chemistry. In a benchmark of binding energies with M05 benchmarks will articially favor methods that are HS-biased, e.g.
which also included TPSSh [167], while M05 had a lower mean post-HF methods and hybrids with >15% HF exchange. As described,
absolute error than TPSSh (33 vs. 41 kJ/mol), TPSSh in fact com- both entropy and ZPE, added as corrections after electronic con-
fortably gave the smallest signed error (5 kJ/mol) of all functionals vergence, almost uniformly favor the looser HS state, because
studied. This could suggest that TPSSh eliminates most of the sys- the occupied ligand-directed (anti-bonding) d-orbitals in the HS
tematic error also for saturated complexes. states weaken and elongate M L bonds, lower ZPE, and increase
Despite the elimination of the systematic error component, vibrational entropy [126]. SCO enthalpies for benchmarking are
large random errors of 40 kJ/mol remain in absolute BDEs across interesting because of available, accurate experimental data [163]
the d-block when modeled with TPSSh (>50 kJ/mol with B3LYP, PBE, and because the process depends critically on balanced Fermi cor-
BP86, etc.). A large part of this random error is due to pathologi- relation that all functionals should strive to achieve. However, the
cal cases confronting all functionals, e.g. the previously mentioned data are restricted to dq congurations in the center of the d-block
4s1 dq+1 4s2 dq promotion energies that embody many of the (q = 47), which precludes probing for universality.
effects discussed above [11], notably the inert pair effect, the SIE
(if the d-conguration is JahnTeller instable), and HF exchange. 5.2. HF exchange and spin states
Therefore, these promotion energies are particularly interesting to
understand in future work. As described above, HF exchange causes hybrid function-
MAEs of 40 kJ/mol for M L BDEs may sound discouraging and als to favor HS whereas non-hybrids favor LS too much
are substantially larger than the average MAEs of main group reac- [100,101,180,181]. Thus, in SCO complexes, use of B3LYP* with
tion energies, typically in the range of 525 kJ/mol for hybrids 15% [178] or B3LYP** [179] or TPSSh [160] with 10% HF exchange
and double hybrids [52]. However, the M L BDEs are extremely provides better results than B3LYP. As the electronic HSLS gap is
challenging cases with substantial differential correlation energies. nearly linear in HF exchange [100], it is not surprising that TPSSh
In larger, substitution-saturated systems, energy differences and with 10% HF exchange usually falls in between non-hybrids and
absolute errors tend to be down-scaled due to the higher density B3LYP [126,160]. For nine iron and cobalt complexes (see Fig. 5),
of states. TPSSh achieved a mean absolute error of 11 kJ/mol, probably close
Many reactions are isodesmic, e.g. conformational and isomer- to the limit of absolute errors attainable from theoretical chemistry
ization energies that have modest differential correlation energies, for these complex systems [160]. For purely main-group molecular
and many other reactions can be studied via their relative effects systems, TPSSh retains an accuracy similar to B3LYP for the G3/99
based on a reference reaction, thus eliminating many of the hid- set [43,159].
den systematic errors mentioned above e.g. differential energies Diverse spin states are particularly widespread in the vast eld
of the type: of porphyrin chemistry, of substantial importance in bioinorganic,
synthetic, and materials chemistry [20,182]. Thus, porphyrins
M L + M L M L + M L (8) are important and challenging test cases for the reliable use of
206 K.P. Kepp / Coordination Chemistry Reviews 257 (2013) 196209

type [187]. B2PLYP-D3, i.e. with the latest dispersion correction by


Grimme, also performed well in a benchmark of articial molecules
with less bias toward specic electronic structures, when compared
against CCSD(T) calculations [188]. B2PLYP accurately describes
some electronic excitation energies [189] and is a very promising
candidate for further development among the new functionals
that are steadily challenging B3LYPs market dominance [190].

6. Larger-system effects not described in this review

Although this paper has been concerned with the systematic


effects in Table 1 and how their corrections affect studies and
benchmarking of methods, the eld of inorganic chemistry is of
course much larger than the chemistry of individual M L bonds.
Many reviews only mentioned in the passing deal with larger
Fig. 5. Enthalpy differences between high- and low-spin states of some iron- molecular systems and broad ranges of properties and should be
and cobalt complexes computed with different functionals, including thermal and consulted for these purposes [5,6,1023].
relativistic effects. Red bars signify experimental ranges. The acronyms refer to
Some notable effects occurring in larger systems are dispersion,
complexes dened in Ref. [160].
which as discussed favors compact states, and coordinative satura-
Copyright 2009: The American Chemical Society.
tion that increases the density of states and thus static correlation,
but also reduces apparent errors by scaling down energy differ-
modern electron-structure methods. In a recent work [183], DFT- ences. This usually means that the errors observed in M L BDEs
calculated energies of a number of electronic congurations of ve (with discouraging, random error components of up to 40 kJ/mol)
iron(III)porphine complexes were compared to the experimentally are larger than in saturated complexes, unless large systematic
observed ground states [184] of these complexes. Given the above errors from e.g. dispersion are present, but it also easily leads to
results, it was surprising to observe that TPSSh is less accurate partial cancelation of errors and obscures the insight that is more
for these systems [126]. However, these conclusions rested on the clear in simpler systems.
premise that direct electronic DFT-energies could probe thermally For larger systems, choosing a realistic model system that
observed ground states, and inclusion of entropy, ZPE, and disper- includes the most signicant physical interactions becomes critical
sion rendered TPSSh the most accurate functional, consistent with [10,13,147,149]. Another important non-systematic effect rarely
other results [160]. dealt with is the sampling or local minimum problem, not of
Another example of erroneous conclusions from neglect of electronic congurations (Section 2.4.), but of conformations and
entropy is the conclusion that B2PLYP [34] performs better than isomers in coordination complexes and clusters, as investigated in
TPSSh for spin-crossover systems [185]. This conicts with the a few cases such as di-iron enzymes [162] and inorganic clusters
results cited above [126,160] and is also anomalous given the [191,192]. The sampling of isomers in these cases clearly affects
large component (53%) of HF exchange in B2PLYP, since HF the identication of species that are considered probes of the
exchange linearly affects spin splitting energies with 20% being experimental system. This problem is even more substantial in the
commonly too large [100,101,104,105]. Again, directly calculated modeling of proteins and cannot currently be dealt with, except in
electronic energies of the relevant congurations were used as relatively small systems.
proxies for experimentally observed ground states, neglecting Also, in larger systems, the chemical complexity may increase
entropy [185]. It is known both experimentally and theoretically e.g. by the presence of metalmetal bonds, which have also not
[105,126,163,186] that HS states are favored by entropy by typi- been discussed here. Metalmetal bonds are often electronically
cally 20 kJ/mol per metal site, and its neglect leads to erroneous even more complex due to involvement of more orbitals, rendering
appraisals of HF exchange and conclusions regarding the relative electron correlation quite distinct from that in most metalligand
merits of functionals. bonds, giving other trends than those discussed here [193]: for
For the archetypical spin-crossover complex Fe(1,10- example, non-hybrids (such as TPSS) perform better than most
phenantroline)2 -(NCS)2 , electronic energies alone would show hybrids, probably due to the massive static correlation that reduces
TPSSh favoring LS by 17 kJ/mol [185], but upon entropy cor- the accuracy of single-conguration HF exchange [167].
rection, the estimates will be very close to SCO as observed When it comes to molecular properties, the situation is equally
experimentally, whereas B2PLYP will over-stabilize the HS state, discouraging as for molecular systems, with different properties
as expected from its large HF exchange component. After cor- requiring different choices of methods [6,1416,48]. However, as
recting by 20 kJ/mol as a reasonable entropy, TPSSh is the most attempted in this review, using the simplest possible systems to
accurate functional also for this data set, removing the conict reduce errors from e.g. dispersion, electronic conguration bias,
between that study and others. A dispersion correction would sampling, entropy, and solvent might be a path forward for each
work to the opposite effect, probably by less than the 515 kJ/mol property separately.
seen for substituted phenyl iron-porphine complexes [126], still
rendering TPSSh substantially more accurate than previously 7. Concluding remarks
found, and such new calculations might in fact be encouraged.
In conclusion, the practice of DFT prediction of thermal ground The Pareto principle (or 80-20 rule) states that in many situ-
states from electronic energies alone while neglecting entropy ations, 80% of the results can be achieved from 20% of the input,
should be discouraged, and 10% HF exchange seems appropriate if the underlying dynamics are governed by power-law behavior
in these cases when the effects of Table 1 are accounted for. (exact numbers depend on the power-law exponent) [194]. Plausi-
Despite this, the general experience with B2PLYP has been bly, modern computational chemistry, as indeed many other elds
positive: variations of the functional perform well for main- of modeling, suffers from its own 80-20 rule: while the exact
group thermochemistry involving also cases with some static distribution depends on methodology and systems studied, a very
correlation, although the best version depends on reaction large part of the electron correlation energy is generally obtained
K.P. Kepp / Coordination Chemistry Reviews 257 (2013) 196209 207

by a small fraction of the computational work, whereas the remain- B3LYP 3-parameter hybrid GGA functional by Becke, using func-
ing small fraction of the correlation energy requires the majority of tionals developed by Colle-Salvetti, Lee, Yang, Parr, and
the total efforts. This is clearly true in wave function theory, where others, with 20% HF exchange
one- and many-electron spaces rapidly restrict accuracy due to B3LYP* B3LYP functional with 15% HF exchange instead of
size restrictions. While this distribution is less extreme in DFT, it is 20%
still apparent from the increased algorithmic complexity associated BDE bond dissociation enthalpy
with new functionals. BP86 combination of Becke 88 exchange and Perdew 1986 cor-
However, as the central message of this review, additional relation functionals
effects, often systematic and partly neglected, play comparable CASPT2 CASSCF subject to second-order perturbation the-
or even larger roles: ZPE, entropy, HF exchange, some relativistic ory to dynamically correlate a CASSCF wave func-
effects (notably scalar relativistic effects and spinorbit coupling on tion: a useful method in inorganic chemistry since
higher-angular momentum bonds), and electronic self-interaction it systematically includes both static and dynamic
tend to weaken bonds either directly (HF exchange, relativistic correlation
effects, and SIE) or indirectly, as a correction to the converged elec- CASSCF complete active space self-consistent eld, a multi-
tronic structure (ZPE, entropy). They mimic static correlation and conguration method by Roos and co-workers that
thus favor higher spin-, radial-, and angular quantum numbers, performs a full CI optimization of all congurations
with resulting bias toward high-spin, excited, atomic, and disso- within an active space of the molecules MOs, thus
ciated states throughout chemistry. In contrast, dispersion, some describing static correlation not obtained in post-HF
solvent effects, and relativistic effects in low angular-momentum single-reference ab initio methods
bonded systems such as M-X -bonds, will often work to the oppo- CCSD(T) coupled-cluster with singlets, doublets, and parameter-
site effect of favoring compact electron densities. ized triplets
As seen in Table 1, errors are often of comparable magnitude and CCSDT coupled-cluster with singlets, doublets, and triplets exci-
usually larger than 10 kJ/mol each, often summing to 50 kJ/mol tations included
for non-isodesmic reaction energies. Lack of attention to any of CI conguration interaction, wave function theory with use
them may lead to systematic errors and erroneous conclusions. For of complex wave functions that are superpositions of
example, studies that neglect entropy, scalar-relativistic effects, several electronic congurations, typically Slater deter-
and ZPE may erroneously favor methods that resemble the HF minants, to incorporate electron correlation
determinant to higher extent, e.g. B3LYP. Thus, accounting for DFT density functional theory
these systematic effects, preferably in benchmarks that elucidate GGA generalized gradient approximation
them carefully, may be worthwhile before complicating density HF HartreeFock, simple wave function method that uses
functionals further. one antisymmetrized Slater determinant as wave func-
As many researchers have testied [13,22,62,81,167], there tion
is no one-size-ts-all in current DFT, although the inclusion of HS high-spin
all the mentioned systematic effects should render it more clear IS intermediate-spin
which functional to use for a given property. However, geometries KS KohnSham procedure, using a HF-like SCF procedure to
can be computed with any standard functional of at least GGA- optimize MOs under the inuence of the functional
type, once the necessary effects (e.g. solvation in charged states) LS low-spin
are included [6264]. For energies, some researchers including M0x Minnesota-type functionals, developed by Truhlar and
the present author hold that HF exchange should be 1015% for co-workers (examples include M05, M06-L, M06-2X,
GGA- and meta-hybrids as in B3LYP* and TPSSh, whereas non- M08-HX)
hybrid GGA functionals tend to perform worse [81,99,126,128]. MO molecular orbital
Scalar-relativistic effects and dispersion e.g. in the form of DFT-D3 MP2 MllerPlesset second order perturbation theory on a HF
[124] should be evaluated routinely. For barriers used to derive reference determinant
rate constants, no general consensus exists, although SIE is often PBE Perdew Burke Ernzerhof a priori GGA functional
a larger problem that suggests use of more HF exchange as e.g. PBE0 25% hybrid version of PBE
in B2PLYP [187,195197]. Finally, new functionals such as double PM3 parameterized model 3, a semi-empirical model by Stew-
hybrids should be considered valid alternatives, although better art
experience is warranted. ROHF restricted open-shell HartreeFock (also applies to DFT),
Most important is arguably the need to pick the low-hanging meaning that all MOs are required to be doubly occupied
fruit by eliminating the discussed systematic errors, to generally except sometimes the highest occupied MOs, depending
use several functionals for reality checking results, to always com- on the state
pare results to experimental data wherever possible during the SCF self-consistent eld, a concept used in HF theory of single
project, to use advanced correlated methods to validate and under- electrons moving in an effective, iteratively solved eld
stand DFT energies, and to design carefully problems to incorporate of other electrons
large extents of error cancelation, e.g. only quantifying energies rel- SCO spin crossover
ative to a reference. Hopefully, such a good practice may remove SIE self-interaction error (or self-interaction energy as this is
some of the confusion that currently resides within the inorganic an energy)
chemistry community regarding the performance and choices of TPSS Tao Perdew Staroverov, Scuseria a priori meta GGA func-
methods for particular modeling tasks. tional
TPSSh 10% hybrid version of TPSS
UHF unrestricted HartreeFock (also applies to DFT, where it
List of abbreviations is called UKS for unrestricted KohnSham), meaning that
- and -spin electrons are treated as separate spaces,
i.e. MOs are singly occupied. This allows spin polariza-
B2PLYP recently developed double hybrid by Grimme and co- tion of core electrons and provides a better description of
workers correlation in most cases
208 K.P. Kepp / Coordination Chemistry Reviews 257 (2013) 196209

WFT wave function theory, usually based on a HartreeFock [59] A. Ruzsinszky, J.P. Perdew, Comput. Theor. Chem. 963 (2011) 26.
reference wave function [60] C.W. Bauschlicher Jr., A. Ricca, H. Patridge, in: D.P. Chong (Ed.), Recent
Advances in Density Functional Methods, World Scientic, 1997, p. 165
ZPE zero point energy (Chapter 6).
[61] J.P. Perdew, A. Ruzsinszky, Int. J. Quantum Chem. 110 (2010) 28012807.
[62] K.P. Jensen, B.O. Roos, U. Ryde, J. Chem. Phys. 126 (2007) 014103.
Acknowledgements [63] M. Bhl, C. Reimann, D.A. Pantazis, T. Bredow, F. Neese, J. Chem. Theory Com-
put. 4 (2008) 14491459.
The Danish Center for Scientic Computing (DCSC) and the Dan- [64] M. Bhl, H. Kabrede, J. Chem. Theory Comput. 2 (2006) 12821290.
[65] S.F. Sousa, E.S. Carvalho, D.M. Ferreira, I.S. Tavares, P.A. Fernandes, M.J. Ras-
ish Science Research Council (FNU, Steno Grant: 272-08-0041) are mos, J.A.N.F. Gomes, J. Comput. Chem. 30 (2009) 27522763.
acknowledged for supporting this work. [66] Y. Zhao, D.G. Truhlar, J. Phys. Chem. A 108 (2004) 69086918.
[67] K.P. Jensen, U. Ryde, J. Mol. Struct. (THEOCHEM) 585 (2002) 239255.
[68] M.H. Olsson, U. Ryde, B.O. Roos, Protein Sci. 7 (1998) 26592668.
References [69] P.J. Hay, J. Chem. Phys. 66 (1977) 43774384.
[70] K.P. Jensen, B.L. Ooi, H.E.M. Christensen, Inorg. Chem. 46 (2007) 87108716.
[1] W. Kohn, A.D. Becke, R.G. Parr, J. Phys. Chem. 100 (1996) 1297412980. [71] J. Grfenstein, D. Cremer, Mol. Phys. 99 (2001) 981989.
[2] W. Kohn, L.J. Sham, Phys. Rev. 140 (1965) A1133A1138. [72] K.P. Jensen, U. Ryde, J. Biol. Chem. 279 (2004) 1456114569.
[3] R.G. Parr, W. Yang, Density Functional Theory of Atoms and Molecules, Oxford [73] J.G. Norman Jr., P.B. Ryan, L. Noodleman, J. Am. Chem. Soc. 102 (1980)
University Press, New York, 1989. 42794282.
[4] http://www.nobelprize.org/nobel prizes/chemistry/laureates/1998/. [74] J.F. Stanton, J. Gauss, Adv. Chem. Phys. 125 (2003) 101146.
[5] P. Geerlings, F. De Proft, W. Langenaeker, Chem. Rev. 103 (2003) 17931873. [75] J. Paldus, X. Li, Adv. Chem. Phys. 110 (1999) 1175.
[6] C.J. Cramer, D.G. Truhlar, Phys. Chem. Chem. Phys. 11 (2009) 1075710816. [76] P. Piecuch, K. Kowalski, I.S.O. Pimienta, S.A. Kucharski, in: M.R. Hoffmann,
[7] E.I. Solomon, M.D. Lowery, Science 259 (1993) 15751581. K.G. Dyall (Eds.), Low-lying Potential Energy Surfaces, ACS Symp. Ser. 828,
[8] E.I. Solomon, T.C. Brunold, M.I. Davis, J.N. Kemsley, S.-K. Lee, N. Lehnert, F. Washington, DC, 2002.
Neese, A.J. Skulan, Y.-S. Yang, J. Zhou, Chem. Rev. 100 (2000) 235349. [77] P. Piecuch, L. Adamowicz, J. Chem. Phys. 102 (1995) 898904.
[9] U. Ryde, K. Nilsson, J. Am. Chem. Soc. 125 (2003) 1423214233. [78] A.I. Krylov, Acc. Chem. Res. 39 (2006) 8391.
[10] U. Ryde, Dalton Trans. (2007) 607625. [79] J. Shen, Z. Kou, E. Xu, S. Li, J. Chem. Phys. 133 (2010) 234106.
[11] M.C. Holthausen, J. Comput. Chem. 26 (2005) 15051518. [80] B.O. Roos, Acc. Chem. Res. 32 (1999) 137144.
[12] T. Ziegler, Chem. Rev. 91 (1991) 651667. [81] F. Neese, D.G. Liakos, S. Ye, J. Biol. Inorg. Chem. 16 (2011) 821829.
[13] P.E.M. Siegbahn, J. Biol. Inorg. Chem. 11 (2006) 695701. [82] B.O. Roos, K. Andersson, M.P. Flscher, P.-. Malmqvist, L. Serrano-Andres,
[14] S.F. Sousa, P.A. Fernandes, M.J. Ramos, J. Phys. Chem. A 111 (2007) Adv. Chem. Phys. 93 (1996) 219331.
1043910452. [83] T.K. Todorova, L. Gagliardi, J.R. Walensky, K.A. Miller, W.J. Evans, J. Am. Chem.
[15] K.E. Riley, B.T. Opt Holt, K.M. Merz Jr., J. Chem. Theory Comput. 3 (2007) Soc. 132 (2010) 1239712403.
407433. [84] K. Pierloot, Mol. Phys. 101 (2003) 20832094.
[16] F. Neese, Coord. Chem. Rev. 253 (2009) 526563. [85] N.B. Amor, S. Zlis, C. Daniel, Int. J. Quantum Chem. 106 (2006) 24582469.
[17] L. Hermosilla, P. Calle, J.M.G. de la Vega, C. Sieiro, J. Phys. Chem. A 109 (2005) [86] S. Zlis, N.B. Amor, C. Daniel, Inorg. Chem. 43 (2004) 79787985.
11141124. [87] A. Ghosh, B.J. Persson, P.R. Taylor, J. Biol. Inorg. Chem. 8 (2003) 507511.
[18] S. Li, D.A. Dixon, J. Phys. Chem. A 111 (2007) 1190811921. [88] A. Ghosh, P.R. Taylor, J. Chem. Theory Comput. 1 (2005) 597600.
[19] E. Oldeld, Phil. Trans. R. Soc. B 360 (2005) 13471361. [89] K.P. Jensen, J. Phys. Chem. B 109 (2005) 1050510512.
[20] A. Ghosh, J. Biol. Inorg. Chem. 11 (2006) 712724. [90] K.P. Jensen, B.O. Roos, U. Ryde, J. Inorg. Biochem. 99 (2005) 4554.
[21] P.E.M. Siegbahn, M.R.A. Blomberg, Chem. Rev. 100 (2000) 421437. [91] K. Pierloot, Int. J. Quantum Chem. 111 (2011) 32913301.
[22] J.N. Harvey, Annu. Rep. Prog. Chem. C 102 (2006) 203226. [92] S. Villaume, A. Strich, C.A. Ndoye, C. Daniel, S.A. Perera, R.J. Bartlett, J. Chem.
[23] C. Daniel, Coord. Chem. Rev. 238239 (2003) 143166. Phys. 126 (2007) 154318.
[24] J.C. Slater, Phys. Rev. 81 (1951) 385390. [93] S. Villaume, C. Daniel, A. Strich, S.A. Perera, R.J. Bartlett, J. Chem. Phys. 126
[25] E. Fermi, Rend. Accad. Naz. Lincei 6 (1927) 602607. (2005) 044313.
[26] L.H. Thomas, Proc. Camb. Philos. Soc. 23 (1927) 542548. [94] A. Vlcek Jr., S. Zlis, Coord. Chem. Rev. 251 (2007) 258287.
[27] P.A.M. Dirac, Proc. Camb. Philos. Soc. 26 (1930) 376. [95] J.P. Perdew, M. Ernzerhof, A. Zupan, K. Burke, in: J.M. Seminario, P.-O. Lwdin,
[28] J.P. Perdew, Phys. Rev. Lett. 55 (1985) 1665. J.R. Sabin, M.C. Zerner (Eds.), Advances in Density Functional Theory, vol. 33,
[29] J.P. Perdew, Phys. Rev. B 33 (1986) 88228824. Academic Press, San Diego, 1999.
[30] A.D. Becke, Phys. Rev. A 38 (1988) 30983100. [96] C.J. Barden, J.C. Rienstra-Kiracofe, H.F. Schaefer III, J. Chem. Phys. 113 (2000)
[31] P. Jemmer, P.J. Knowles, Phys. Rev. A 51 (1995) 35713575. 690700.
[32] A.D. Becke, J. Chem. Phys. 104 (1996) 1040. [97] S.R. Langhoff, C.W. Bauschlicher Jr., Annu. Rev. Phys. Chem. 39 (1988)
[33] A.D. Becke, J. Chem. Phys. 98 (1993) 56485652. 181212.
[34] S. Grimme, J. Chem. Phys. 124 (2006) 034108. [98] J. Harris, R.O. Jones, J. Phys. F 4 (1974) 11701186.
[35] Y. Tawada, T. Tsuneda, S. Yanagisawa, T. Yanai, K. Hirao, J. Chem. Phys. 120 [99] K.P. Jensen, Inorg. Chem. 47 (2008) 1035710365.
(2004) 8425. [100] M. Reiher, Inorg. Chem. 41 (2002) 69286935.
[36] J.D. Chai, M. Head-Gordon, J. Chem. Phys. 128 (2008) 084106. [101] H. Paulsen, L. Duelund, H. Winkler, H. Toftlund, A.X. Trautwein, Inorg. Chem.
[37] D. Joubert, M. Levy, Phys. Rev. A 54 (1996) 961963. 40 (2001) 22012203.
[38] M. Levy, A. Grling, Phys. Rev. A 51 (1995) 28512856. [102] S. Zein, S.A. Borshch, P. Fleurat-Lessard, M.E. Casida, H. Chermette, J. Chem.
[39] K. Burke, J.P. Perdew, M. Levy, Phys. Rev. A 53 (1996) R2915R2917. Phys. 126 (2007) 014105.
[40] J.P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett. 77 (1996) 38653868. [103] L.M.L. Daku, A. Vargas, A. Hauser, A. Fouqueau, M.E. Casida, ChemPhysChem
[41] J. Tao, J.P. Perdew, V.N. Staroverov, G.E. Scuseria, Phys. Rev. Lett. 91 (2003) 6 (2005) 13931410.
146401. [104] M. Reiher, O. Salomon, B.A. Hess, Theor. Chem. Acc. 107 (2001) 4855.
[42] J.P. Perdew, M. Ernzerhof, K. Burke, J. Chem. Phys 105 (1996) 99829985. [105] H. Paulsen, A.X. Trautwein, Top. Curr. Chem. 235 (2004) 197219.
[43] V.N. Staroverov, G.E. Scuseria, J. Tao, J.P. Perdew, J. Chem. Phys. 119 (2003) [106] G. Schenk, M.Y.M. Pau, E.I. Solomon, J. Am. Chem. Soc. 126 (2004) 505515.
1212912137. [107] M.M. Quintal, A. Karton, M.A. Iron, A.D. Boese, J.M.L. Martin, J. Phys. Chem. A
[44] S. Kurth, J.P. Perdew, P. Blaha, Int. J. Quantum Chem. 75 (1999) 889909. 110 (2006) 709716.
[45] P.J. Stephens, F.J. Devlin, C.F. Chabalowski, M.J. Frisch, J. Phys. Chem. 98 (1994) [108] V. Polo, E. Kraka, D. Cremer, Mol. Phys. 100 (2002) 17711790.
1162311627. [109] O.A. Vydrov, Correcting the self-interaction error of approximate density
[46] C. Lee, W. Yang, R.G. Parr, Phys. Rev. B 37 (1988) 785789. functionals, Ph.D. Thesis, Rice University, Houston, TX, 2007.
[47] S.H. Vosko, L. Wilk, M. Nusair, Can. J. Phys. 58 (1980) 12001211. [110] O.A. Vydrov, G.E. Scuseria, J. Chem. Phys. 121 (2004) 8187.
[48] Y. Zhao, D.G. Truhlar, Theor. Chem. Acc. 120 (2008) 215241. [111] P. Mori-Sanchez, A.J. Cohen, W. Yang, J. Chem. Phys. 125 (2006) 201102.
[49] J.J.P. Stewart, J. Comput. Chem. 10 (1989) 209220. [112] Y. Zhang, W. Yang, J. Chem. Phys. 109 (1998) 26042608.
[50] J.J.P. Stewart, J. Comput. Chem. 12 (1991) 320341. [113] J. Grfenstein, E. Kraka, D. Cremer, J. Chem. Phys. 120 (2004) 524539.
[51] Y. Zhao, N.E. Schultz, D.G. Truhlar, J. Chem. Phys. 123 (2005) 161103. [114] M. Lundberg, P.E.M. Siegbahn, J. Chem. Phys. 122 (2005) 224103.
[52] J.A. Pople, in: D.W. Smith (Ed.), Proceedings of the Summer Research Confer- [115] D. Cremer, Mol. Phys. 99 (2001) 18991940.
ence on Theoretical Chemistry, Energy Structure and Reactivity, John Wiley [116] J. Grfenstein, D. Cremer, Theor. Chem. Acc. 123 (2009) 171182.
& Sons, Ltd., New York, 1973. [117] E. Ruiz, S. Alvarez, J. Cano, V. Polo, J. Chem. Phys. 123 (2005) 164110.
[53] R.H. Hertwig, W. Koch, Chem. Phys. Lett. 268 (1997) 345351. [118] A.J. Johansson, M.R.A. Blomberg, P.E.M. Siegbahn, J. Chem. Phys. 129 (2008)
[54] L. Goerigk, S. Grimme, Phys. Chem. Chem. Phys. 13 (2011) 66706688. 154301.
[55] M.J.G. Peach, D.J. Tozer, N.C. Handy, Int. J. Quantum Chem. 111 (2011) [119] F. Furche, J.P. Perdew, J. Chem. Phys. 124 (2006) 044103.
563569. [120] S. Grimme, J. Comput. Chem. 25 (2004) 14631473.
[56] G. Frenking, N. Frhlich, Chem. Rev. 100 (2000) 717774. [121] S. Grimme, J. Comput. Chem. 27 (2006) 17871799.
[57] P.-O. Lwdin, Adv. Chem. Phys. 2 (1959) 207322. [122] S.N. Steinmann, C. Corminboeuf, J. Chem. Theory Comput. 7 (2011)
[58] J.C. Slater, Phys. Rev. 34 (1929) 12931322. 35673577.
K.P. Kepp / Coordination Chemistry Reviews 257 (2013) 196209 209

[123] A. Tkatchenko, M. Schefer, Phys. Rev. Lett. 192 (2009) 073005. [161] G. Roos, P. Geerlings, J. Messens, J. Phys. Chem. B 113 (2009) 1346513475.
[124] S. Grimme, J. Antony, S. Ehrlich, H. Krieg, J. Chem. Phys. 132 (2010) [162] K.P. Jensen, C.B. Bell III, M.D. Clay, E.I. Solomon, J. Am. Chem. Soc. 131 (2009)
154104. 1215512171.
[125] S. Grimme, S. Ehrlich, L. Goerigk, J. Comput. Chem. 32 (2011) 14561465. [163] J.W. Turner, F.A. Schultz, Coord. Chem. Rev. 219221 (2001) 8197.
[126] K.P. Kepp, J. Inorg. Biochem. 105 (2011) 12861292. [164] G. Brewer, M.J. Olida, A.M. Schmiedekamp, C. Viragh, P.Y. Zavalij, Dalton Trans.
[127] K.P. Jensen, U. Ryde, J. Phys. Chem. A 107 (2003) 75397545. (2006) 56175629.
[128] P.E.M. Siegbahn, M.R.A. Blomberg, S.-L. Chen, J. Chem. Theor. Comput. 6 (2010) [165] M.P. Waller, H. Braun, N. Hojdis, M. Bhl, J. Chem. Theory Comput. 3 (2007)
20402044. 22342242.
[129] M.D. Wodrich, C. Corminboeuf, P. von Rage Schleyer, Org. Lett. 8 (2006) [166] J.P. Perdew, A. Ruzsinszky, L.A. Constantin, J. Sun, G.I. Csonka, J. Chem. Theory
36313634. Comput. 5 (2009) 902908.
[130] L.A. Burns, A. Vazquez-Mayagoitia, B.G. Sumpter, C.D. Sherrill, J. Chem. Phys. [167] Y. Zhao, D.G. Truhlar, J. Chem. Phys. 124 (2006) 224105.
134 (2011) 084107. [168] K.P. Jensen, in: E.I. Solomon, R.A. Scott, R. Bruce King (Eds.), Computational
[131] Y. Zhao, D.G. Truhlar, J. Chem. Theory Comput. 1 (2005) 415432. Inorganic and Bioinorganic Chemistry, John Wiley & Sons, Ltd., Chichester,
[132] P. Pyykko, Chem. Rev. 88 (1988) 563594. UK, 2009, pp. 373386.
[133] D.J. Gorin, F.D. Toste, Nature 446 (2007) 395403. [169] K.P. Jensen, U. Ryde, ChemBioChem 4 (2003) 413424.
[134] F. Neese, E.I. Solomon, Inorg. Chem. 37 (1998) 65686582. [170] J.A. Real, A.B. Gaspar, M.C. Munoz, Dalton Trans. 12 (2005) 20622079.
[135] P.A. Christiansen, W.C. Ermler, K.S. Pitzer, Annu. Rev. Phys. Chem. 36 (1985) [171] H.A. Goodwin, Coord. Chem. Rev. 18 (1976) 293325.
407432. [172] N. Strickland, J.N. Harvey, J. Phys. Chem. B 111 (2007) 841852.
[136] K.P. Jensen, J. Phys. Chem. A 113 (2009) 1013310141. [173] E. Knig, K. Madeja, Inorg. Chem. 7 (1967) 18481855.
[137] M. Hargittai, Acc. Chem. Res. 42 (2009) 453462. [174] P. Gtlich, A. Hauser, H. Spiering, Angew. Chem. Int. Ed. Engl. 33 (1994)
[138] I.B. Bersuker, Electronic Structure and Properties of Transition Metal Com- 20242054.
pounds, Wiley & Sons, New Jersey, 2010. [175] R.J. Deeth, A.E. Anastasi, M.J. Wilcockson, J. Am. Chem. Soc. 132 (2010)
[139] D.J. Grant, M.H. Matus, J.R. Switzer, D.A. Dixon, J.S. Francisco, K.O. Christe, J. 68766877.
Phys. Chem. A 112 (2008) 31453156. [176] A.V. Postnikov, G. Bihlmayer, S. Blugel, Comput. Mater. Sci. 36 (2006) 9195.
[140] C.W. Bauschlicher Jr., Theor. Chem. Acc. 101 (1999) 421425. [177] A. Sorkin, M.A. Iron, D.G. Truhlar, J. Chem. Theory Comput. 4 (2008) 307315.
[141] J.M.L. Martin, A. Sundermann, P.L. Fast, D.G. Truhlar, J. Chem. Phys. 113 (2000) [178] J. Conradie, A. Ghosh, J. Phys. Chem. B 111 (2007) 1262112624.
13481358. [179] I. Respondek, L. Bressel, P. Saalfrank, H. Kampf, A. Grohmann, Chem. Phys. 347
[142] S.-G. Wang, W. Liu, W.H.E. Schwarz, J. Phys. Chem. A 106 (2002) 795803. (2008) 514522.
[143] K.S. Thanthiriwatte, E.G. Hohenstein, L.A. Burns, C.D. Sherrill, J. Chem. Theory [180] M. Swart, A.R. Groenhof, A.W. Ehlers, K. Lammertsma, J. Phys. Chem. A 108
Comput. 7 (2011) 8896. (2004) 54795483.
[144] K.P. Jensen, J. Inorg. Biochem. 102 (2008) 87100. [181] R.J. Deeth, N. Fey, J. Comput. Chem. 25 (2004) 18401848.
[145] K.P. Jensen, M. Rykr, Dalton Trans. 39 (2010) 96849695. [182] W.R. Scheidt, C.A. Reed, Chem. Rev. 81 (1981) 543555.
[146] A. Klamt, G.J. Schrmann, J. Chem. Soc. Perkin Trans. 2 (1993) 799805. [183] M.M. Conradie, J. Conradie, A. Ghosh, J. Inorg. Biochem. 105 (2011) 8491.
[147] P.E.M. Siegbahn, F. Himo, J. Biol. Inorg. Chem. 14 (2009) 643651. [184] M.K. Kadish, E. Van Caemelbecke, E. Gueletii, S. Fukuzumi, K. Miyamoto, T.
[148] P.E.M. Siegbahn, F. Himo, Wiley Interdiscip. Rev.: Comput. Mol. Sci. 1 (2011) Suenobu, A. Tabard, R. Guilard, Inorg. Chem. 37 (1998) 17591766.
323336. [185] S. Ye, F. Neese, Inorg. Chem. 49 (2010) 772774.
[149] L. Hu, J. Eliasson, J. Heimdal, U. Ryde, J. Phys. Chem. A 113 (2009) 1179311800. [186] G. Baranovic, Chem. Phys. Lett. 369 (2003) 668672.
[150] K.P. Kepp, J. Inorg. Biochem. 107 (2012) 1524. [187] A. Karton, A. Tarnopolsky, J.-F. Lamere, G.C. Schatz, J.M.L. Martin, J. Phys. Chem.
[151] A. Warshel, J. qvist, Annu. Rev. Biophys. Biophys. Chem. 20 (1991) 267298. A 112 (2008) 1286812886.
[152] M. Naja, M. Zahedi, E. Klein, Comput. Theor. Chem. 978 (2011) 1628. [188] M. Korth, S. Grimme, J. Chem. Theory Comput. 5 (2009) 9931003.
[153] T. Dudev, Y.-I. Lin, M. Dudev, C. Lim, J. Am. Chem. Soc. 125 (2003) 31683180. [189] S. Grimme, F. Neese, J. Chem. Phys. 127 (2007) 154116.
[154] K.P. Jensen, U. Ryde, J. Am. Chem. Soc. 125 (2003) 1397013971. [190] T. Schwabe, S. Grimme, Acc. Chem. Res. 41 (2008) 569579.
[155] H. Goff, G.N. La Mar, C.A. Reed, J. Am. Chem. Soc. 99 (1977) 36413646. [191] A.N. Alexandrova, A.I. Boldyrev, J. Chem. Theory Comput. 1 (2005) 566580.
[156] J.P. Collman, J.L. Hoard, N. Kim, G. Lang, C.A. Reed, J. Am. Chem. Soc. 97 (1975) [192] D.Y. Zubarev, A.N. Alexandrova, A.I. Boldyrev, L.F. Cui, X. Li, L.S. Wang, J. Chem.
26762681. Phys. 124 (2006) 124305.
[157] M. Radon, K. Pierloot, J. Phys. Chem. 112 (2008) 1182411832. [193] S. Yanagisawa, T. Tsuneda, K. Hirao, J. Chem. Phys. 112 (2000) 545553.
[158] T. Kitagawa, J. Teraoka, Chem. Phys. Lett. 63 (1979) 443446. [194] A. Grosfeld-Nir, B. Ronen, N. Kozlovsky, Int. J. Prod. Res. 45 (2007) 23172325.
[159] L.A. Curtiss, K. Raghavachari, P.C. Redfern, J.A. Pople, Chem. Phys. Lett. 270 [195] H.C. Fang, Z.H. Li, K.N. Fan, Phys. Chem. Chem. Phys. 13 (2011) 1335813369.
(1997) 419426. [196] T.R. Ramadhar, R.A. Batey, Comput. Theor. Chem. 976 (2011) 167182.
[160] K.P. Jensen, J. Cirera, J. Phys. Chem. A 113 (2009) 1003310039. [197] S. Goel, A.E. Masunov, Int. J. Quantum Chem. 111 (2011) 42764287.

Potrebbero piacerti anche