Kao 2016

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Molecular Physics

An International Journal at the Interface Between Chemistry and


Physics

ISSN: 0026-8976 (Print) 1362-3028 (Online) Journal homepage: http://www.tandfonline.com/loi/tmph20

Use of Löwdin orthogonalised Fermi orbitals for


self-interaction corrections in an iron porphyrin

Der-you Kao & Mark R. Pederson

To cite this article: Der-you Kao & Mark R. Pederson (2016): Use of Löwdin orthogonalised
Fermi orbitals for self-interaction corrections in an iron porphyrin, Molecular Physics, DOI:
10.1080/00268976.2016.1225992

To link to this article: http://dx.doi.org/10.1080/00268976.2016.1225992

Published online: 08 Sep 2016.

Submit your article to this journal

Article views: 34

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=tmph20

Download by: [University of California, San Diego] Date: 12 December 2016, At: 04:52
MOLECULAR PHYSICS, 
http://dx.doi.org/./..

56TH SANIBEL SYMPOSIUM - SPECIAL ISSUE

Use of Löwdin orthogonalised Fermi orbitals for self-interaction corrections in an


iron porphyrin
Der-you Kaoa and Mark R. Pederson b

a
Mechanical and Aerospace Engineering, The George Washington University, Washington, DC, USA; b Department of Chemistry, Johns Hopkins
University, Baltimore, MD, USA

ABSTRACT ARTICLE HISTORY


A new approach for formulating the self-interaction correction, referred to as the Fermi-Löwdin Received  April 
orbital-based self-interaction correction (FLO-SIC), is briefly reviewed and applied to the Fe(II)- Accepted  August 
porphyrin molecule. The Fermi–Löwdin orbitals are localised but due to the fact they are constructed
KEYWORDS
directly from the spin-density matrices, they lead to a set of orbitals that have the property that Fermi–Löwdin orbital;
each and every orbital is explicitly invariant to unitary transformations. These orbitals allow one to self-interaction correction;
constrain the Perdew–Zunger self-interaction correction in a manner that ensures that the energy Fe(II) porphyrin; density
is explicitly invariant to unitary transformations and size-extensive. We review these improvements functional theory
and show how the multiple solutions, intrinsic to the original formulation, are removed through the
constraint offered by the Fermi–Löwdin construction. The locations of descriptors of Fermi–Löwdin
orbital imitate the formation of classical chemical bonds. The very first application of this method to
a system containing a transition metal atom is presented here.

1. Introduction Fermi–Löwdin orbitals (FLOs). Such orbitals depend


on first constructing normalised Fermi orbitals (FOs)
An approach for implementing the self-interaction cor-
which are explicitly dependent on a single-particle
rection (SIC) to density-functional approximations and
density matrix but depend parametrically on a set of
recovering an asymptotically exact effective potential is
quasi-classical electronic positions or descriptors. When
due to Perdew and Zunger and dates back to 1981 [1].
orthogonalised via Löwdin’s method of symmetric
However, the explicit orbital dependence, or alternatively
orthonormalisation, the formalism allows for the deriva-
unitary non-invariance, associated with this formu-
tion of an N × N unitary matrix that depends explicitly
lation leads to large increases in computational costs
on the density matrix and N quasi-classical electronic
due to the need to solve the localisation equations for
positions that we refer to as electronic descriptors. Thus,
SIC and also leads to an energy expression that is not
for any set of Kohn–Sham orbitals, the self-interaction
size-extensive. In a recent work, a size-extensive unitarily
energy is reduced to a function of 3N parameters that
invariant reformulation of the SIC has been introduced
can be minimised using standard gradient methods.
which proceeds through the construction of energy-
Recent work on applications to small molecules [2],
localised orthonormal orbitals that are referred to as

CONTACT Mark R. Pederson mpeder@jhu.edu


©  Informa UK Limited, trading as Taylor & Francis Group
2 D.-Y. KAO AND M. R. PEDERSON

medium-sized molecules [3], and large molecules [4]


have provided very encouraging results. For example,
using the local density approximation [5] of Perdew and
Wang (PW92), it was found that cohesive energies in
small molecules were significantly improved, especially
for systems with π bonds, and that the HOMO/LUMO
gaps are increased relative to their Kohn–Sham counter-
parts which is the trend one needs to shift the eigenvalues
in a manner that allows the energy of the HOMO state
to match the ionisation energy. To date, there have not
been any applications of the method to systems contain-
ing transition-metal atoms or ions which represent the
types of atoms where the need for SICs was originally
noted by Lindgren [6] and Janak et al. [7]. As such, this
work applies this new formulation to a spin unpolarised
metal porphyrin and to a spin-ordered metal porphyrin
both of which have no net spin. While we are primarily
interested in demonstrating the use of FLO-SIC on the
spin-unpolarised system, we include PW92 results from
a second (lower energy) orbitally ordered system, with
no net moment, since this structure has different fillings
of the 3d orbitals.
Porphyrins, which consist of four pyrrole groups Figure . A bent Fe(II)-porphyrin: hydrogens are white, carbons are
linked by methene bridges with a transition-metal centre, grey, nitrogens are black, and iron is larger central atom. The upper
have been selected by nature as the functional pigment of panel is the top view of Fe(II)-porphyrin and the lower panel is the
side view.
biology. Because transition-metal atoms can exist in mul-
tiple spin and charge states and due to their large natu-
rally occurring terrestrial abundance, nature has chosen rather strong relationships between the spin structure of
such systems to form the basis of many life-sustaining and the electrons and distortions of the porphyrins. Demon-
catalytically interesting molecular systems. The redox strating this for systems with no net moment is instruc-
property of irons centred in porphyrins in heme leads tive, since, in addition to interesting physics, it is a factor
to a charge-induced polarisation of nearby closed shell that must be accounted for if comparisons between dif-
molecules (such as O2 ), and hence the Fe-porphyrin (Fig- ferent calculations are to be meaningful.
ure 1), is able to carry oxygen [8–13]. Although approximations based upon the DFT pro-
Several porphyrin complexes exhibit novel magnetic, vide excellent predictive power for many fundamental
electronic, optical, and photoelectrical properties. For properties, an outstanding challenge has been to reliably
instance, in a study of polypyrrole, containing porphyrin predict some properties of transition-metal-containing
moieties, it has been shown that the molecular system molecules, especially when undercoordinated, such as for
experiences an immediate increase in electric resistance the case of Fe(II)-porphyrin. Lower spin states are known
when carbon monoxide gas binds to the iron at the cen- to be favoured by GGA exchange-correlation function-
tre of the porphyrin [14]. In addition, it has been demon- als, while hybrid functionals that include a fraction of
strated that a self-assembled oligomeric porphyrin ladder Hartree–Fock exchange often prefer high-spin states.
can mimic light harvesting [15] that is found in natu- Furthermore, different energy gaps between spin states
rally occurring photosystems. A network of Fe(II) meso- are obtained with different exchange-correlation func-
tetra(4-pyridyl) porphyrin with an effective moment of tionals [17]. One suggested deficiency of DFT in describ-
5.52 B has been reported with high thermal stability at ing transition metals is that relatively localised 3d valence
room temperature [16]. Preliminary understanding of electrons suffer strongly from self-interaction error pre-
this experimental investigation posits that the observed senting in pure density functionals. Self-interaction error
spin-glass-like behaviour could be caused by either local is only approximately corrected in hybrid functionals.
lattice distortions or the existence of vacancies at the While the work discussed here is on a transition-metal
Fe(II) sites. Both of these effects are expected to modify system, it should be noted that such problems are even
the local exchange coupling. An additional result from more severe in the heavier atoms where even stronger
the work discussed here shows that there are, indeed, errors in the Coulomb interaction cause qualitatively
MOLECULAR PHYSICS 3

incorrect fillings in the f-electron series. A discussion of been determined. Consider, for example, the neon atom
this problem can be found in [18]. which is a closed shell with a 2s2p3 valence. Within this
The need for the use of SICs for systems containing shell, one can construct at least three sets of localised
transition-metal ions was first suggested by Lindgren [6] orbitals that automatically satisfy the localisation equa-
in applications to the Cu atoms. The formulation pro- tions due to symmetry considerations. These sets are the
posed by Lindgren introduced the self-interaction at the cartesian set given by {|2s 〉, |2px 〉, |2py 〉,
level of the single-electron Hamiltonian and proposed |2pz 〉}, an explicitly complex spherical harmonic set
|2p +i|2p  |2p −i|2p 
correcting the error in the interelectronic exchange func- that is given by {|2s, x√(2) y , x√(2) y , |2pz }, and
tional, derived from the free-electron-gas model, in a the standard sp3 hybrids. It was demonstrated in 1988
local-density approximation. In 1981, Perdew and Zunger by Pederson and Lin [20] that the hybrids minimise the
proposed the SIC at the level of the total energy expres- total energy for the Kohn–Sham exchange only func-
sion, so it could be applied to functionals designed to tional. It has been further discussed that this is the only
include correlation and exchange effects [1] and also serve choice that automatically leads to three-fold degenera-
as an ab initio alternative to the LSDA+U method for the cies for the canonical 2p orbitals and it is now recog-
description of strongly correlated systems. nised that this set is the minimising set determined from
Perdew and Zunger posited that the orbital densi- the FLO-SIC formulation [34]. Here, we argue further
ties, ρ i (r) = |φ iσ (r)|2 , could be constructed from a set that since the choice of the nodal features of the min-
of localised orbitals that are similar to the Edmiston– imising set of localised orbitals, in the original PZ func-
Ruedenberg [19] energy-localised orbitals in the Hartree tional, will also depend on the features of the functional
approximation. Later, Pederson et al. [20–22] suggested itself (GGA, LDA, SCAN [35], etc.), the FLO-SIC formu-
the use of localisation equations as a variational means lation would remove this seemingly undesirable trait of
for addressing the unitary non-invariance that existed the original formulation. Based on the considerations of
in the original PZ formulation. This approach has been this paragraph, we argue strongly for the hypothesis that
the primary means for variational SICs as discussed the FLO-SIC formulation must be further considered and
in recent reviews [23–30]. While the original version that arguments for further modification of the PZ for-
of SIC [20–22] only considered real orbitals, several mulation, presented in [34] based on considerations of
researchers [27,31,32] have noted that there is no a pri- results from many authors, may be a better way to further
ori reason to expect that complex orbitals could not lead achieve the nearly nodeless orbital densities discussed in
to lower SIC energies in the original PZ formulation and [33].
have subsequently derived additional formula for finding In Section 2, the FLO-SIC [34] is reviewed. The results
stationary solutions. Pederson and Perdew noted in [33] from a calculation on a spin-unpolarised porphyrin are
that real localised orbitals which minimise the SIC energy reported in Section 3, with additional remarks about this
are stationary with respect to complex unitary transfor- study in Section 4.
mations but again did not guarantee that the real solu-
tions could not be a saddle point or local maximum. More
recently, a modification of the functional has been made 2. Fermi–Löwdin orbital-based self-interaction
through the use of FLOs for the SIC. correction
While we have stressed the fact that this new version of
In PZ self-interaction-corrected expression for a spin-
SIC has the same functional form as the original Perdew–
polarised system, the localised orbital density can be
Zunger (PZ) SIC but offers formal improvements by con-
represented with both localised {φ i,σ (r)} and canonical
straining the functional to be an explicit density func-
{ψ i,σ (r)} orbitals [20–22]:
tional, it is appropriate to reconsider some past discus-
sions on one of the simplest atoms where the original    
formulation leads to multiple stationary solutions for a ρσ (r) = φi,σ (r)2 = ψi,σ (r)2 . (1)
given set of Kohn–Sham orbitals. Except for cases that i i

explicitly violate size-extensivity, the existence of multi-


ple stationary solutions is not a formal problem since it The orbitals used for constructing the SIC energy must
is only the lowest energy configuration of these multiple satisfy O(N2 ) localisation equations given by
solutions that should be compared to experiment. How-

ever, the fact that they exist in the simplest isolated atoms     σ
Hoσ + ViσSIC  φiσ = λ  φ jσ (2)
creates intrinsic uncertainties in applications, since, for ij
j
energy landscapes with multiple stationary points, it is 
always difficult to guarantee that the ground-state has φiσ | ViσSIC − V jσ
SIC
| φ jσ = 0. (3)
4 D.-Y. KAO AND M. R. PEDERSON

Based on the complexity associated with the above Table . For the two-orbital case, there are two ways to use Gram–
equations, a new version based upon FLOs [34] which Schmidt to reorthogonalise orbitals but only one way to deter-
mine new orbitals with the Löwdin’s method. The generalisation
satisfies many desirable properties for self-interaction-
to N orbitals is that there would be N! different ways to construct
corrected functionals has been developed. This advan- orthonormal orbitals if Gram–Schmidt is employed.
tageous new formulation [34] of the self-interaction-
Gram–Schmidt  Gram–Schmidt  Löwdin
corrected density-functional expression, that evades the
requirement for solving the computational demanding φ F F F1 − 21  F1 |F2 F2
localisation equation (Equation (3)), is now discussed. φ F −  F |F 〉F F −  F |F 〉F F2 − 21  F1 |F2 F1
This expression is explicitly unitarily invariant and size-
extensive [2] because it is based on the FO [36,37], con-
structed from the spin-density matrix, which is explicitly The intermediate Löwdin orbitals (ILO), which struc-
invariant under unitary transformations. Given a trial set turally are similar to molecular orbitals and Bloch
of Kohn–Sham orbitals, a normalised FO can be defined Functions, are eigenstates of the overlap matrix and are
at any point in space aiσ according to therefore constructed from a unitary transformation
on the FOs according to


μ ψμσ (aiσ )ψμσ (r) ρσ (aiσ , r) 
Fiσ (r) =
=√ . (4) |Tα  = Tα j |Fj  (7)
|ψ (a )|2 ρσ (aiσ ))
μ iσ j

The eigenvalues of the FO-overlap matrix, Qα , determine


The positions used for construction of the FOs, now
the total amount of charge each ILO captures. From Equa-
referred to as descriptors (aiσ ), provide a semiclassical
tions (5)–(7), the FLOs, designated by φ k , are constructed
picture of an electron as they look like electronic posi-
with the ILO transformation matrix and associated eigen-
tions [2,34]. For plane waves enclosed in a cubic Bril-
values according to
louin zone, ψkσ = √1V eik·r , Equation (4) reduces to Wan-

nier function, √1V k e−ik·aiσ eik·r . When an FO is val-  1 


|φk  = √ Tkα Tα j |Fj  ≡ φkFj |Fj . (8)
ued at aiσ , the FO captures all the spin density at the Qα
αj j
point in space. The FO is reduced to a delta function if
it is constructed with a full orbital set and therefore fully
Because the overlap matrix is real and symmetric, the
recovers the classical ideal of an electronic position. For
inverse transformation between FLOs and FOs is
special sets of points, the FOs can be immediately orthog-
onal (e.g. Wannier functions), but in general, they are not. 
|Fl  = Qβ Tlβ Tβn |φn . (9)
In order to derive an orthonormalised orbital set of
βn
localised orbitals from the initial FOs, Löwdin’s method
of symmetric orthonormalisation is applied. The Löwdin The set of FLOs is guaranteed to be an orthonormal set
symmetric reorthonormalisation procedure guarantees and therefore a unitary transformation of Kohn–Sham
that each localised orbital {φ iσ } is spatially similar to its orbitals. To illustrate the uniqueness of the Löwdin proce-
parent FO in a least-squares sense [38] but each localised dure, we consider another popular systematic method for
orbital now depends parametrically on all the FO descrip- orthonormalising a set of orbitals known as the Gram–
tors and on the spin-density matrix, or equivalently, on Schmidt orthonormalisation. Gram–Schmidt orthonor-
the Kohn–Sham orbitals (Equation (4)). According to malisation does not construct the new set of orbitals
Löwdin’s method of symmetric orthonormalisation, an in a symmetric manner. Considering a two-dimensional
overlap matrix is constructed with FOs according to example, there are two sets of ‘new’ orbitals that can be
constructed using the Gram–Schmidt procedure but only
Si j =  Fi |Fj , (5) one set that can be determined from the Löwdin’s method.
These sets of orbitals are listed in Table 1. Löwdin’s
method treats every orbital in the same pattern and there
which is Hermitian and positive definite (eigenvalues are
is only one outcome regardless of the number of FOs.
positive) since FOs are generally expected to be linearly
On the other hand, if the Gram–Schmidt orthonormali-
independent. Therefore, the overlap matrix can be diag-
sation is adopted, there are N! different ways to construct
onalised by a unitary matrix, Tjα :
orthonormalised orbitals. Regardless of what set of the
 Gram-Schmidt orbitals are chosen, there are always some
Tαi∗ Si j T jα = Qα . (6) orbitals that are very delocalised.
ij
MOLECULAR PHYSICS 5

Once the FLOs are determined for a given set of Table . Structural and energetic data of flat Dd Fe(II)-porphyrin
descriptors, the energy is minimised by varying these without constraining spin-up and spin-down electrons in the
descriptors. A set of Kohn–Sham orbitals is obtained same orbitals. This is the lowest energy of Ms =  electronic con-
figuration, so the relative energy is  eV (the first row). The super-
through standard DFT self-consistent field procedure and script on the left of symmetry label is the occupation number and
the positions of descriptors can be determined with the on the right is spin index of occupied electron. The Fe atom is at
methods for optimising molecular geometries. Minimisa- the centre.
tion of the energy in molecules by brute force adjustment E = . eV
of the FO descriptors leads to improved atomisation ener-
gies [2–4,30,34]. However, it was noted that with deriva- Coordinates of atoms (Å)
tives it would be possible to use methods such as conju- N (., ., .)
gate gradients to variationally determine the positions of C (., ., .)
C (., ., .)
FO descriptors. In order to do this, the derivatives of SIC C (., ., .)
energy with respect to the position of FO descriptors is H (., ., .)
H (., ., .)
needed:
Configurations of iron d orbitals (eV)
    
dE SIC dφk  SIC    dφk
= Vk φk + φk VkSIC  . d↓
z2
.
dam dam dam . d↑ /. d↑ .
k xz yz
       . d↓ − .
dφk  dφl  xy
. d↑ |
=
klk  φl − φk , (10) − . |
dam 
xy
dam . d↓ /. d↓
xz yz − .
kl
. d↑ − .
z2

with
klk =  φl |VkSIC |φk . Due to the fact that the localised
orbitals are constrained to lie in the space of the Kohn– Table . Structural and energetic data of flat Dd Fe(II)-porphyrin
Sham orbitals, this is a generally correct formula and with constraining spin-up and spin-down electrons in the same
there is no requirement that Kohn–Sham orbitals need to orbitals. The relative energy is . eV (the first row). The super-
script on the left of symmetry label is occupation number and on
be self-consistent solutions of a Hamiltonian. A complete
the right is spin index of occupied electron.
derivation of the FO derivatives may be found in [30,34].
The calculations reported here were performed with the E = . eV
NRLMOL suite of codes [39]. Coordinates of atoms (Å)

N (., ., .)


C (., ., .)
3. Results and discussion C (., ., .)
C (., ., .)
Fe(II)-porphyrin plays the key role for carrying oxygen in H (., ., .)
mammals and its geometry and energy differ as a function H (., ., .)
of spin states of the iron. The difficulty in determining Configurations of iron d orbitals (eV)
the geometries is due to the small difference in energy . d↑↓ .
and the detail of d-orbital filling. Here, only the MS x 2 −y2
. LUMO .
= 0 geometries and spin configurations are discussed, . d↑↓ /. d↑↓ .
xz yz
because the primary effort here is to demonstrate that this . d↑↓ − .
z2
version of SIC maybe be used on systems containing d . d↑↓ − .
xy
electrons.
In this work, the D4d symmetric flat geometry is cal-
culated within PBE-GGA to serve as the reference. Once spin-up and three spin-down d electrons associated with
the geometry is optimised, the symmetry is removed iron. Comparing the filling of d orbitals in Tables 2 and 3,
while the FLO-SIC calculations are performed. The full it is energetically less favoured when the electrons carry-
symmetry of the flat porphyrin can be generated by (x, y, ing both spins are constrained to be in the same d orbitals.
z) → (y, x, z), (x, y, z) → (x, y, −z), and (x, y, z) → (x, The energy of geometry in Table 3 is 0.54 eV higher than
−y, z). The inequivalent atomic positions for the relaxed the energy of the geometry in Table 2, which allows spin-
geometry and electron configuration for the low-energy up and spin-down electrons in different orbitals, because
structure is reported in Table 2 and the higher energy the Coulomb energy dominates the electron filling in d
structure is reported in Table 3. The major difference orbitals. The Kohn–Sham orbital density in one of the
between these two geometries is the filling of d orbitals. spin channels is shown in Figure 2, which corresponds
In the spin-unpolarised Fe(II)-porphyrin, there are three to the calculation in Table 3. The dxy orbital has the
6 D.-Y. KAO AND M. R. PEDERSON

Figure . Kohn–Sham orbital density of iron d-orbitals and occu-


pation numbers of spin-unpolarised calculation within PBE-GGA.

lowest energy among all the iron d-orbitals, because the


electronic lobes of dxy avoid the lobes of the nitrogens. A
0.01 eV gap between dz2 and dxz /dyz is found. The frac-
tional occupation numbers are observed for degenerate
orbitals dxz and dyz at Fermi level [7]. This is generally
Figure . Positions of FLO descriptors in the spin-unpolarised
an indicator of self-interaction error. dx2 −y2 is the orbital
Fe(II)-porphyrin are shown in as the smallest spheres. These are
with the highest energy since the electronic lobes of iron the descriptors of spin-up channel while those of the spin-down
collide with nitrogens and form anti-bonding orbitals. channel are identical. Iron is at the centre, nitrogens are light grey,
Since the positions of FO descriptors are transfer- carbons are dark grey, and hydrogens are in white.
able, the initial positions of the electrons associated
with the atoms on the ring are adopted from Mg(II)- configuration appears to be broken [3,30]. The carbon–
porphyrins [4]. Since many initial sets of FO descrip- carbon and carbon–hydrogen σ bonds and π bonds
tors are possible, identification of good guesses for the are reproduced by FO descriptors. From the side
initial positions of FO descriptors are screened through view, we observe that some of the FO descriptors can
consideration of the lowest eigenvalue of the intermedi- hover above/below the porphyrin plane. The hovering
ate Löwdin orbitals, the summation of the magnitude of behaviour indicates that those descriptors are flexible
the square of force, and the SIC energy. Generally speak- to move with very small changes in the total energy.
ing, several initial positions are chosen to compete for This appears to be a classical means for understanding
the geometry with the lowest energy, which is minimised the concept of aromaticity [4,30]. The resonance struc-
with respect to the positions of FO descriptors. The geo- tures observed in typical Kekule’ structures involve nine
metric mean of the intermediate Löwdin orbitals, the nat- delocalised π bonds (five carbon–carbon, two carbon–
ural log of the norm of FO derivatives, and the difference nitrogen, and two nitrogen–iron bonds). As discussed
of the SIC energy relative to the minimum are introduced in [40], our lowest energy geometry is in accord with
to determine the stages of the evolution of the positions this understanding. This paper shows that there are nine
of FO descriptors [34]. descriptors, referred to as ‘hovering descriptors’ [4,30],
The FLO-SIC calculation is performed within the that are displaced far above the plane of the molecule.
PW92-LDA approximation. The fractional occupied Four of the hovering descriptors are on top of the vertices
solutions are removed once the SIC is applied. The posi- of the tetrahedral coordination of nitrogens and they are
tions of FO descriptors are the red circles in Figure 3. similar to the ammonium ion (NH4+ ) tetrahedral config-
The FO descriptors near the plane imitate the classi- uration. In addition, there are always 24 FO descriptors
cal chemical bonds, and the symmetry of the electronic in close proximity to the iron centre with the other two
MOLECULAR PHYSICS 7

relocated to somewhere on the ring. This implies that [6] I. Lindgren, Int. J. Quant. Chem. 5(S5), 411–420 (1971).
the centred iron atom is +2 charged. These observations [7] J.F. Janak, Phys. Rev. B 18(12), 7165–7168 (1978).
imply that the behaviour of FO descriptors coincide with [8] J.P. Collman, J.L. Hoard, N. Kim, G. Lang, and C.A. Reed,
J. Am. Chem. Soc. 97, 2676–2681 (1975).
chemical intuition. [9] H. Goff, G.N. La Mar, and C.A. Reed, J. Am. Chem. Soc.
99, 3641–3646 (1977).
[10] M.S. Liao and S. Scheiner, J. Chem. Phys. 117, 205–219
4. Summary (2002).
In this paper, the FLO-based SIC is applied to the Ms = [11] H. Nakashima, J.-Y. Hasegawa, and H. Nakatsuji, J. Com-
put. Chem. 27(4), 426–433 (2006).
0 Fe(II)-porphyrin to demonstrate that this new theoret-
[12] C.A. Reed, T. Mashiko, W.R. Scheidt, K. Spartalian, and
ical method is viable for large molecules containing tran- G. Lang, J. Am. Chem. Soc. 102, 2302–2306 (1980).
sition metal systems. FOs depend explicitly on the den- [13] E. Sigfridsson and U. Ryde, J. Inorg. Biochem. 91, 101–
sity matrix and are therefore unitarily invariant highly 115 (2002).
localised orbitals. Use of Löwdin symmetric orthonor- [14] S. Paul, F. Amalraj, and S. Radhakrishnan, Synth. Met.
malisation preserves size-extensivity and the transferabil- 159(11), 1019–1023 (2009).
[15] C. She, S.J. Lee, J.E. McGarrah, J. Vura-Weis, M.R.
ity of descriptors. In this study, we show that the SIC is Wasielewski, H. Chen, G.C. Schatz, M.A. Ratner, and J.T.
able to lower the energy and resolve the fractional occu- Hupp, Chem. Commun. 46(4), 547–549 (2010).
pation. The similarity between FLO descriptors and the [16] D. Danilovic, C.L. Lin, T. Yuen, L. Pan, and J. Li, J. Appl.
classical chemical bonds is explored. In addition, aro- Phys. 101(9), 09E103 1–2 (2007).
maticity of Fe(II)-porphyrin ring is interpreted by the [17] E.I. Ioannidis and H.J. Kulik, J. Chem. Phys. 143(3),
034104 1–11 (2015).
hovering FLO descriptors.
[18] A. Svane, L. Petit, Z. Szotek, and W.M. Temmerman, Phys.
Rev. B 76(11), 115116 (2007).
[19] C. Edmiston and K. Ruedenberg, Rev. Mod. Phys. 35,
Acknowledgments 457–465 (1963).
Mark R. Pederson acknowledges the long-standing support [20] M.R. Pederson and C.C. Lin, J. Chem. Phys. 88(3), 1807–
from NRL which allowed development of NRLMOL. Der-you 1817 (1988).
Kao was supported by a fellowship at The George Washing- [21] M.R. Pederson, R.A. Heaton, and C.C. Lin, J. Chem. Phys.
ton University Institute of Nanotechnology. The authors thank 80(5), 1972–1975 (1984).
Drs Torsten Hahn, Jens Kortus, Tunna Baruah, Koblar Jackson, [22] M.R. Pederson, R.A. Heaton, and C.C. Lin, J. Chem. Phys.
and Mr Simon Liebing for many interesting discussions. 82(6), 2688–2699 (1985).
[23] J.P. Perdew, A. Ruzsinszky, J. Sun, and M.R. Pederson, in
Advances in Atomic, Molecular, and Optical Physics, edited
Disclosure statement by E. Arimondo C.C. Lin S.F. Yelin (Burlington Academic
Press, 2015), Vol. 64, Chap. 1, pp. 1–14.
No potential conflict of interest was reported by the authors. [24] K.A. Jackson, in Advances in Atomic, Molecular, and Opti-
cal Physics, edited by E. Arimondo C.C. Lin S.F. Yelin
Funding (Burlington Academic Press, 2015), Vol. 64, Chap. 2,
pp. 15–27.
This work was supported in part by Office of Naval Research [25] A. Pertsova, C.M. Canali, M.R. Pederson, I. Rungger, and
(ONR) [N00014-16-1-2464]; and The George Washington Uni- S. Sanvito, in Advances in Atomic, Molecular, and Opti-
versity Institute of Nanotechnology. cal Physics, edited by E. Arimondo C.C. Lin S.F. Yelin
(Burlington Academic Press, 2015), Vol. 64, Chap. 3,
pp. 29–86.
ORCID [26] P.M. Dinh, P. Reinhard, E. Suraud, and M. Vincendon, in
Advances in Atomic, Molecular, and Optical Physics, edited
Mark R. Pederson http://orcid.org/0000-0003-2489-4849 by E. Arimondo C.C. Lin S.F. Yelin (Burlington Academic
Press, 2015), Vol. 64, Chap. 4, pp. 87–103.
[27] N. Poilvert, G. Borghi, N.L. Nguyen, N.D. Keilbart,K.
References Wang, and I. Dabo, in Advances in Atomic, Molecular,
and Optical Physics, edited by E. Arimondo C.C. Lin S.F.
[1] J.P. Perdew and A. Zunger, Phys. Rev. B 23(10), 5048–
Yelin (Burlington Academic Press, 2015), Vol. 64, Chap.
5079 (1981).
5, pp. 105–127.
[2] M.R. Pederson, A. Ruzsinszky, and J.P. Perdew, J. Chem.
[28] N. Gidopoulos and N.N. Lathiotakis, in Advances in
Phys. 140(12), 121103 1–4 (2014).
Atomic, Molecular, and Optical Physics, edited by E. Ari-
[3] T. Hahn, S. Liebing, J. Kortus, and M.R. Pederson,
mondo C.C. Lin S.F. Yelin (Burlington Academic Press,
J. Chem. Phys. 143(22), 224104 1–7 (2015).
2015), Vol. 64, Chap. 6, pp. 129–142.
[4] M.R. Pederson, T. Baruah, D. Kao, and L. Basurto,
[29] S. Kümmel, in Advances in Atomic, Molecular, and Opti-
J. Chem. Phys. 144(16), 164117 1–8 (2016).
cal Physics, edited by E. Arimondo C.C. Lin S.F. Yelin
[5] J.P. Perdew and Y. Wang, Phys. Rev. B 45, 13244–13249
(Burlington Academic Press, 2015), Vol. 64, Chap. 7,
(1992).
pp. 143–151.
8 D.-Y. KAO AND M. R. PEDERSON

[30] M.R. Pederson and T. Baruah, in Advances in Atomic, [35] J. Sun, A. Ruzsinszky, and J.P. Perdew, Phys. Rev. Lett. 115,
Molecular, and Optical Physics, edited by E. Arimondo 036402 1–6 (2015).
C.C. Lin S.F. Yelin (Burlington Academic Press, 2015), [36] W. Luken and D.N. Beratan, Theor. Chim. Acta 61(3),
Vol. 64, Chap. 8, pp. 153–180. 265–281 (1982).
[31] S. Klüpfel, P. Klüpfel, and H. Jónsson, Phys. Rev. A. 84(5), [37] W. Luken and J.C. Culberson, Theor. Chim. Acta 66(5),
050501R 1–4 (2011). 279–283 (1984).
[32] S. Lehtola and H. Jónsson, J. Chem. Theory Comput. [38] B.C. Carlson and J.M. Keller, Phys. Rev. 105(1), 102–106
10(12), 5324–5337 (2014). (1957).
[33] M.R. Pederson and J.P. Perdew, K Newsletter Scientific [39] M.R. Pederson, D.V. Porezag, J. Kortus, and D.C. Patton,
Highlight of the Month (February, 2012). <http://www. Phys. Stat. Solidi B 217, 197–218 (2000).
psi-k.org/newsletters/News_109/Highlight_109.pdf >. [40] D. Kao, M.R. Pederson, T. Baruah, T. Hahn, S. Liebing,
[34] M.R. Pederson, J. Chem. Phys. 142(6), 064112 1–7 (2015). and J. Kortus (submitted).

You might also like