Ijms 23 00939

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

International Journal of

Molecular Sciences

Article
DFT Study of the Molecular and Electronic Structure of
Metal-Free Tetrabenzoporphyrin and Its Metal Complexes
with Zn, Cd, Al, Ga, In
Alexey V. Eroshin 1 , Arseniy A. Otlyotov 1,2 , Ilya A. Kuzmin 1 , Pavel A. Stuzhin 1 and Yuriy A. Zhabanov 1, *

1 Research Institute of Chemistry of Macroheterocyclic Compounds, Ivanovo State University of Chemistry


and Technology, 153000 Ivanovo, Russia; alexey.yeroshin@gmail.com (A.V.E.);
arseniy.otlyotov@chph.ras.ru (A.A.O.); wonderful_37@list.ru (I.A.K.); stuzhin@isuct.ru (P.A.S.)
2 N.N. Semenov Institute of Chemical Physics of Russian Academy of Sciences, Kosygina Street 4,
119991 Moscow, Russia
* Correspondence: zhabanov@isuct.ru; Tel.: +7-4932-35-98-74

Abstract: The electronic and molecular structures of metal-free tetrabenzoporphyrin (H2 TBP) and
its complexes with zinc, cadmium, aluminum, gallium and indium were investigated by density
functional theory (DFT) calculations with a def2-TZVP basis set. A geometrical structure of ZnTBP
and CdTBP was found to possess D4h symmetry; AlClTBP, GaClTBP and InClTBP were non-planar
complexes with C4v symmetry. The molecular structure of H2 TBP belonged to the point symmetry
group of D2h . According to the results of the natural bond orbital (NBO) analysis, the M-N bonds
had a substantial ionic character in the cases of the Zn(II) and Cd(II) complexes, with a noticeably
increased covalent contribution for Al(III), Ga(III) and In(III) complexes with an axial –Cl ligand.
 The lowest excited states were computed with the use of time-dependent density functional theory

(TDDFT) calculations. The model electronic absorption spectra indicated a weak influence of the
Citation: Eroshin, A.V.; Otlyotov,
nature of the metal on the Q-band position.
A.A.; Kuzmin, I.A.; Stuzhin, P.A.;
Zhabanov, Y.A. DFT Study of the
Keywords: tetrabenzoporphyrin; DFT study; molecular and electronic structure; chemical bonding
Molecular and Electronic Structure of
Metal-Free Tetrabenzoporphyrin and
Its Metal Complexes with Zn, Cd, Al,
Ga, In. Int. J. Mol. Sci. 2022, 23, 939.
https://doi.org/10.3390/ijms23020939
1. Introduction
Tetrapyrrole macroheterocycles, such as porphyrines, phthalocyanines and their analo-
gies and metal complexes, have found a number of applications [1–4]. The possibility of
Academic Editor: Alexande Baykov
fine-tuning their properties [5] by modification of the peripheral substituents [6] or atoms
Received: 28 December 2021 in the central ring allows us to use them as organic semiconductors, light-emitting diods [7],
Accepted: 11 January 2022 in photodynamic therapy [8,9], sensors of molecular oxygen [10–12] and in medicine,
Published: 15 January 2022 particularly theranostic [13–15].
Publisher’s Note: MDPI stays neutral
Tetrapyrrolic macrocycles can be used as photosensitizers in photodynamic therapy
with regard to jurisdictional claims in (PDT), and they often offer imaging capabilities [16–18].
published maps and institutional affil- The peripheral modification of porphyrin molecules allows the fine-tuning of their
iations. spectral luminescence properties, determining their fluorescence and photosensitizing
properties and the possibility of attaching peripheral moieties to target penetration in
tumor cells [18].
The intense absorption band (Q-band) of tetrabenzoporphyrin generally lies at
Copyright: © 2022 by the authors. 600–750 nm, meeting the requirements of an ideal photosensitizer (700–850 nm) [19]. The
Licensee MDPI, Basel, Switzerland. choice of the range is based on two main factors: the absorption and the scattering of light
This article is an open access article by tissue decrease as the wavelength increases. Moreover, if the absorption band is too far
distributed under the terms and in the red region, the oxidation potential and photobleaching will be decreased [20,21]. At
conditions of the Creative Commons
the same time, the central metal in porphyrins often determines the ratio of the competing
Attribution (CC BY) license (https://
fluorescence processes, and internal conversion processes, i.e., efficiency of the singlet
creativecommons.org/licenses/by/
oxygen generation. In this regard, along with Zn(II), complex macrocycles containing other
4.0/).

Int. J. Mol. Sci. 2022, 23, 939. https://doi.org/10.3390/ijms23020939 https://www.mdpi.com/journal/ijms


Int. J. Mol. Sci. 2022, 23, 939 2 of 12

heavy metal ions, such as Ga(III) and In(III), are very perspective as photosensitizers. In
addition, complexes of phthalocyanines are used as active layers in organic electronic de-
vices, e.g., In(III) phthalocyanine is used as an effective donor in photovolotaic cells [22,23].
The substitution of electronegative meso-nitrogens by meso-CH groups in TPB complexes
might be favorable for the enhancement of donor properties.
Complexes of benzo-fused porphyrins, which have more intense absorption in the
visible region than common porphyrines, remain much less studied [24,25]; moreover, no
information was reported on Ga and In tetrabenzoporphyrins.
Earlier in our laboratory, the properties of Ca(II), Zn(II) Y, La, Lu, metal-free tetra(1,2,5-
thiadiazole)porphyrazine [26–28], and other analogies of porphyrins, were investigated by
quantum chemical calculations.
In this work, we describe the influence of the molecular and electronic structures
on the properties of the series of tetrabenzoporphyrin complexes. Quantum chemical
calculations were carried out by means of the density functional theory (DFT) [29,30], since
our experience shows that it describes macrocyclic metal complexes fairly well [31–37]. The
Ahlrichs’-type def2-basis sets are commonly used for the metal-containing systems [38,39].
At the same time, triple-zeta quality basis sets are normally good enough and less computa-
tionally expensive, compared with their quadruple-zeta analogues. Therefore, a def2-TZVP
basis set was chosen for the calculations performed in the present work. The nature of
the chemical bonding between metal atoms and nitrogen atoms has been described using
the NBO-analysis of electron density distribution. The lowest excited states were also
calculated in order to explain the peculiarities and tendencies observed in the experimental
electronic absorption spectra available for the Zn, Cd and metal-free tetrabenzoporphyrins.

2. Results
2.1. Molecular Structure
The metal-free tetrabenzoporphyrin molecule H2 TBP possessed a D2h symmetry of
equilibrium configuration, according to quantum chemical calculations. All calculated
metal complexes had a fourth order symmetry axis: planar (D4h ) zinc (ZnTBP) and cad-
mium (CdTBP) complexes, whereas non-planar with doming-distorted porphyrin skeleton
(C4v ) complexes of aluminum, gallium and indium comprised a chlorine as an axial lig-
and (AlClTBP, GaClTBP и InClTBP). The optimized molecular
Int. J. Mol. Sci. 2022, 23, x FOR PEER REVIEW 3 of 13
structures are depicted in
Figure 1.

(a) (b)

(c)
Figure 1. Molecular models of metal-free tetrabenzoporphyrin (a), its complexes MTBP with Zn,
Figure
Cd 1. Molecular
(b), MClTBP with Al, Ga, models
In (c). of metal-free tetrabenzoporphyrin (a), its complexes MTBP with Zn, Cd
(b), MClTBP with Al, Ga, In (c).
The degree of the doming distortion for MClTBP can be described by the dihedral
angle α between planes of opposite pyrrole rings. Its values were 176°, 176° and 172° for
AlClTBP, GaClTBP and InClTBP, respectively. The main geometric parameters are
listed in Table 1.

Table 1. Molecular parameters 1 of H2-tetrabenzoporphyrin and its metal complexes optimized at


PBE0/def2-TZVP level.
Int. J. Mol. Sci. 2022, 23, 939 3 of 12

The degree of the doming distortion for MClTBP can be described by the dihedral
angle α between planes of opposite pyrrole rings. Its values were 176◦ , 176◦ and 172◦ for
AlClTBP, GaClTBP and InClTBP, respectively. The main geometric parameters are listed
in Table 1.

Table 1. Molecular parameters 1 of H2 -tetrabenzoporphyrin and its metal complexes optimized at


PBE0/def2-TZVP level.

H2 TBP ZnTBP CdTBP AlClTBP GaClTBP InClTBP


Symmetry D2h D4h D4h C4v C4v C4v
Distances
M-N 1.012 (2.340) 2 2.063 2.152 2.044 2.075 2.184
M-Cl - - - 2.154 2.196 2.360
N . . . Nopp 4.268 (4.106) 4.125 4.304 4.006 4.060 4.195
N . . . Nadj 2.961 2.917 3.043 2.833 2.871 2.966
N-Cα 1.362 (1.353) 1.363 1.355 1.370 1.366 1.361
Cα -Cβ 1.439 (1.457) 1.446 1.453 1.440 1.441 1.446
Cβ -Cβ 1.407 (1.399) 1.401 1.409 1.395 1.397 1.404
Cβ -Cγ 1.394 (1.389) 1.393 1.391 1.394 1.393 1.392
Cγ-Cδ 1.379 (1.385) 1.381 1.383 1.380 1.380 1.381
Cδ -Cδ 1.404 (1.398) 1.402 1.399 1.403 1.402 1.401
Cα -Cm 1.379 (1.390) 1.383 1.398 1.373 1.376 1.387
Cα -Cm -Cα 128.0 127.4 130.3 125.3 126.0 128.3
r(M-X) 3 - 0 0 0.407 0.453 0.678
Bond angles
N-Cα -Cm 126.2 (125.9) 125.6 125.6 125.6 125.8 125.8
N-Cα -Cβ 106.3 (110.7) 109.5 107.6 110.6 109.9 108.6
A4 180.0 180.0 180.0 176.3 176.4 171.6
1 All internuclear distances are in Angstroms (Å), valence angle are in degrees (◦ ). 2 The values in parentheses
correspond to the corresponding values for the isoindolenine fragments of the metal-free tetrabenzoporphyrin.
3 X is a dummy atom located in the center between N atoms. 4 α is the dihedral angle between planes of opposite

pyrrole rings.

The metal out-of-plane distance, defined as the distance between a metal atom and
dummy-atom placed in the center of the square of nitrogen atoms, increased from Al to In,
which corresponded to an increase in its ionic radii [40]. The influence of the metal nature
on the structure of the macrocyclic framework was minor, as in the works [26–28]. Indeed,
internuclear distances were predominantly intermediate between the corresponding val-
ues for the isoindole and isoindolenine fragments of the metal-free tetrabenzoporphyrin.
However, the cadmium complex broke out of this trend. There was a significant shortening
of the N-Cα distance and the Cα -Cm distance elongation) in CdTBP.
The M-N distance and Cα -Cm -Cα valence angle correlated with the ionic radii of
metals [40]. The tetrabenzoporphyrines structure resembled both porphyrins and phthalo-
cyanines. It is not surprising that the M-N distance was closer to that in porphyrines [41],
rather than in phthalocyanines [42,43], since for the latter the replacement of carbon atoms
with nitrogen atoms into meso-positions leads to a decrease in the size of the macrocycle
cavity. Our previous studies show that the introduction of peripheral substituents does not
significantly affect the structure of the macrocyclic ligand, whereas a change of the metal
nature can lead to substantial changes in the inner macrocycle ring internuclear distances.

2.2. NBO-Analysis
According to the results of the NBO analysis, the complexes of TBP ligand with
Zn(II) and Cd(II) were stabilized by strong interactions of the types: LP(N) → ns(M) and
LP(N) → np(M) (Figure 2; n = 4 for Zn, and n = 5 for Cd). In the case of the MClTBP
complexes (M = Al, Ga, In) the out-of-plane position of a metal atom led to an additional
favorable overlap LP(N) → npz (M) (Figure 3, c; n = 3 for Al, n = 4 for Ga and n = 5 for In).
2.2. NBO-Analysis
According to the results of the NBO analysis, the complexes of TBP ligand with
Zn(II) and Cd(II) were stabilized by strong interactions of the types: LP(N) → ns(M) and
LP(N) → np(M) (Figure 2; n = 4 for Zn, and n = 5 for Cd). In the case of the MClTBP
Int. J. Mol. Sci. 2022, 23, 939 4 of 12
complexes (M = Al, Ga, In) the out-of-plane position of a metal atom led to an additional
favorable overlap LP(N) → npz(M) (Figure 3, c; n = 3 for Al, n = 4 for Ga and n = 5 for In).

(a) (b)
Figure 2.
Figure 2. Schemes
Schemesofofthe dominant
the dominant donor–acceptor interactions
donor–acceptor between
interactions Zn and
between the
Zn TBPthe
and ligand.
TBP (a)ligand.
The result of the orbital interaction of the type LP(N) → 4s(Zn) (E(2) = 42.6 kcal mol−1); (b) the−result
1 ); (b) the
(a) The result of the orbital interaction of the type LP(N) → 4s(Zn) (E(2) = 42.6 kcal mol
of the orbital interaction of the type LP(N) → 4p(Zn) (E(2) = 35.5 kcal mol−1). Only one
Int. J. Mol. Sci. 2022, 23, x FOR PEER REVIEW of the four 5 of 13
result of the orbital
corresponding interaction
interactions of the type LP(N) → 4p(Zn) (E(2) = 35.5 kcal mol−1 ). Only one of the
is demonstrated.
four corresponding interactions is demonstrated.

(a) (b)

(c)
Figure 3.
Figure 3. Schemes
Schemesof ofthe
thedominant
dominantdonor–acceptor
donor–acceptorinteractions between
interactions between thethe
Al Al
andand
TBPTBPligand. (a)
ligand.
The result of the orbital interaction of the type LP(N) → 3s(Al) (E(2) = 40.7 kcal mol−1); (b) the result
− 1
(a) The result of the orbital interaction of the type LP(N) → 3s(Al) (E(2) = 40.7 kcal mol ); (b) the
of the orbital interaction of the type LP(N) → 3py(Al) (E(2) = 60.0 kcal mol−1); (c) the result of the
result of the orbital interaction of the type LP(N) → 3py (Al) (E(2) =−160.0 kcal mol−1 ); (c) the result
orbital interaction of the type LP(N) → 3pz(Al) (E(2) = 12.2 kcal mol ). Only one of the four corre-
of the orbital interaction of the type LP(N) → 3pz (Al) (E(2) = 12.2 kcal mol−1 ). Only one of the four
sponding interactions is demonstrated.
corresponding interactions is demonstrated.
It should
It shouldbebenoted
notedthat
thatthethecovalent
covalent contribution
contribution into
into thethe bonds
bonds M–NM–N depends
depends on theon
the oxidation state of the central metal, and is higher for
III M IIICl complexes than for com-
oxidation state of the central metal, and is higher for M Cl complexes than for complexes
plexes
of of bivalent
bivalent ions According
ions According to theofvalues
to the values of thecharges
the natural naturalq(M) charges
andq(M) and theof
the energies ener-
the
gies of the donor–acceptor
donor–acceptor interactions (∑ interactions (∑ E(d-a))
E(d-a)) between between
lone pairs on the lone pairs atoms
nitrogen on theand
nitrogen
s- and
atoms andofs-the
p-orbitals and p-orbitals
metal atoms, of thethe metal
bond Cd–N atoms,
was the bondmore
slightly Cd–N was
ionic slightly with
compared moreZn–Nionic
compared with Zn–N (Table 2). The trend is different for the complexes
(Table 2). The trend is different for the complexes MClTBP: the covalent contribution to the MClTBP: the
covalent contribution to the bond M–N increased within the series
bond M–N increased within the series Al → Ga → In. The enhanced covalent properties ofAl → Ga → In. The
enhanced covalent properties of Ga–N bond correlated with the well-known [44,45] al-
ternation in the electronegativities of Al (χ = 1.47), Ga (χ = 1.82) and In (χ = 1.49).

Table 2. Selected parameters of MTBP complexes from NBO calculations.


Int. J. Mol. Sci. 2022, 23, 939 5 of 12

Ga–N bond correlated with the well-known [44,45] alternation in the electronegativities of
Al (χ = 1.47), Ga (χ = 1.82) and In (χ = 1.49).

Table 2. Selected parameters of MTBP complexes from NBO calculations.

ZnTBP CdTBP AlClTBP GaClTBP InClTBP


E(HOMO), eV −5.06 −5.13 −5.19 −5.22 −5.30
E(LUMO), eV −2.34 −2.38 −2.54 −2.55 −2.62
∆E, eV 2.72 2.75 2.65 2.67 2.68
q(M) NPA, e 1.304 1.333 1.776 1.673 1.718
q(N) NPA, e −0.573 −0.570 −0.607 −0.584 −0.579
q(Cl) NPA, e −0.574 −0.545 −0.563
configuration 4s0.35 3d9.97 4p0.37 5s0.41 5d9.95 5p0.31 3s0.42 3p0.76 4s0.55 4p0.76 5s0.54 5p0.73
∑ E(d-a),
312.2 303.8 451.7 495.6 471.9
kcal mol−1
Q(M–N) 0.283 0.274 0.320 0.335 0.330
r(M–N), Å 2.063 2.152 2.044 2.075 2.184
Int. J. Mol. Sci. 2022, 23, x FOR PEER REVIEW 6 of 13
2.3. Electronic Spectra
The calculated spectra of the investigated compounds were quite similar, which
In the spectrum
demonstrates theofinsignificant
H2TBP, the influence
splitting of of the Q-band
the metal into two
nature bands
to the was of
position observed.
the Q-band.
Q-band
In splitting occurred
the spectrum of H2 TBP, because of the double-degenerated
the splitting of the Q-band intoLUMOs of eg*was
two bands symmetry of Q-
observed.
MTBP,
band which split
splitting into two
occurred b1g* and
because b2g* double-degenerated
of the orbitals of H2TBP, since the latter
LUMOs of ehad a lower of
g * symmetry
symmetry
MTBP, compared
which with
split into twotheb1g
metal
* andcomplexes. A negligible
b2g * orbitals of H2 TBP,batochromic shifthad
since the latter for athe
lower
complexes with
symmetry axial ligands,
compared with the and a hypsochromic
metal complexes. shift for Cd and
A negligible Zn occurredshift
batochromic withfora the
metal being with
complexes introduced into the H
axial ligands, 2TBP
and molecule (Figure
a hypsochromic 4). for Cd and Zn occurred with a
shift
metal being introduced into the H2 TBP molecule (Figure 4).

InClTBP

GaClTBP

AlClTBP

CdTBP

ZnTBP

H2TBP

250 300 350 400 450 500 550 600 650 700
λ, nm

Figure 4. Calculated TDDFT


4. Calculated TDDFTelectronic
electronic absorption
absorptionspectra
spectrafor
forH
H22TBP,
TBP,MTBP andMClTBP
MTBPand complexes.
MClTBPcom-
plexes.
The calculated oscillator strengths (f) for the lowest-allowed excited states, along with
their The
composition
calculated(inoscillator
terms of strengths
one-electron transition)
(f) for are given in excited
the lowest-allowed Table 3. states,
A full version
along of
the
withtable
theiriscomposition
listed in Table(in S1.
terms of one-electron transition) are given in Table 3. A full
Theoflong-wave
version the table isabsorption maxima
listed in Table S1. (Q band) in the spectra of MTBP can be assigned
to theThe
transitions
long-wavefrom HOMOmaxima
absorption and HOMO-1
(Q band)to inthe
the double-degenerated
spectra of MTBP can be LUMO. This is
assigned
to the transitions
typical from HOMO such
for macroheterocycles and HOMO-1 to the double-degenerated
as porphyrazines with Ca, Zn [27] LUMO. This is
and -tetrakis(1,2,5-
typical for macroheterocycles
thiadiazole)porphyzarines with such
Y, La, asLu porphyrazines
[28]. The shift with of theCa, Zn [27]maximum
long-wave and
-tetrakis(1,2,5-thiadiazole)porphyzarines
towards higher values in the case of MClTBP, with Y, La, Lu [28]. The
compared withshift of the
MTBP, waslong-wave
quite small
maximum
(ca. 20 nm) towards
and might higher values ineither
be attributed the case of different
to the MClTBP,nature
compared with MTBP,
of a central was or
metal atom,
quite small (ca. 20 nm) and might be attributed either to the different nature of
to the presence of the axial chlorine substituent. The most intensive peak in the spectra, the a central
metal band,
Soret atom, predominantly
or to the presence of the axial chlorine
corresponded substituent.
to the electron The most
transitions intensive
from peak a1
the occupied
in the spectra, the Soret band, predominantly corresponded to the electron transitions
from the occupied a1 (Al, Ga, In complexes) a2u (Zn, Cd complexes)-type MOs to the
LUMOs. The composition of the Q- and Soret-bands were quite similar for all complexes,
and can be described by the model of Gouterman [46,47].
Int. J. Mol. Sci. 2022, 23, 939 6 of 12

(Al, Ga, In complexes) a2u (Zn, Cd complexes)-type MOs to the LUMOs. The composition
of the Q- and Soret-bands were quite similar for all complexes, and can be described by the
model of Gouterman [46,47].

Table 3. Calculated composition of the lowest excited states and corresponding oscillator strengths
for H2 TBP and MTBP complexes.

State Composition (%) λ, nm f exp λ, nm


H2 TBP
∗ (31)
2b3u → 1b2g
11 B1u ∗ (69) 578 0.11 663.5 (Py) [48]
3au → 1b1g
∗ (19)
2b3u → 1b1g
11 B2u ∗ 568 0.23
3au → 1b2g (81)
2b3u → 1b2g∗ (8)

21 B1u ∗
2b3u → 1b2g (59) 387 1.24 431.8 (Py) [48]
∗ (28)
3au → 1b1g
∗ (74)
2b3u → 1b1g
21 B2u ∗ (18) 372 1.10 416.1 (Py) [48]
3au → 1b2g
∗ (88)
2b3u → 1b2g
31 B1u ∗ (6) 332 0.35
2b3u → 1b2g
31 B2u ∗
3au → 2b2g (93) 331 0.15
ZnTBP
2a2u → 1e∗g (21) 613(Ar matrix) [49]
11 Eu 563 0.19
2a1u → 1e∗g (79) 628.5 (Py) [48]
2a2u → 1e∗g (73)
21 Eu 372 1.17 433.3(Py) [48]
2a1u → 1e∗g (20)
31 Eu 2a1u → 2e∗g (94) 326 0.13
CdTBP
2a2u → 1e∗g (23)
11 Eu 560 0.18 628 (Py) [50]
2a1u → 1e∗g (77)
2a2u → 1e∗g (71)
21 Eu 376 1.21 434 (Py) [50]
2a1u → 1e∗g (22)
1b2u → 1e∗g (18)
31 Eu 320 0.08
2a1u → 2e∗g (78)
AlClTBP
2a1 → 1e∗ (20)
11 E 580 0.18
2a2 → 1e∗ (80)
2a1 → 1e∗ (73)
21 E 382 1.01
2a2 → 1e∗ (19)
31 E 2a2 → 2e∗ (93) 329 0.14
GaClTBP
3a1 → 1e∗ (21)
11 E 577 0.17
2a2 → 1e∗ (79)
3a1 → 1e∗ (73)
21 E 385 0.99
2a2 → 1e∗ (20)
2b1 → 1e∗ (5)
31 E 326 0.10
2a2 → 2e∗ (93)
InClTBP
2a1 → 1e∗ (22)
11 E 577 0.17
2a2 → 1e∗ (77)
2a1 → 1e∗ (72)
21 E 388 1.03
2a2 → 1e∗ (22)
1b1 → 1e∗ (23)
31 E 322 0.07
2a2 → 2e∗ (73)
Int. J. Mol. Sci. 2022, 23, 939 7 of 12

Shapes of the highest occupied molecular orbitals (HOMO), HOMO-1 and lowest
unoccupied
Int. J. Mol. Sci. 2022, 23, x FOR PEER REVIEW
molecular orbitals (LUMO) were similar (Figure 5) for all the8 of investigated
13
complexes. HOMO is an au (H2 ), a1u (Zn, Cd) or a2 (Al, Ga, In), LUMOs are doubly-
degenerated e* or eg * orbitals. The symmetry of these orbitals is typical for porphyrines
and porphyrazines [27,31,51,52].

H2TBP ZnTBP CdTBP

LUMO+1

b1g* b2u* b1u*

LUMO

b2g* eg * eg *

HOMO

au a1u a1u

HOMO-1

b3u a2u a2u


AlClTBP GaClTBP InClTBP

LUMO+1

b1 * b1 * b1*

LUMO

e* e* e*

HOMO

a2 a2 a2

HOMO-1

a1 a1 a1

Figure 5. Shapes of the frontier molecular orbitals.


Int. J. Mol. Sci. 2022, 23, x FOR PEER REVIEW 9 of 13

Int. J. Mol. Sci. 2022, 23, 939 8 of 12


Figure 5. Shapes of the frontier molecular orbitals.

The HOMO predominantly represented the linear combination of atomic orbitals


(AOs) The HOMO
of the pyrrolepredominantly
rings, whereasrepresented
the HOMO-1 thewas
linear combination
localized of atomic
on the carbon orbitals
atoms in
(AOs) of the pyrrole rings, whereas the HOMO-1 was localized on the
the meso-positions. Furthermore, HOMO-1 and HOMO-2 orbitals contained both AOs carbon atoms inof
the
meso-positions. Furthermore, HOMO-1 and HOMO-2 orbitals contained both
the macrocycle and axial ligand Cl for Al, Ga and In complexes. The contribution of the AOs of the
macrocycle
metal and axial
to this orbital wasligand
slight.Cl for Al, Ga and In complexes. The contribution of the metal
to this orbital was slight.
Despite the fact that the HOMO–LUMO gap was the lowest for metal-free H2TBP
Despite the fact that the HOMO–LUMO gap was the lowest for metal-free H2 TBP
(Figure 6), the wavelength of its Q-band was not the largest among the molecules con-
(Figure 6), the wavelength of its Q-band was not the largest among the molecules consider-
sidering both the transitions from HOMO to LUMO, and from HOMO-1 to LUMO, con-
ing both the transitions from HOMO to LUMO, and from HOMO-1 to LUMO, contributed
tributed to Q-band, according to the quantum chemical calculations performed.
to Q-band, according to the quantum chemical calculations performed.

H2TBP ZnTBP CdTBP AlClTBP GaClTBP InClTBP


0
b1g* a2u*
b3u* eg* eg* e*
b2g* a1*
-1 au* b1u* b2*
b3u* b2u* b1*

-2 eg*
b1g*
e*
b2g*
-3
E, eV

-4 2.64 2.73 2.75 2.65 2.67 2.68

au a1u
-5 a2
b3u a2u
a1
-6
b1g
b2u eg b2u
b1g eg e
-7 eg b1
b3u
b1g e
b2g a2u a1
b1u
b2

Figure 6. Molecular orbital (MO) level diagram for H2 TBP and MTBP complexes. The values of
Figure 6. occupied
higher Molecularmolecular
orbital (MO) level diagram
orbital–lowest for H2TBP
unoccupied and MTBP
molecular complexes.
orbital The values gaps
(HOMO–LUMO) of are
higher
givenoccupied
in eV. molecular orbital–lowest unoccupied molecular orbital (HOMO–LUMO) gaps are
given in eV.
3. Computational Details
3. Computational Details calculations of the tetrabenzoporphyrin and its metal complexes
Quantum chemical
with Zn, Cd, Al,
Quantum chemicalGa, and In were performed
calculations using the Gaussian09
of the tetrabenzoporphyrin and its[53] program.
metal complexesPBE0
with Zn, Cd, Al, Ga, and In were performed using the Gaussian09 [53] program. PBE0D3
exchange-correlation (XC) functional with the density functional dispersion correction
provided by Grimme
exchange-correlation [54],functional
(XC) in combination withdensity
with the the def2-TZVP basis
functional set [55] for
dispersion all atoms
correction
D3taken from the
provided by EMSL
Grimme BSE library
[54], [56–58], waswith
in combination applied
the for the structure
def2-TZVP basisoptimization and
set [55] for all
computation
atoms of harmonic
taken from the EMSL vibrations.
BSE library Analytic
[56–58], Hessian calculations
was applied indicated
for the structuretheoptimiza-
absence of
theand
tion imaginary vibrational
computation frequencies
of harmonic and, therefore,
vibrations. Analytic the optimized
Hessian structuresindicated
calculations corresponded
the
to the minima on the PES. The optimized Cartesian coordinates
absence of the imaginary vibrational frequencies and, therefore, the optimized of H 2 TBP and its metal
structures
complexes with
corresponded Zn,minima
to the Cd, Al, on Ga,the
andPES.In are
Theavailable
optimized in the Supplementary
Cartesian coordinatesMaterials.
of H2TBP
For describing the core electron shells of the cadmium and indium
and its metal complexes with Zn, Cd, Al, Ga, and In are available in the Supplementary atoms, pseudopo-
tentials combined with a corresponding basis set were used. The doubly occupied orbitals
Materials.
corresponding to the
For describing the1s, 2s, 2p,
core 3s, 3pshells
electron and 3doforbitals were described
the cadmium by multiconfiguration-
and indium atoms, pseudo-
Dirac–Hartree–Fock-adjusted pseudopotentials [59,60].
potentials combined with a corresponding basis set were used. The doubly occupied or-
Gaussian03 [61]
bitals corresponding towas
the employed for 3p
1s, 2s, 2p, 3s, theandNBO-analysis
3d orbitals of electron
were density
described by distribution.
multicon-
TDDFT calculations of the electronic absorption
figuration-Dirac–Hartree–Fock-adjusted pseudopotentials [59,60]. spectra were performed with the use of
theGaussian03
Firefly QC package [62], which is partially based on the GAMESS (US)
[61] was employed for the NBO-analysis of electron density distribution. [63] source code,
since it supports separate computations of the electronic transitions for each irreducible
TDDFT calculations of the electronic absorption spectra were performed with the use of
representation which, in turn, results in the proper automatic determination of the wave
function symmetry. The number of the calculated excited states was 30.
Int. J. Mol. Sci. 2022, 23, 939 9 of 12

The molecular models and orbitals demonstrated in the paper were visualized by
means of the Chemcraft program [64].

4. Conclusions
The influence of the nature of the metal (M = Zn, Cd, Al(Cl), Ga(Cl), In(Cl)) on the
molecular and electronic structure of the tetrabenzoporphyrin molecule H2 TBP was studied
using the DFT method (PBE0 functional) with a def2-TZVP basis set. A weak influence of
the metal nature on the structure of the macrocyclic framework was observed, whereas the
dimensions of the coordination cavity of the macrocycle increased in proportion to the ionic
radii of metals. The electron density distribution was considered in terms of the natural
bond orbitals (NBO). The low values of the Wyberg bond indices indicated that the M-N
bonds had an ionic character with a noticeable covalent contribution, which depended on
the oxidation state of the metal. The complexes were stabilized by strong interactions of
the types: LP(N) → ns(M), LP(N) → np(M). For non-planar complexes with Al, Ga, and
In, an additional favorable overlap LP(N) → npz (M) appeared, where n was a principal
quantum number.
The Q-band position was weakly dependent on the nature of the metal. The fron-
tier orbitals were mainly localized on the atoms constituting the internal 16-membered
macrocycle. HOMO and LUMOs were found to be ordinary Gouterman-type orbitals.
The HOMO–LUMO gap was the lowest for metal-free H2 TBP; however, an insignificant
batochromic shift of the Q-band for the complexes with axial ligand, and a hypsochromic
shift of the Q-band for Cd and Zn occurred, compared with the metal-free H2 TBP molecule,
since both the transitions from HOMO to LUMO and from HOMO-1 to LUMO contributed
to the Q-band.

Supplementary Materials: The following supporting information can be downloaded at: https:
//www.mdpi.com/article/10.3390/ijms23020939/s1.
Author Contributions: Conceptualization, P.A.S. and Y.A.Z.; methodology, P.A.S. and Y.A.Z.; inves-
tigation, I.A.K., A.V.E. and A.A.O.; resources, Y.A.Z.; data curation, A.V.E.; writing—original draft
preparation, A.V.E., A.A.O. and I.A.K. All authors have read and agreed to the published version of
the manuscript.
Funding: This work is supported by the Russian Science Foundation (grant № 21-73-10126).
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: The data presented in this study are available on request from the
corresponding author.
Acknowledgments: The research was carried out using the resources of the Center for Shared Use of
Scientific Equipment of the ISUCT (with the support of the Ministry of Science and Higher Education
of Russia, grant No. 075-15-2021-671).
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Sorokin, A.B. Recent progress on exploring µ-oxo bridged binuclear porphyrinoid complexes in catalysis and material science.
Coord. Chem. Rev. 2019, 389, 141–160. [CrossRef]
2. Yu, Z.; Hagfeldt, A.; Sun, L. The application of transition metal complexes in hole-transporting layers for perovskite solar cells:
Recent progress and future perspectives. Coord. Chem. Rev. 2020, 406, 213143. [CrossRef]
3. Almeida-Marrero, V.; Van De Winckel, E.; Anaya-Plaza, E.; Torres, T.; De La Escosura, A. Porphyrinoid biohybrid materials as an
emerging toolbox for biomedical light management. Chem. Soc. Rev. 2018, 47, 7369–7400. [CrossRef]
4. Koifman, O.I.; Ageeva, T.A.; Beletskaya, I.P.; Averin, A.D.; Yakushev, A.A.; Tomilova, L.G.; Dubinina, T.V.; Tsivadze, A.Y.;
Gorbunova, Y.G.; Martynov, A.G.; et al. Macroheterocyclic compounds-a key building block in new functional materials and
molecular devices. Macroheterocycles 2020, 13, 311–467. [CrossRef]
5. Longevial, J.; Clément, S.; Wytko, J.A.; Ruppert, R.; Weiss, J.; Richeter, S. Peripherally Metalated Porphyrins with Applications in
Catalysis, Molecular Electronics and Biomedicine. Chem.-A Eur. J. 2018, 24, 15442–15460. [CrossRef]
Int. J. Mol. Sci. 2022, 23, 939 10 of 12

6. Zhou, Q.; Xu, S.; Yang, C.; Zhang, B.; Li, Z.; Deng, K. Modulation of peripheral substituents of cobalt thioporphyrazines and their
photocatalytic activity. Appl. Catal. B Environ. 2016, 192, 108–115. [CrossRef]
7. Borek, C.; Hanson, K.; Djurovich, P.I.; Thompson, M.E.; Aznavour, K.; Bau, R.; Sun, Y.; Forrest, S.R.; Brooks, J.; Michalski, L.; et al.
Highly efficient, near-infrared electrophosphorescence from a Pt-metalloporphyrin complex. Angew. Chem.-Int. Ed. 2007,
46, 1109–1112. [CrossRef] [PubMed]
8. Srivatsan, A.; Missert, J.R.; Upadhyay, S.K.; Pandey, R.K. Porphyrin-based photosensitizers and the corresponding multifunctional
nanoplatforms for cancer-imaging and phototherapy. J. Porphyr. Phthalocyanines 2015, 19, 109–134. [CrossRef]
9. Abrahamse, H.; Hamblin, M.R. New photosensitizers for photodynamic therapy. Biochem. J. 2016, 473, 347–364. [CrossRef]
[PubMed]
10. Vinogradov, S.A.; Wilson, D.F. Porphyrin dendrimers as biological oxygen sensors. In Designing Dendrimers; John Wiley & Sons:
New York, NY, USA, 2012; pp. 465–503. ISBN 978-0-470-43355-3.
11. Quaranta, M.; Borisov, S.M.; Klimant, I. Indicators for optical oxygen sensors. Bioanal. Rev. 2012, 4, 115–157. [CrossRef] [PubMed]
12. Papkovsky, D.B.; Dmitriev, R.I. Biological detection by optical oxygen sensing. Chem. Soc. Rev. 2013, 42, 8700–8732. [CrossRef]
[PubMed]
13. Agostinis, P.; Berg, K.; Cengel, K.A.; Foster, T.H.; Girotti, A.W.; Gollnick, S.O.; Hahn, S.M.; Hamblin, M.R.; Juzeniene, A.;
Kessel, D.; et al. Photodynamic therapy of cancer: An update. CA Cancer J. Clin. 2011, 61, 250–281. [CrossRef] [PubMed]
14. Kelkar, S.S.; Reineke, T.M. Theranostics: Combining imaging and therapy. Bioconjug. Chem. 2011, 22, 1879–1903. [CrossRef]
15. Juweid, M.E.; Mottaghy, F.M. Current and future aspects of nuclear molecular therapies: A model of theranostics. Methods 2011,
55, 193–195. [CrossRef]
16. Scherer, R.L.; McIntyre, J.O.; Matrisian, L.M. Imaging matrix metalloproteinases in cancer. Cancer Metastasis Rev. 2008, 27, 679–690.
[CrossRef]
17. Danhier, F.; Ansorena, E.; Silva, J.M.; Coco, R.; Le Breton, A.; Préat, V. PLGA-based nanoparticles: An overview of biomedical
applications. J. Control. Release 2012, 161, 505–522. [CrossRef] [PubMed]
18. Van De Wiele, C.; Oltenfreiter, R. Imaging probes targeting matrix metalloproteinases. Cancer Biother. Radiopharm. 2006,
21, 409–417. [CrossRef]
19. O’Connor, A.E.; Gallagher, W.M.; Byrne, A.T. Porphyrin and nonporphyrin photosensitizers in oncology: Preclinical and clinical
advances in photodynamic therapy. Photochem. Photobiol. 2009, 85, 1053–1074. [CrossRef]
20. Bonnett, R. Photosensitizers of the porphyrin and phthalocyanine series for photodynamic therapy. Chem. Soc. Rev. 1995,
24, 19–33. [CrossRef]
21. Balaz, M.; Collins, H.A.; Dahlstedt, E.; Anderson, H.L. Synthesis of hydrophilic conjugated porphyrin dimers for one-photon and
two-photon photodynamic therapy at NIR wavelengths. Org. Biomol. Chem. 2009, 7, 874–888. [CrossRef] [PubMed]
22. Williams, G.; Sutty, S.; Klenkler, R.; Aziz, H. Renewed interest in metal phthalocyanine donors for small molecule organic solar
cells. Sol. Energy Mater. Sol. Cells 2014, 124, 217–226. [CrossRef]
23. Travkin, V.V.; Stuzhin, P.A.; Okhapkin, A.I.; Korolyov, S.A.; Pakhomov, G.L. Organic tandem Schottky junction cells with high
open circuit voltage. Synth. Met. 2016, 212, 51–54. [CrossRef]
24. Kadish, K.M.; Smith, K.M.; Guilard, R. The Porphyrin Handbook: Inorganic, Organometallic and Coordination Chemistry; Elsevier:
Amsterdam, The Netherlands, 2003; Volume 3, p. 425.
25. Milgrom, L.R. The Colours of Life: An Introduction to the Chemistry of Porphyrins and Related Compounds; Oxford University Press:
New York, NY, USA, 1997.
26. Otlyotov, A.A.; Ryzhov, I.V.; Kuzmin, I.A.; Zhabanov, Y.A.; Mikhailov, M.S.; Stuzhin, P.A. Peculiarities of electronic structure
and chemical bonding in iron and cobalt metal complexes of porphyrazine and tetra(1,2,5-thiadiazole)porphyrazine. J. Porphyr.
Phthalocyanines 2020, 24, 1146–1154. [CrossRef]
27. Otlyotov, A.A.; Ryzhov, I.V.; Kuzmin, I.A.; Zhabanov, Y.A.; Mikhailov, M.S.; Stuzhin, P.A. Dft study of molecular and electronic
structure of ca(Ii) and zn(ii) complexes with porphyrazine and tetrakis(1,2,5-thiadiazole)porphyrazine. Int. J. Mol. Sci. 2020,
21, 2923. [CrossRef]
28. Zhabanov, Y.A.; Ryzhov, I.V.; Kuzmin, I.A.; Eroshin, A.V.; Stuzhin, P.A. DFT Study of Molecular and Electronic Structure of Y, La
and Lu Complexes with Porphyrazine and Tetrakis(1,2,5-thiadiazole)porphyrazine. Molecules 2020, 26, 113. [CrossRef] [PubMed]
29. Hohenberg, P.; Kohn, W. Inhomogeneous electron gas. Phys. Rev. 1964, 136, B864. [CrossRef]
30. Kohn, W.; Sham, L.J. Self-consistent equations including exchange and correlation effects. Phys. Rev. 1965, 140, A1133. [CrossRef]
31. Zhabanov, Y.A.; Tverdova, N.V.; Giricheva, N.I.; Girichev, G.V.; Stuzhin, P.A. DFT Study of molecular and electronic structure
of magnesium (II) tetra(1,2,5-chalcogenadiazolo) porphyrazines, [TXDPzMg] (X = O, S, Se, Te). J. Porphyr. Phthalocyanines 2017,
21, 439–452. [CrossRef]
32. Lebedeva (Yablokova), I.A.; Ivanova, S.S.; Novakova, V.; Zhabanov, Y.A.; Stuzhin, P.A. Perfluorinated porphyrazines. 3. Synthesis,
spectral-luminescence and electrochemical properties of perfluorinated octaphenylporphyrazinatozinc(II). J. Fluor. Chem. 2018,
214, 86–93. [CrossRef]
33. Stuzhin, P.A.; Skvortsov, I.A.; Zhabanov, Y.A.; Somov, N.V.; Razgonyaev, O.V.; Nikitin, I.A.; Koifman, O.I. Subphthalocyanine
azaanalogues—Boron(III) subporphyrazines with fused pyrazine fragments. Dye. Pigment. 2019, 162, 888–897. [CrossRef]
Int. J. Mol. Sci. 2022, 23, 939 11 of 12

34. Otlyotov, A.A.; Zhabanov, Y.A.; Pogonin, A.E.; Kuznetsova, A.S.; Islyaikin, M.K.; Girichev, G.V. Gas-phase structures of
hemiporphyrazine and dicarbahemiporphyrazine: Key role of interactions inside coordination cavity. J. Mol. Struct. 2019,
1184, 576–582. [CrossRef]
35. Hamdoush, M.; Nikitin, K.; Skvortsov, I.; Somov, N.; Zhabanov, Y.; Stuzhin, P.A. Influence of heteroatom substitution in benzene
rings on structural features and spectral properties of subphthalocyanine dyes. Dye. Pigment. 2019, 170, 107584. [CrossRef]
36. Eroshin, A.V.; Otlyotov, A.A.; Zhabanov, Y.A.; Veretennikov, V.V.; Islyaikin, M.K. Complexes of ca(Ii), ni(ii) and zn(ii) with hemi-
and dicarbahemiporphyrazines: Molecular structure and features of metal-ligand bonding. Macroheterocycles 2021, 14, 119–129.
[CrossRef]
37. Zhabanov, Y.A.; Eroshin, A.V.; Stuzhin, P.A.; Ryzhov, I.V.; Kuzmin, I.A.; Finogenov, D.N. Molecular structure, thermodynamic
and spectral characteristics of metal-free and nickel complex of tetrakis(1,2,5-thiadiazolo)porphyrazine. Molecules 2021, 26, 2945.
[CrossRef]
38. Bursch, M.; Hansen, A.; Pracht, P.; Kohn, J.T.; Grimme, S. Theoretical study on conformational energies of transition metal
complexes. Phys. Chem. Chem. Phys. 2021, 23, 287–299. [CrossRef] [PubMed]
39. Maurer, L.R.; Bursch, M.; Grimme, S.; Hansen, A. Assessing Density Functional Theory for Chemically Relevant Open-Shell
Transition Metal Reactions. J. Chem. Theory Comput. 2021, 17, 6134–6151. [CrossRef]
40. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta
Crystallogr. Sect. A 1976, 32, 751–767. [CrossRef]
41. Sliznev, V.V.; Pogonin, A.E.; Ischenko, A.A.; Girichev, G.V. Vibrational spectra of cobalt(II), nickel(II), copper(II), zinc(II)
Etioporphyrins-II, MN4C32H36. Macroheterocycles 2014, 7, 60–72. [CrossRef]
42. Luc, N.Q.; Dang, V.S.; Tran, Q.T.; Pham, V.T.; Mai, A.T. Density Function Theory calculation, and phthalonitrile process for a
synthesis of single crystal zinc phthalocyanine. Mater. Sci. Semicond. Process. 2020, 113, 105025. [CrossRef]
43. Ruan, C.Y.; Mastryukov, V.; Fink, M. Electron diffraction studies of metal phthalocyanines, MPc, where M = Sn, Mg, and Zn
(reinvestigation). J. Chem. Phys. 1999, 111, 3035–3041. [CrossRef]
44. Allred, A.L.; Rochow, E.G. A scale of electronegativity based on electrostatic force. J. Inorg. Nucl. Chem. 1958, 5, 264–268.
[CrossRef]
45. Mann, J.B.; Meek, T.L.; Allen, L.C. Configuration energies of the main group elements. J. Am. Chem. Soc. 2000, 122, 2780–2783.
[CrossRef]
46. Gouterman, M. Spectra of porphyrins. J. Mol. Spectrosc. 1961, 6, 138–163. [CrossRef]
47. Gouterman, M.; Wagnière, G.H.; Snyder, L.C. Spectra of porphyrins. Part II. Four orbital model. J. Mol. Spectrosc. 1963, 11, 108–127.
[CrossRef]
48. Ehrenberg, B.; Johnson, F.M. Spectroscopic studies of tetrabenzoporphyrins: MgTBP, ZnTBP and H2TBP. Spectrochim. Acta Part A
Mol. Spectrosc. 1990, 46, 1521–1532. [CrossRef]
49. VanCott, T.C.; Koralewski, M.; Metcalf, D.H.; Schatz, P.N.; Williamson, B.E. Magnetooptical spectroscopy of zinc tetrabenzopor-
phyrin in an argon matrix. J. Phys. Chem. 1993, 97, 7417–7426. [CrossRef]
50. Koehorst, R.B.M.; Kleibeuker, J.F.; Schaafsma, T.J.; De Bie, D.A.; Geurtsen, B.; Henrie, R.N.; Van Der Plas, H.C. Preparation and
spectroscopic properties of pure tetrabenzoporphyrins. J. Chem. Soc. Perkin Trans. 1981, 7, 1005–1009. [CrossRef]
51. Stillman, M.; Mack, J.; Kobayashi, N. Theoretical aspects of the spectroscopy of porphyrins and phthalocyanines. J. Porphyr.
Phthalocyanines 2002, 6, 296–300. [CrossRef]
52. Nemykin, V.N.; Hadt, R.G.; Belosludov, R.V.; Mizuseki, H.; Kawazoe, Y. Influence of molecular geometry, exchange-correlation
functional, and solvent effects in the modeling of vertical excitation energies in phthalocyanines using time-dependent density
functional theory (TDDFT) and polarized continuum model TDDFT methods: Can modern computational chemistry methods
explain experimental controversies? J. Phys. Chem. A 2007, 111, 12901–12913. [CrossRef]
53. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.;
Petersson, G.A.; et al. Gaussian 09, Rev A.1; Gaussian Inc.: Wallingford, CT, USA, 2009.
54. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion
correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [CrossRef]
55. Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn:
Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297–3305. [CrossRef] [PubMed]
56. Schuchardt, K.L.; Didier, B.T.; Elsethagen, T.; Sun, L.; Gurumoorthi, V.; Chase, J.; Li, J.; Windus, T.L. Basis set exchange: A
community database for computational sciences. J. Chem. Inf. Model. 2007, 47, 1045–1052. [CrossRef]
57. Feller, D. The role of databases in support of computational chemistry calculations. J. Comput. Chem. 1996, 17, 1571–1586.
[CrossRef]
58. Pritchard, B.P.; Altarawy, D.; Didier, B.; Gibson, T.D.; Windus, T.L. New Basis Set Exchange: An Open, Up-to-Date Resource for
the Molecular Sciences Community. J. Chem. Inf. Model. 2019, 59, 4814–4820. [CrossRef] [PubMed]
59. Andrae, D.; Häußermann, U.; Dolg, M.; Stoll, H.; Preuß, H. Energy-adjusted ab initio pseudopotentials for the second and third
row transition elements. Theor. Chim. Acta 1990, 77, 123–141. [CrossRef]
60. Metz, B.; Stoll, H.; Dolg, M. Small-core multiconfiguration-Dirac-Hartree-Fock-adjusted pseudopotentials for post-d main group
elements: Application to PbH and PbO. J. Chem. Phys. 2000, 113, 2563–2569. [CrossRef]
Int. J. Mol. Sci. 2022, 23, 939 12 of 12

61. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Montgomery, J.A.; Vreven, J.T.; Kudin, K.N.;
Burant, J.C.; et al. Gaussian 03, Revision B.03; Gaussian Inc.: Wallingford, CT, USA, 2003.
62. Granovsky, A.A. Firefly Version 8. Available online: http://classic.chem.msu.su/gran/firefly/index.html (accessed on 15
April 2021).
63. Schmidt, M.W.; Baldridge, K.K.; Boatz, J.A.; Elbert, S.T.; Gordon, M.S.; Jensen, J.H.; Koseki, S.; Matsunaga, N.; Nguyen, K.A.;
Su, S.; et al. General atomic and molecular electronic structure system. J. Comput. Chem. 1993, 14, 1347–1363. [CrossRef]
64. Zhurko, G.A.; Zhurko, D.A. ChemCraft Version 1.6 (Build 312). Available online: http://www.chemcraftprog.com/index.html
(accessed on 1 December 2021).

You might also like