J. Serb. Chem. Soc. 84 (8) 779–800 (2019)
JSCS–5225
UDC 546.302:532.14:548.4+519.677
Authors’ Review
AUTHORS’ REVIEW
Calculation of the Jahn–Teller parameters with DFT•
MATIJA ZLATAR1# and MAJA GRUDEN2*#
1University
of Belgrade-Institute of Chemistry, Technology and Metallurgy, Department of
Chemistry, Njegoševa 12, Belgrade, Serbia and 2University of Belgrade, Faculty of
Chemistry, Studentski trg 12–16, Belgrade, Serbia
(Received 10 May, accepted 25 June 2019)
Abstract: In this review, а density functional theory (DFT) procedure is presented to calculate the Jahn–Teller (JT) parameters in a non-empirical way,
which does not depend on the system at hand. Moreover, the intrinsic distortion
path (IDP) model that gives further insight into the mechanism of the distortion
is presented. The summarized results and their comparison to experimentally
estimated values and high-level ab initio calculations, not only prove the good
ability of the used approach, but also provide many answers to the intriguing
behavior of JT active molecules.
Keywords: vibronic coupling; density functional theory; intrinsic distortion
path; distortion; transition metal complexes; organic ions and radicals.
CONTENTS
1. INTRODUCTION
2. METHODOLOGY
3. RESULTS AND DISCUSSION
4. CONCLUSIONS
5. REFERENCES
1. INTRODUCTION
Since its discovery more than 80 years ago, the Jahn–Teller (JT) effect1 has
drawn much attention in the scientific community, usually among scientists described as: “When a baby cries without anyone knowing why, we say - these are
the teeth. Similarly, in the chemistry of transition metal compounds, when experimental phenomenon cannot be interpreted easily, it is often attributed to the
Jahn–Teller effect. Moreover, the Jahn–Teller effect is simultaneously a source
* Corresponding author. E-mail:
[email protected]
#
Serbian Chemical Society member.
•
Dedicated to Professor Miljenko Perić on the occasion of his 70th birthday.
https://doi.org/10.2298/JSC190510064Z
779
Available on line at www.shd.org.rs/JSCS/
________________________________________________________________________________________________________________________
(CC) 2019 SCS.
780
ZLATAR and GRUDEN
of comfort and of anxiety. Comfort because it is still there as a last resort to explain a result. Anxiety because it is challenging to reliably prove it, without remaining ambiguity”.2 A witty quotation, translated from French, reflects the difficulties and challenges in both experimental and computational determination of
the JT effect. It echoes another common misapprehension – that the JT effect is
related only to transition metal complexes, which is certainly not the case. The JT
effect also occurs in highly symmetric open shell molecules, such as the cyclopentadienyl radical, benzene cation, fullerene ions, to name a few. After the
Nobel Prize for physics in 1987,3 when its importance in high-Tc superconductivity was highlighted, it became more apparent that this phenomenon is not only
of fundamental interest, but also applicable in many areas of chemistry and
physics.4
The JT theorem states that in a non-linear molecule with a degenerate
ground state, structural distortion must occur, which removes the degeneracy,
lowers the symmetry and stabilizes the system. In linear molecules because of
symmetry constraints degenerate, electronic states cannot couple with appropriate vibrations. However, these molecules can still be distorted, as described by
the Renner–Teller (RT) effect.5 Although here the JT effect is mentioned first, it
is noteworthy that historically it was preceded by the discovery of the RT effect.5
A requirement that a molecule must be in a degenerate electronic state to be a
subject of the JT and RT effects is excluded in the pseudo-Jahn–Teller (PJT)
formalism.4,6,7 All these effects belong to so-called “vibronic interactions.”4,8,9
Vibronic coupling is a quantum mechanical description of the influence of vibrations on the electronic structure, and vice versa. These effects are the origin of
various molecular properties, such as colossal magnetoresistance,10,11 high-Tc
superconductivity,3,12 nonconventional superconductivity,13 single-molecule
transport,14 ferroelectricity,4,15 spin-crossover16 and dynamics.9 They are essential in the design of single molecular magnets.17–19 The vibronic coupling is responsible for the observation of symmetry-forbidden transitions in spectroscopy,20
conical intersections in photochemistry, and is indispensable for understanding
the spectroscopy of linear molecules.21–26 The cause of any structural distortion
of a polyatomic system is due to the JT, RT, PJT effect, or their combination.4,7,27 Thus, the shape of molecules, both linear28 and non-linear,29,30 as
well as solids,31 is explained within the framework of vibronic coupling theories.
The coupling of electronic structure and nuclear movements is also essential for a
proper understanding of chemical reactions.32–36
Due to the coupling of electronic and vibrational motions, the Born–Oppenheimer (BO) approximation fails at the point of electronic degeneracy. That was
one of the main reasons why it has been wrongly assumed that first principle
calculations cannot be applied to analyze JT or RT effects. However, the breakdown of the BO approximation brings an apparent paradox. JT type effects
Available on line at www.shd.org.rs/JSCS/
________________________________________________________________________________________________________________________
(CC) 2019 SCS.
DFT AND JT EFFECT
781
emerge because of the non-validity of the BO approximation, but the effects as
such would not exist if the BO approximation is not used in the first place. The
JT effect exists only in the realm of the BO approximation. Without the BO
approximation, we do not have a picture of the molecule and there would be no
sense in discussing a distortion of a molecular structure. Standard concepts in
theoretical chemistry, such as force or force constants, would lose their meaning.
Nonetheless, a perturbation approach restores the BO approximation. A Hamiltonian is written as a Taylor series in nuclear coordinates. Coefficients in the
polynomial expression of an adiabatic potential energy surface are the vibronic
coupling constants, and they quantify the strength of the coupling between the
electronic structure and nuclear displacements.4 These coefficients are complicated integrals that have physical interpretation, such as force and force constant
at a high symmetry configuration.4 Conventional computational methods could
still be used if the adiabatic potential energy surface is accurately determined,
which is nowadays possible with most computational methods. Many studies
confirmed the excellent ability of both wave-function based methods and density
functional theory (DFT) based methods to analyze vibronic coupling.4,37 Typical
studies do not calculate the vibronic coupling constants directly, but the calculated adiabatic potential curves along the distortion are fitted to the polynomial
expression that resulted from the vibronic coupling model.30,38–41 Computational
studies enabled a more in-depth insight into the properties of JT active compounds and the understanding of many manifestations of the JT effect.
In the last decades, DFT has emerged as the mainstream among computational methods, preferable because it gives a good compromise between accuracy and computational time. The theory of DFT is well documented,42–48 and
here, just a brief explanation is given in a non-mathematical way of what the density functional theory is and how it can be applied to analyze the JT effect. In its
theory, the DFT gives the exact energy of the system as a method of obtaining a
solution to the Schrödinger equation of a many-body system, using electron density. The Hohenberg–Kohn theorem49 asserts that the density of any system determines all ground-state properties of the system, so the total ground state energy
of a multi-electron system is a functional of the density. Consequently, if the
electron density functional is known, the exact total energy of the system is
given. In practical computational work, as there is still no universal functional,
approximations have to be made, leading to many functionals designed for certain properties.43,50
There is a controversy in the literature whether DFT can be used for the
analysis of the JT effect,51–53 as the original Hohenberg–Kohn theorems49 were
formulated only for non-degenerate ground states in the absence of a magnetic
field. However, the second reformulation of this theorem gives formally proof
that the DFT can handle degenerate states.54,55 This issue is also elaborated
Available on line at www.shd.org.rs/JSCS/
________________________________________________________________________________________________________________________
(CC) 2019 SCS.
782
ZLATAR and GRUDEN
within the sub-system ensemble DFT,56,57 the spin-restricted ensemble-referenced Kohn–Sham (KS) DFT,58 the nonadiabatic generalization of DFT59 and
the sign-change in DFT.60 Additionally, DFT naturally encompasses the mechanism of the JT effect.61,62 Electron density must be totally symmetric in a point
group of a molecule. This is not possible if the orbitals belonging to a degenerate
irreducible representation (irrep) are not equally occupied. In such cases, non-totally symmetric density will lead to non-isotropic forces on nuclei, driving the distortion to the molecular structure with lower point-group than the initial one. A
molecular structure with degenerate electronic states in a high-symmetry (HS)
point group is, therefore, not a stationary point. In the case of the PJT effect, the
change in the distribution of electrons along a distortion is accompanied by a
lowering of the force constant of the high-symmetry configuration.4,7,29,63
Admittedly, from a practical point of view, care should be taken when dealing
with JT type effects with DFT.37,64 A pragmatic way for a detailed calculation of
the JT effect was developed by Daul et al.,37,64,65 and it has been successfully
applied for the analysis of many JT active molecules.64–75
2. METHODOLOGY
Within vibronic coupling theory, Hamiltonian, ℋ, is expanded as a Taylor series around
the HS molecular structure, along the normal modes, Qi:
ℋ = ℋ0 + Σi(∂ℋ/∂Qi)Qi+ 1/2Σij(∂2ℋ/(∂Qi∂Qj))QiQj + …
(1)
where ℋ 0 is the Hamiltonian for the reference nuclear configuration, and the sum that follows
it is the JT or vibronic Hamiltonian, W. The adiabatic potential energy surface for an f-fold
degenerate electronic state takes the form
Ek = 1/2 ΣiKiQi2 + εk
(2)
where k = 1,2,…f, and Ki is the force constant for the vibration Qi and εk are the roots of the
secular equation:
|W – εI| = 0
(3)
W is an f×f vibronic matrix, and I is a unit matrix of the same dimensions. The matrix
elements of the vibronic operator are the vibronic coupling constants. For instance, the linear
vibronic coupling constant has the form Fi,k = <Ψk|∂ℋ/∂Qi|Ψk>, where Ψk is the wavefunction
belonging to the k-th component of the degenerate state. Group theory can be used to judge
whether these matrix elements are different from zero. Irrep of the JT active normal modes
must belong to the same representation as the symmetric direct product of the components of
the degenerate electronic state, to have a linear coupling constant different from zero.
For example, in the case of the JT active Sb4-, electronic ground state in square planar
(D4h) configuration is 2Eg, and vibrations that belong to B1g and B2g irreps distort the structure
to rhombus and rectangle, respectively. Both distorted structures belong to the D2h point
group. The adiabatic potential energy surface of Sb4- in the space of two JT active vibrations
is depicted in Fig. 1. The parameters used to construct the surface were obtained by DFT.70
Another example is the VCl4 tetrahedral molecule (Td point group) with an E ground
electronic state coupled with double degenerate vibrations also belonging to E irrep. Distortion lowers the symmetry to D2d, and the degenerate state splits into 2A1 and 2B1. The adiabatic
Available on line at www.shd.org.rs/JSCS/
________________________________________________________________________________________________________________________
(CC) 2019 SCS.
DFT AND JT EFFECT
783
potential energy surface has the famous wrapped “Mexican-hat-like” form, with three equivalent minima and three equivalent transition states, Fig. 2.
Fig. 1. Contour plot of the adiabatic potential
energy surface of Sb4- in the space of JT active
b1g and b2g normal modes. Minima have
rectangular and transition states rhombic geometries.70
Fig. 2. Contour plot of the adiabatic potential
energy surface of VCl4 in the space of two
components of the JT active e vibration.
Minima have 2A1, and transition states 2B1
state.64
Vibronic coupling constants determine the shape of the adiabatic potential energy surface
(Figs. 1 and 2). However, from a computational perspective, it is more natural to define it with
the JT stabilization energy (EJT), warping barrier (Δ) and JT radius (RJT). A qualitative cut
through the adiabatic potential energy surface, along the JT active distortion Qa, is given in
Fig. 3, indicates how this set of JT parameters define the adiabatic potential energy surface of
JT active molecules. The meaning of the parameters is clear – the value of EJT gives the
energy stabilization due to the distortion, the energy difference between the minimum (min in
Fig. 3) and transition state (TS in Fig. 3) is Δ, and the direction and magnitude of the distortion are given with RJT.
The set of JT parameters, Fig. 3, is, at least in principle, easy to determine from the first
principle calculations. To obtain the JT parameters, one needs to know the proper geometry
and, consequently, the energy of the HS configuration, as well as the geometries and energies
of the distorted, electronically non-degenerate low symmetry (LS) structures. Experimental
determination of the parameters is less straightforward. One needs to fit the experimentally
obtained results to the proposed model.
The procedure for calculating the JT parameters using DFT consists of three steps:37,64
1. Geometry optimization constraining symmetry to the HS point group. This yields a
geometry of the high-symmetry configuration, with an electronic structure in which the electrons of degenerate orbitals are distributed equally over the components of the degenerate
Available on line at www.shd.org.rs/JSCS/
________________________________________________________________________________________________________________________
(CC) 2019 SCS.
784
ZLATAR and GRUDEN
irreps, e.g., for e1-configuration this will mean to place 0.5 electrons into each of the two
(alpha) e-orbitals; for e3-configuration, 0.5 electrons will be in each of two beta e orbitals.
2. Performing a single-point calculation constraining the HS on the nuclear geometry and
the LS on the electron density. This gives the energy of a Slater determinant with an integer
electron orbital occupancy. It is necessary to evaluate the energies of all possible Slater determinants with integer electron occupations. In other words, all possible modes of occupation
of the molecular orbitals need to be evaluated. This step is achieved by introducing an adequate occupation scheme of the molecular orbitals (MO).
3. A geometry optimization constraining the structure to the low-symmetry point group,
with the proper occupancy of the KS orbitals. These calculations yield different geometries
and energies that correspond to a minimum and a transition state on the adiabatic potential
energy surface.
Fig. 3. Qualitative cross-section through the
adiabatic potential energy surface, along the
JT active distortion Qa from the high-symmetry point (HS) toward the minimum (min)
and transition state (TS); definition of the JT
parameters: the JT stabilization energy (EJT),
the warping barrier (Δ) and the JT radius
(RJT).
This calculation scheme is schematically drawn in Fig. 4 for the calculation of the JT
parameters in octahedral Cu(II) complexes. EJT is the difference between the energies
obtained in steps 2 and 3. As stated above, steps 2 and 3 are repeated for different combin-
Fig. 4. DFT approach for the calculation of the JT parameters. The JT effect in an
octahedral Cu(II) complex is taken as an example.
Available on line at www.shd.org.rs/JSCS/
________________________________________________________________________________________________________________________
(CC) 2019 SCS.
DFT AND JT EFFECT
785
ations of electronic states in the LS point group. Hence there will be different EJT for a
minimum and transition state structures. The difference between the two EJT values gives the
warping barrier, Δ. RJT is given by the length of the distortion vector between the high symmetry and the minimum energy configuration, i.e., the difference in the geometries from step
1 and 3.
To avoid calculating the geometry and energy at the HS configuration, it is possible to
use the energy of the vertical (Franck–Condon) transition, EFC, from the LS distorted structures.74 This energy is easily obtained by promoting the unpaired electron from the ground
state to the first excited state for the ground state (LS) geometry. When anharmonicity is negligible, i.e., in the linear and quadratic JT models, EJT = 4EFC.
The first step consists of the calculation with a non-integer electron configuration. Since
MOs themselves have no physical meaning, using the partial occupation is just a way of
obtaining the electron density of the A1 symmetry in the HS point group. As the electron
density is totally symmetric, it exerts the same force on all atoms. After geometry optimization in step 1, the total force is null. Passing from this configuration with fractional
occupation numbers, to the one with integer occupation, in the second and third step, leads to
a change in the electronic density responsible for the non-zero force. Distortion thus occurs
because a change in the electronic density requires modifying the equilibrium geometry.42,61
This is illustrated in Fig. 5, where the difference in electron densities between steps 2 and 1
for the case of [CuF6]4- is depicted, together with the resulting force (calculated at the LDA/
/TZP level of theory). The resulting forces lead to tetragonal distortion and stronger equatorial
bonds compared to the axial ones. Lowering of the symmetry in step 2 is necessary because
this is the only way that the electron density is totally symmetric with integer occupations of
the MOs.
Fig. 5. Change of the electron density from the orbital configuration
with fractional occupation to the one with integer occupation in the Oh
configuration of [CuF6]4-. Light/dark color represents electron density
depletion/concentration (iso-value 0.002 a.u). Forces on atoms are
represented as black arrows.
Calculation of the Hessian at step 3 indicates the character of the stationary points –
minimum or transition state. This correlates with the relative magnitudes of the two EJT. Interestingly, it is possible to calculate the Hessian also in step 1, and all eigenvalues should be
positive. This means that geometry from step 1 is the minimum for such an electronic configuration (with fractional occupation numbers). This nuclear configuration is a minimum
structure (hypothetically) in the absence of the JT effect.
The JT parameters that are obtained from the calculations, Figs. 3 and 4, relate to the vibronic coupling constants.4 However, the exact relations will depend on the vibronic coupling
model used. For example, in the case of Sb4-, for each of the two JT active vibrations (b1g and
b2g), Fig. 1, the relations are EJT = F2/2K and RJT = F/K, where F is the linear vibronic coup-
Available on line at www.shd.org.rs/JSCS/
________________________________________________________________________________________________________________________
(CC) 2019 SCS.
786
ZLATAR and GRUDEN
ling constant, and K is the force constant of the normal mode.70 In the case of VCl4, Fig. 2,
second-order vibronic coupling is operational, and the relations are: EJT = F2/2(K–|G|),
RJT(min) = |F|/(K–2|G|), RJT(TS)=|F|/(K+2|G|), Δ=4EJT/(K+2|G|), where F, K and G are the
linear vibronic coupling constant, the force constant and the quadratic vibronic coupling constants, respectively.64
It is worth mentioning that the calculated JT parameters are not very sensitive to the
choice of exchange-correlation functional.73 However, the JT parameters strongly depend on
geometry. Hence the proper choice of the functional for the system at hand is of utmost importance. For organic molecules, this is not a difficult task, but for transition metal compounds
particular caution (especially concerning spin states) is required.76
Further insight into what is occurring during distortion is based on the analysis of RJT. It
has already been mentioned that RJT describes the direction and magnitude of the distortion. In
the case of the triangular Na3 cluster, it is represented by changing the angle.40 In other simple
cases, the distortion corresponds to the movements of nuclei along one normal mode that
belongs to a non-totally symmetric irrep of the HS point group of a molecule. RJT is then the
magnitude of the distortion along that normal mode. In simple octahedral Cu(II) complexes
often the following relation is used:65 RJT2=Σi δdi, where δdi = di – d0, where di is the distance
from the Cu(II) ion to the i-th ligand, and d0 is the average metal–ligand distance. In more
complex molecules, the JT distortion is a consequence of many active modes. The evaluation
of the influence of different normal modes on the JT effect is a “multimode problem,” and
suitable models have to be employed to analyze the distortion.4,37,41,77 The two most successful models for the analysis of the multimode problem are the “interaction mode”41,78,79 and
the “intrinsic distortion path” (IDP).37,67,77
The interaction mode41,78,79 is a single, effective mode that is built as a linear combination of all the JT active normal modes of the HS configuration. Coefficients in this linear
combination depend on the vibronic coupling constants of each mode, and the vibronic coupling model used. Typically, the interaction mode describes the straight-line path between the
HS and LS geometries, i.e., its direction coincides with the direction of RJT. It is crucial to
notice that the interaction mode is a reducible representation in the HS point group, while it
belongs to the totally symmetric irreducible representation in the LS point group. The same is
true for RJT.
The essence of the IDP model37,67,77 is to express the distortion along a model minimal
energy path starting from the DFT, or the first-principle HS geometry, and ending in the LS
minimum, or the LS transition state. The reference point in IDP is the LS configuration, and
the energy surface is approximated to be harmonic, i.e., quadratic in the LS vibrational modes.
This model allows the determination of both the presence and contribution to the EJT and RJT
of all normal modes involved. All totally symmetric normal modes in the LS configuration
contribute to the distortion because IDP as a “reaction path” must be totally symmetric.36 The
analytical expression of the energy allows the total force at the HS point to be obtained
directly. This force can be projected on either HS or LS normal modes. Knowing that the
physical interpretation of the linear vibronic coupling constants is the force along an HS normal mode,4 the IDP directly evaluates all the linear vibronic coupling constants.77 The contributions of each mode along the path change, thus providing a very detailed picture of the
mechanism of the distortion.
Both, interaction mode and IDP have their advantages and drawbacks, as discussed in the
literature,77 and reduce a multidimensional energy surface that is difficult to visualize to a
simple cross-section.
Available on line at www.shd.org.rs/JSCS/
________________________________________________________________________________________________________________________
(CC) 2019 SCS.
787
DFT AND JT EFFECT
3. RESULTS AND DISCUSSION
Using the procedure mentioned above, many JT active molecules were analyzed, ranging from small metal clusters and organic radicals to organo-metallic
compounds and coordination compounds (Table I). All these molecules differ in
the number of atoms, nature of the chemical bonding, symmetry of the distortion,
range of the JT stabilization energy. However, in all cases, we obtained good
agreement with experimentally evaluated values or high-level ab initio data.
TABLE I. Results of the DFT calculations performed to analyze the JT effect of a set of compounds, and comparison to the available reference values (ref.)
37,64
67
73
70
70
37,66
66
66
66
77
92
68
68
68
68
67
96
96
71
71
71
98
98
EJT (ref.)
cm-1
30–80
280–550
760–1160
–
–
1237
700–1000
–
1043–1374
557
280–600
–
952
484
645
1050
350
–
–
406
–
1800
1609
80–83
4
40,84,85
–
–
86
87–89
–
90
91
4
–
93
94
94
80,95
80,95
–
–
97
–
99
100
1995
98
–
–
D3→C2
2168
98
–
–
D3→C2
D3→C2
D3→C2
2200
1980
253
75
75
18
1800–2200
1620–2000
–
101
101
–
Molecule
Distortion
VCl4
Cu3
Na3
Sb4As4•
C5H5
C6H6+
C6H6•
C7H7
C60+
C60C20H10+
C20H10C24H12+
C24H12CoCp2
MnCp2
[FeCp2]+
Pc3MgPcMnPc
[Cu(TACN)2]2+
[Ni(TACN)2]3+
(low spin)
[Co(TACN)2]2+
(low spin)
[Co(TACN)2]3+
(intermediate spin)
[Cu(en)3]2+
[Cu(eg)3]2+
[NiCl3(Hdabco)2]+
Td→D2d
D3h→C2v
D3h→C2v
D4h→D2h
D4h→D2h
D5h→C2v
D6h→D2h
D6h→C2v
D7h→C2v
Ih→D5d
Ih→C2h
C5v→Cs
C5v→Cs
D6h→D2h
D6h→D2h
D5h→C2v
D5h→C2v
D5h→C2v
D4h→D2h
D4h→D2h
D4h→D2h
D3→C2
D3→C2
EJT(DFT)
cm-1
50
530
870
579
840
1235
879
1187
853
600
300–315
548–665
404–540
287–401
290–410
870
500
462
319–454
399–614
137–152
1790
1504
D3→C2
Ref. (DFT)
Ref.
VCl4 and Sb4–, Figs. 1 and 2, have already been mentioned. The modes that
can contribute to the distortion are a1 and e for VCl4 and a1g, b1g and b2g for
Available on line at www.shd.org.rs/JSCS/
________________________________________________________________________________________________________________________
(CC) 2019 SCS.
788
ZLATAR and GRUDEN
Sb4–. Totally symmetric modes do not change the symmetry. However, in principle, need to be taken into consideration since their vibronic coupling constants
are not zero by symmetry. Our analysis based on the IDP model indicated that
totally symmetric modes do not contribute to the distortion in these molecules.37,70 Thus, the JT effect in VCl4 is in essence an example of ideal, single
mode JT distortion, and Sb4– is the simplest multimode problem. However, in
Sb4– only one mode is responsible for the distortion to the minimum and the
other one to the transition state. IDP is designed for more complicated situations77 but is also perfectly suitable for these cases. IDP analysis justifies illustration of adiabatic potential energy surfaces for these molecules as depicted in
Figs. 1. and 2. As4– is completely an analog to Sb4–.70 Vibronic coupling constants for all three molecules were calculated by IDP, by their relation to the DFT
obtained JT parameters, and by fit to the potential energy curves along the JT
active normal modes. The results of the different approaches are consistent. DFT
results for VCl4 (EJT = 50 cm–1)37,64 are in good agreement with experimental
values (EJT = 30–80 cm–1),80–83 and together with the values of the vibronic
coupling constants, confirm the dynamic character of the JT effect.
In VCl4, for the first time, some interesting features related to the DFT
scheme for the calculation of the JT parameters were observed.58 These observations are found to be general in other cases. The first one is related to the
energies of different electron distributions in step 2 (Fig. 4), which should be
equal for all the possibilities. However, in practice, this is not the case due to the
nature of the exchange-correlation functional involved in practical DFT calculations. The second finding is that ordering of orbitals in step 2 is non-aufbau. In
other words, the singly occupied MO stays, after the energy minimization, above
the lowest empty orbital. The expected orbital ordering is usually restored after
relaxation of the geometry (step 3). When distortion is very small, e.g., in VCl4,
the occupation stays even after the geometry optimization in step 3. It should be
pointed out that the JT effect is a change in geometry due to a lowering of the
total energy and the KS MO energies do not need to reflect that. In DFT, the total
energy is not equal to the sum of the KS orbital energies. The third remark is that
partial occupation of degenerate orbitals (step 1) usually leads to lower energy
than the one-electron-one-orbital occupancy (step 2), because of different selfinteraction errors in the two cases. In the case of VCl4, the energy from step 1
can be even lower than the final energies from step 3, giving a misleading result.
The cyclopentadienyl radical (•C5H5) is one of the most studied molecules
from the JT perspective, and the theoretical methods employed range from
simple MO calculations102 to high-level ab-initio calculations (Table II). Studies
of Mueller et al. who used complete active space methods (EJT = 2147 cm–1 )86
and dispersed fluorescence spectroscopy (EJT = 1237 cm–1),86 as well as fitting
ab-initio calculations to the spectra (EJT = 1463 cm–1)86 are considered to be
Available on line at www.shd.org.rs/JSCS/
________________________________________________________________________________________________________________________
(CC) 2019 SCS.
789
DFT AND JT EFFECT
benchmark results for the determination of the JT parameters in the cyclopentadienyl radical. The authors also identified three dominant normal modes necessary to explain their results. Results obtained by DFT are comparable,37,66 or in
some cases even better than other theoretical methods (Table II). With the IDP
model, it was possible to identify the three most important vibrations (C–C stretch,
C–C–C bending, and C–C–H bending),66 in agreement with the computational
and experimental studies by Miller et al.86
TABLE II. Summary of various computational methods used to study the JT effect in the
cyclopentadienyl radical
Methoda
MO
MO
MO
HF/STO-3G
CI/STO-3G
HF/6-311+G*
MP2/6-311+G*
MP4/6-311+G*
CCSD/6-311+G*
CCSD(T)/6-311+G*
CASSCF/cc-PVDZ
CASSCF/6-31G*
CASSCF/6-31G*
CASSCF/cc-PVDZ
CISD/cc-PVDZ
EOMIP-CCSD/DZP
DFT(LDA)/TZP
DFT (B3LYP)/TZP
IDP
Exp.
EJT / cm-1
560
728
495
5072
2484
1452
3065
2581
1613
1613
2139
2147
1463
1665
2553
1581
1253
1685
1239
1237
Reference
Liehr 1956102
Snyder 1960103
Hobey et al. 1960104
Meyer et al. 1979105
Borden 1979106
Cuhna et al. 1999107
Cuhna et al. 1999107
Cuhna et al. 1999107
Cuhna et al. 1999107
Cuhna et al. 1999107
Bearpark et al. 1999108
Miller et al., 200186
Miller et al., 200186
Kiefer et al. 2001109
Zilberg et al. 2002110
Stanton et al. 2008111
Zlatar et al., 200937
Zlatar et al., 201373
Zlatar et al., 2010
Miller et al., 200186
aHF – Hartree–Fock; CI – configuration Interaction; MPN – Moeller–Plesset perturbation theory of order N;
CCSD(T) – coupled cluster single, double (triple) excitations; CASSCF – complete-active-space SCF; CISD –
single and double excitations, single reference CI method; EOMIP-CCSD – equation-of-motion ionization potential coupled cluster single, double excitations; LDA – local density approximation; B3LYP – Becke’s 3-parameters Lee–Yang–Parr hybrid functional
DFT results show that the energy difference between the two distorted LS
structures of the cyclopentadienyl radical is only around 1.5 cm–1.37,66 Hence, it
is an example of a dynamic JT effect. According to DFT calculations, the 2A2
structure (en-alyl) is a transition state, while 2B1 (dienyl) is a minimum.37,66 This
is in agreement with CASSCF(5,5)/cc-pVTZ112 and CASSCF(5,5)/6-31G*86
calculations that report the dienyl structure to be a minimum and Δ to be 4.5 and
3.6 cm–1, respectively. The conversion between the 2A2 and 2B1 states goes
around the JT cusp, Fig. 6. Due to the five-fold permutational symmetry of the
cyclopentadienyl radical, there needs to be five equivalent dienyl and five equi-
Available on line at www.shd.org.rs/JSCS/
________________________________________________________________________________________________________________________
(CC) 2019 SCS.
790
ZLATAR and GRUDEN
valent en-alyl structures,113 which interconvert to each other with practically no
energy barrier.
Fig. 6. Schematic representation of the
pseudorotation on the lowest sheet of
the adiabatic potential energy surface
of the cyclopentadienyl radical. C–C
bond length, Å, which is chosen to be
the linear transit parameter is marked.
According to DFT calculations, the 2A2 conformations have one imaginary
frequency belonging to the B2 irrep in C2v point group, which lowers the symmetry to Cs. Pseudorotation of the cyclopentadienyl radical goes along the path
that has Cs symmetry, except at the ten configurations of C2v symmetry that
correspond to the stationary points on the potential energy surface. To test
whether such a small Δ is just due to the numerical noise in DFT calculations, the
lowest sheet of the adiabatic potential energy surface was explored using relaxed
linear transit calculations. The results are shown in Fig. 7. One of the C–C bond
lengths was varied from 1.469 Å, i.e., single C–C bond in 2B1, to 1.336 Å, i.e.,
double C=C bond in 2A2. The symmetry was constrained to Cs. All the remaining
coordinates were allowed to be optimized. This process corresponds to half the
pseudorotation depicted in Fig. 6.
In addition to the cyclopentadienyl radical, the JT effect in the family of
cyclic conjugated hydrocarbons was analyzed,66 including the benzene cation
(C6H6+), the benzene anion (C6H6–) and the tropyl radical (•C7H7). All three
molecules have a double degenerate electronic ground state (2E1g, 2E1g and E2ʹʹ,
respectively) in the HS configuration (D6h, D6h and D7h, respectively) that is
coupled with double degenerate JT active normal modes (e1g, e1g and e3ʹ, respectively). The DFT results are in excellent agreement with previous studies
(Table I). The IDP model showed that the C–C stretch is most important for
energy stabilization, while C–C–C bend, and C–C–H bend contribute more to the
Available on line at www.shd.org.rs/JSCS/
________________________________________________________________________________________________________________________
(CC) 2019 SCS.
DFT AND JT EFFECT
791
RJT. The energy component analysis along IDP revealed two distinct regions.
Most of the energy stabilization is achieved by electron–nuclear attractions in the
first region, where the C–C stretch prevails. Other forces due to electron–electron, nuclear–nuclear, and kinetic interactions are dominant in the second region.
In the second region, little stabilization is achieved, and the molecules relax
toward the LS stationary point via bending deformations.
Fig. 7. Linear transit calculations used for the
exploration of the lowest sheet of the potential energy surface of the cyclopentadienyl
radical.
In the benzene cation and anion, the symmetry of the HS nuclear configuration, the symmetry of the ground electronic state, and symmetry of the JT active
vibrations are all the same. However, the minimum on the potential energy surface is different. While in C6H6+ the minimum is D2h conformation, in C6H6–
due to pseudo-Jahn–Teller coupling of the ground π∗ electronic state with the
excited σ∗ state, the C2v conformation is found to be the global minimum. The
out-of-plane C2v geometry obtained is a consequence of both JT and pseudo-JT
distortion.
Corannulene (C20H10) and coronene (C24H12) and their ions are interesting
molecules because besides being building blocks for fullerene, they also have
application in electronic devices. Distortion in these molecules,68 as in other carbohydrates always starts with the modes that are predominantly C–C stretching,
and these modes contribute the most to the overall distortion due to the descent in
symmetry.
The neutral fullerene C60 molecule has very high symmetry. It belongs to the
Ih point group, with a high degree of degeneracies. The highest occupied molecular orbital is fivefold degenerate (hu), and the lowest unoccupied molecular
orbital is triply degenerate (t1u), Fig. 8. The distortion in the JT active C60+ and
C60– may lead to the structures belonging to the three different epikernel subgroups, i.e., to the structures with D5d, D3d or D2h symmetry. The JT active
modes in C60– belong to Hg irrep. The stable geometry of the C2h symmetry,
resulting from JT distortions of double degenerate states in D5d or D3d, is predicted to be the most stable.92 In C60–, the energy differences between different
Available on line at www.shd.org.rs/JSCS/
________________________________________________________________________________________________________________________
(CC) 2019 SCS.
792
ZLATAR and GRUDEN
distorted structures are very small.92 According to DFT calculations, the 2A1u
conformation in the D5d point group is the global minimum for C60+.77 The vibronic coupling in C60+ is very complicated. The D5d structure is a consequence
of the hg modes described by one set of the linear vibronic coupling coefficients.
D3d structures are due to the distortion led by the hg modes with a different set of
vibronic coupling coefficients, but gg modes are also active. In addition, the totally symmetric normal modes also have non-zero vibronic coupling. Utilizing the
IDP model, all the vibronic coupling constants of all normal modes in C60+ were
calculated77 – two ag, six gg and two sets of constants for eight hg modes. The
obtained values agree with previously calculated vibronic coupling constants
from the gradient of the HOMO level.91 The vibronic coupling constants of the
eight hg modes and two ag modes in C60– were also calculated92 by the IDP
model, which agreed with previous reports.91,114 The choice of the geometry of
the Ih structure was also highlighted, in particular, related to the contributions of
totally symmetric normal modes. According to IDP, most of the stabilization
energy is achieved close to the Ih point by the hardest JT active modes. These
modes have a large C–C stretch character. The relaxation of the geometry in the
final part of the distortion path is encountered by softer modes. The situation is
completely analog to the JT effect in the cyclopentadienyl radical and the benzene cation.
Fig. 8. Qualitative MO scheme of the fullerene anion, neutral fullerene, and the fullerene cation.
JT instable metallocenes, e.g., low spin d7 cobaltocene (CoCp2, Cp = C5H5),
low spin d5 manganocene (MnCp2) and low spin ferrocenyl cation, ([FeCp2]+)
are typical examples of multimode JT distortions. The HS conformation of metallocenes can be either D5h if the two cyclopentadienyl rings are eclipsed, or D5d if
two rings are staggered. According to DFT calculations, the eclipsed conformation is the more stable one,72 and descent in symmetry due to the distortion
goes to C2v LS structures. The EJT agrees with the experimental results, and do
not differ much for the three metallocenes (Table I).80,95 In the C2v minimum
energy conformation, both cobaltocene and manganocene have 16 totally symmetrical normal modes that all could in principle contribute to the distortion.
Available on line at www.shd.org.rs/JSCS/
________________________________________________________________________________________________________________________
(CC) 2019 SCS.
DFT AND JT EFFECT
793
With IDP model, not only their contribution to the RJT, but also their contribution
to the EJT was determined.67
The ground state of cobaltocene is 2E1ʹʹ with a single electron in e1ʹʹ MO.
This MO is antibonding between metal dxz (or dyz) and the ligand π-system.
Using group theory, it is easy to show that the distortion coordinate is e2ʹ. Out of
16 vibrations, 4 contribute more than 95 % to the EJT, and all four are first order
JT active, i.e., e2ʹ, and all are located on cyclopentadienyl ring, which is no
longer planar. Although in most cases the electronic distortion due to the JT
effect is mainly localized on the central metal ion, here it is delocalized over the
ligands. The symmetry of the electronic ground state in the HS point pushed the
distortion towards perturbation of the aromaticity of the Cp rings. In other words,
JT induces conformational changes of the ligands, and out-of-plane ring distortion and C–H wagging minimize the antibonding interaction between the
single occupied d orbital of the Co(II) ion and the π system of the Cp rings.
Manganocene with 5 electrons in d-orbitals is in the 2E2ʹ ground state (low
spin is the most stable one) with a single hole in the nonbonding e2ʹ orbital in the
D5h symmetry, with the very close lying 2A1ʹ state. Distortion lowers the symmetry to C2v along the e1ʹ normal coordinates. Skeletal vibrations contribute the
most to the EJT and enhance small bonding interactions between the d-orbitals
and the π-system of Cp rings.67 These distortions are localized around the metal.
However, IDP analysis revealed that there is one more vibrational mode that
bends the two Cp rings and has an influence on the JT distortion, but does not
bring a dominant energetic contribution.67,96
Another intriguing feature is the unusual behavior of manganocene compared to its isoelectronic analog ferrocenium cation and other metallocenes.
Namely, in the solid state, manganocene forms a zig-zag polymer in which the
ground state is the high spin state, which was not observed in any other case.115
Our thorough study in which DFT, energy decomposition analysis and IDP
methods were applied, revealed the reason behind this peculiar behavior of
MnCp2.96 It was shown that the close lying 6A1ʹ state allows conversion to the
HS state present in the zigzag polymer. Moreover, the close lying excited state
2A ʹ enables ligand deformation through pseudo-JT coupling and triggers the
1
polymerization. Hence, the unique behavior of manganocene has been rationalized and explained by its degenerate ground state and close lying electronic and
spin states.96
Phthalocyanine trianion (Pc3−), magnesium phthalocyanine ion (MgPc–) and
manganese phthalocyanine (MnPc) are all prone to JT distortion. Detailed analysis71 showed that the coordinated metal dictated the way of breaking the symmetry: while in MgPc–, the distortion is mainly localized on the ligand, in MnPc,
central metal ion presents the trigger for the occurrence of JT distortion over the
whole system. This difference was rationalized with IDP analysis that indicated
Available on line at www.shd.org.rs/JSCS/
________________________________________________________________________________________________________________________
(CC) 2019 SCS.
794
ZLATAR and GRUDEN
different normal modes responsible for the JT effect in these similar systems. The
linear vibronic coupling constants of all 24 JT active modes were determined by
IDP model that agreed with previous work.71
A study98 on complexes of bis(1,4,7-triazacyclononane) (TACN) with some
first-row transition metal ions with the ground term subject to JT distortion, revealed that vibronic coupling must be taken into consideration, even when the JT
distortion is negligible because it can contribute to all the properties of the system. In all cases, the EJT values obtained by DFT calculations were in excellent
agreement with experimentally observed ones, confirming the excellent ability of
the DFT approach to calculate JT parameters even when the complicated electronic structure of transition metal compounds have to be considered. Furthermore, IDP analysis gave further insight into what is happening during the distortion and distinguished the most important normal modes out of 71 to the EJT and
RJT. It has been shown that the second coordination sphere does not influence the
values of the JT parameters,98 as in the case of [Cu(en)3]2+ (en is ethylenediamine) and [Cu(eg)3]2+ (eg is ethylene glycol).75
Pt(PF3)4 is tetrahedral d0 molecule in the 1A1 ground electronic state. This
molecule is an important precursor for focused electron beam induced processing.116 In this technique, a focused beam of electrons is used to initiate the
dissociation of ligands in the precursor, ideally resulting in a pure metal deposit.
Several processes occur during this process, and the importance of so-called
neutral dissociation was first highlighted in Pt(PF3)4.117 Neutral dissociation is
the dissociation of electronically excited states initiated by incoming electrons.
At first sight, this small symmetric molecule is not related to the JT effect. However, its lowest excited states are degenerate (lowest 1,3T1 and 1,3T2) and are
consequently subject to JT distortion. Distortion in these excited states is asymmetric Pt–P stretch of T2 symmetry and this is exactly the dissociation coordinate. Therefore, the JT effect in the excited states triggers dissociation.117
These four excited states are directly dissociative, while other excited states are
dissociative via many conical intersections.117
Five-coordinated Ni(II) trigonal–bipyramidal complexes (D3h) in a high spin
are in 3Eʹ ground state, hence prone to the JT distortion to the C2v structure. It
has been shown that controlling the geometry of a transition metal complex is the
way to chemically control their magnetic properties, particularly that spin–orbit
coupling (SOC) could be used to tune the magnetic anisotropy.18,19 If the SOC is
higher, the magnetic anisotropy has a greater value. Knowing the fact that both
coordination of the ligands and the partial quenching of the SOC due to JT
distortions reduce the magnetic anisotropy from its ideal, the free ion value, the
system ([NiCl3(Hdabco)2]+, dabco is 1,4-diazabicyclo[2.2.2]-octane), was found
in which the bulky dabco axial ligands prevent distortion.18 As a consequence of
quenched JT, the calculations revealed,18 and later it was experimentally con-
Available on line at www.shd.org.rs/JSCS/
________________________________________________________________________________________________________________________
(CC) 2019 SCS.
DFT AND JT EFFECT
795
firmed,118 that this complex displays the largest magnetic anisotropy ever observed for a mononuclear Ni(II) complex.
4. CONCLUSIONS
Cast in the form of authors’ review paper, this work summarizes the results
obtained by a special procedure within the framework of DFT, for the analysis of
the JT effect in many different molecules. The theoretical background and practical computational recipe of this non-empirical and effective method are given.
Treatment of the multimode JT problem with the IDP approach provides microscopic insight into the symmetry breaking process and rationalize many manifestations of vibronic coupling. Therefore, the fast and reliable method presented
herein could be considered as the preferable tool for the investigation of adiabatic
potential energy surface of the JT active molecules.
Acknowledgements. This work is supported by the Ministry of Education, Science and
Technological Development of the Republic of Serbia (Grant No. 172035).
ИЗВОД
ИЗРАЧУНАВАЊЕ ЈАН–ТЕЛЕРОВИХ ПАРАМЕТАРА ПРИМЕНОМ ТЕОРИЈЕ
ФУНКЦИОНАЛА ГУСТИНЕ
МАТИЈА ЗЛАТАР1 и МАЈА ГРУДЕН2
1
Универзитет у Београду-Институт за хемију, технологију и металургију, Центар за хемију,
1
Његошева 12, Београд и Универзитет у Београду-Хемијски факултет, Студентски трг 12–16,
Београд
У овом прегледном раду, представљена је не-емпиријска процедура за израчунавање
Јан–Телерових параметара применом теорије функционала густине, која не зависи од конкретног система који се проучава. Представљен је и модел својственог пута дисторзије, који
даје додатни увид у механизам дисторзије. Сумирани резултати и њихово поређење са експериментално процењеним вредностима, као и поређење са резултатима ab initio прорачуна
високог нивоа, доказују тачност и велику применљивост коришћене процедуре. Такође, овде
приказани рачунарски приступ даје многе одговоре на интригантне особине Јан–Телерактивних молекула.
(Примљено 10. маја, прихваћено 25. јуна 2019)
1.
2.
3.
4.
5.
6.
REFERENCES
H. A. Jahn, E. Teller, Proc. R. Soc., A 161 (1937) 220
(https://dx.doi.org/10.1098/rspa.1937.0142)
O. Kahn, Structure électronique des éléments de transition: ions et molécules complexes,
Presses universitaires de France, 1977
J. G. Bednorz, K. A. Müler, Perovskite-Type Oxides – The New Approach to High-Tc
Superconductivity - Nobel Lecture, in Nobel Lect. Phys. 1981–1990, G. Ekspang, Ed.,
World Scientific Publishing Co, Singapore, 1993, pp. 424–457
I. B. Bersuker, The Jahn–Teller Effect, Cambridge University Press, Cambridge, 2006
(https://dx.doi.org/10.1017/CBO9780511524769)
R. Renner, Z. Physik 92 (1934) 172 (https://dx.doi.org/10.1007/BF01350054)
U. Öpik, M. H. L. Pryce, Proc. R. Soc., A 238 (1957) 425
(https://dx.doi.org/10.1098/rspa.1957.0010)
Available on line at www.shd.org.rs/JSCS/
________________________________________________________________________________________________________________________
(CC) 2019 SCS.
796
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
ZLATAR and GRUDEN
I. B. Bersuker, Chem. Rev. 113 (2013) 1351 (https://dx.doi.org/10.1021/cr300279n)
T. Azumi, K. Matsuzaki, Photochem. Photobiol. 25 (1977) 315
(https://dx.doi.org/10.1021/cr300279n)
H. Köppel, W. Domcke, L. S. Cederbaum, Multimode Molecular Dynamics Beyond the
Born–Oppenheimer Approximation, Adv. Chem. Physics, Vol. 57, in I. Prigogine, S. A.
Rice, Eds., Wiley 2007, pp. 59–246 (https://dx.doi.org/10.1002/9780470142813.ch2)
A. J. Millis, P. B. Littlewood, B. I. Shraiman, Phys. Rev. Lett. 74 (1995) 5144
(https://dx.doi.org/10.1103/PhysRevLett.74.5144)
A. J. Millis, B. I. Shraiman, R. Mueller, Phys. Rev. Lett. 77 (1996) 175
(https://dx.doi.org/10.1103/PhysRevLett.77.175)
J. G. Bednorz, K. A. Müller, Z. Physik, B: Cond. Matter 64 (1986) 189
(https://dx.doi.org/10.1007/BF01303701)
R. H. Zadik, Y. Takabayashi, G. Klupp, R. H. Colman, A. Y. Ganin, A. Potočnik, P.
Jeglič, D. Arčon, P. Matus, K. Kamarás, Y. Kasahara, Y. Iwasa, A. N. Fitch, Y. Ohishi,
G. Garbarino, K. Kato, M. J. Rosseinsky, K. Prassides, Sci. Adv. 1 (2015) e1500059
(https://dx.doi.org/10.1126/sciadv.1500059)
M. G. Schultz, T. S. Nunner, F. von Oppen, Phys. Rev., B 77 (2008) 075323
(https://dx.doi.org/10.1103/PhysRevB.77.075323)
I. B. Bersuker, Phys. Lett. 20 (1966) 589 (https://dx.doi.org/10.1016/00319163(66)91127-9)
P. Garcia-Fernandez, I. B. Bersuker, Phys. Rev. Lett. 106 (2011) 246406
(https://dx.doi.org/10.1103/PhysRevLett.106.246406)
M. Atanasov, J. M. Zadrozny, J. R. Long, F. Neese, Chem. Sci. 4 (2013) 139
(https://dx.doi.org/10.1039/c2sc21394j)
M. Gruden-Pavlović, M. Perić, M. Zlatar, P. Garcia-Fernandez, Chem. Sci. 5 (2014) 1453
(https://dx.doi.org/10.1039/C3SC52984C)
M. Perić, A. García-Fuente, M. Zlatar, C. Daul, S. Stepanović, P. García-Fernández, M.
Gruden-Pavlović, Chem. Eur. J. 21 (2015) 3716
(https://dx.doi.org/10.1002/chem.201405480)
W. Domcke, H. Köppel, L. S. Cederbaum, Mol. Phys. 43 (1981) 851
(https://dx.doi.org/10.1080/00268978100101721)
M. Perić, S. D. Peyerimhoff, J. Mol. Spectrosc. 212 (2002) 153
(https://dx.doi.org/10.1006/JMSP.2002.8534)
M. Perić, S. D. Peyerimhoff, J. Chem. Phys. 102 (1995) 3685
(https://dx.doi.org/10.1063/1.468599)
M. Mitić, R. Ranković, M. Milovanović, S. Jerosimić, M. Perić, Chem. Phys. 464 (2016)
55 (https://dx.doi.org/10.1016/J.CHEMPHYS.2015.11.002)
M. Perić, Mol. Phys. 105 (2007) 59 (https://dx.doi.org/10.1080/00268970601129076)
M. Perić, B. Ostojić, J. Radić-Perić, Phys. Rep. 290 (1997) 283
(https://dx.doi.org/10.1016/S0370-1573(97)00018-5)
M. Perić, S. D. Peyerimhoff, R. J. Buenker, Mol. Phys. 49 (1983) 379
(https://dx.doi.org/10.1080/00268978300101241)
I. B. Bersuker, Spontaneous Symmetry Breaking in Matter Induced by Degeneracies and
Pseudodegeneracies, in Adv. Chem. Physics, Vol. 160, S. A. Rice, A. R. Dinner, Eds.,
Wiley, 2016, pp. 159–208 (https://dx.doi.org/10.1002/9781119165156.ch3)
P. Garcia-Fernandez, I. B. Bersuker, Int. J. Quantum Chem. 112 (2012) 3025
(https://dx.doi.org/10.1002/qua.24204)
Available on line at www.shd.org.rs/JSCS/
________________________________________________________________________________________________________________________
(CC) 2019 SCS.
DFT AND JT EFFECT
797
29. P. García-Fernández, J. A. Aramburu, M. Moreno, M. Zlatar, M. Gruden-Pavlovic, J.
Chem. Theory Comput. (2014) 1824 (https://dx.doi.org/10.1021/ct4011097)
30. Y. Liu, I. B. Bersuker, P. Garcia-Fernandez, J. E. Boggs, J. Phys. Chem., A 116 (2012)
7564 (https://dx.doi.org/10.1021/jp3032836)
31. P. Garcia-Fernandez, J. A. Aramburu, M. Moreno, Phys. Rev., B 83 (2011) 174406
(https://dx.doi.org/10.1103/PhysRevB.83.174406)
32. R. F. W. Bader, Mol. Phys. 3 (1960) 137
(https://dx.doi.org/10.1103/10.1080/00268976000100161)
33. R. F. W. Bader, Can. J. Chem. 40 (1962) 1164 (https://dx.doi.org/10.1139/v62-178)
34. R. F. W. Bader, A. D. Bandrauk, J. Chem. Phys. 49 (1968) 1666
(https://dx.doi.org/10.1063/1.1670293)
35. R. G. Pearson, J. Am. Chem. Soc. 91 (1969) 4947
(https://dx.doi.org/10.1021/ja01046a001)
36. R. G. Pearson, Symmetry Rules for Chemical reactions, A Wiley-Interscience
Publication, New York,1976 (https://dx.doi.org/10.1002/bbpc.19790830620)
37. M. Zlatar, C.-W. Schläpfer, C. Daul, A New Method to Describe Multimode Jahn–Teller
Efect Using Density Functional Theory, in The Jahn–Teller-Effect Fundamentals and
Implications for Physics and Chemistry, H. Koeppel, D. R. Yarkoni, H. Barentzen, Eds.,
Springer Series in Chemical Physics, Vol. 97, 2009, pp. 131
(https://dx.doi.org/10.1007/978-3-642-03432-9_6)
38. A. R. Ilkhani, N. N. Gorinchoy, I. B. Bersuker, Chem. Phys. 460 (2015) 106
(https://dx.doi.org/10.1016/J.CHEMPHYS.2015.07.015)
39. A. R. Ilkhani, W. Hermoso, I. B. Bersuker, Chem. Phys. 460 (2015) 75
(https://dx.doi.org/10.1016/J.CHEMPHYS.2015.02.017)
40. P. Garcia-Fernandez, I. B. Bersuker, J. A. Aramburu, M. T. Barriuso, M. Moreno, Phys.
Rev., B 71 (2005) 184110 (https://dx.doi.org/10.1103/PhysRevB.71.184117)
41. Y. Liu, I. B. Bersuker, W. Zou, J. E. Boggs, J. Chem. Theory Comput. 5 (2009) 2679
(https://dx.doi.org/10.1021/ct9002515)
42. R. G. Parr, W. Yang, Density-Functional Theory of Atoms and Molecules, Oxford
University Press, Oxford, 1989 (https://global.oup.com/academic/product/densityfunctional-theory-of-atoms-and-molecules-9780195092769?cc=rs&lang=en&)
43. C. J. Cramer, D. G. Truhlar, Phys. Chem. Chem. Phys. 11 (2009) 10757
(https://dx.doi.org/10.1039/b907148b)
44. F. Neese, Coord. Chem. Rev. 253 (2009) 526
(https://dx.doi.org/10.1016/j.ccr.2008.05.014)
45. J. P. Perdew, A. Ruzsinszky, Int. J. Quantum Chem. 110 (2010) 2801
(https://dx.doi.org/10.1002/qua.22829)
46. K. Burke, L. O. Wagner, Int. J. Quantum Chem. 113 (2013) 96
(https://dx.doi.org/10.1002/qua.24259)
47. A. D. Becke, J. Chem. Phys. 140 (2014) 18A301 (https://dx.doi.org/10.1063/1.4869598)
48. H. S. Yu, S. L. Li, D. G. Truhlar, J. Chem. Phys. 145 (2016) 130901
(https://dx.doi.org/10.1063/1.4963168)
49. P. Hohenberg, W. Kohn, Phys. Rev. 136 (1964) B864
(https://dx.doi.org/10.1103/PhysRev.136.B864)
50. L. Goerigk, A. Hansen, C. Bauer, S. Ehrlich, A. Najibi, S. Grimme, Phys. Chem. Chem.
Phys. 19 (2017) 32184 (https://dx.doi.org/10.1039/C7CP04913G)
Available on line at www.shd.org.rs/JSCS/
________________________________________________________________________________________________________________________
(CC) 2019 SCS.
798
ZLATAR and GRUDEN
51. I. B. I. B. Bersuker, J. Comp. Chem. 18 (1997) 260
(https://dx.doi.org/10.1002/(SICI)1096-987X(19970130)18:2<260::AIDJCC10>3.0.CO;2-M)
52. I. G. G. Kaplan, J. Mol. Struct. 838 (2007) 39
(https://dx.doi.org/10.1016/J.MOLSTRUC.2007.01.012)
53. I. B. Bersuker, J. Phys. Conf. Ser. 833 (2017) 012001 (https://dx.doi.org/10.1088/17426596/833/1/012001)
54. M. Levy, Phys. Rev., A 26 (1982) 1200 (https://dx.doi.org/10.1103/PhysRevA.26.1200)
55. M. Levy, Int. J. Quantum Chem. 110 (2010) 3140 (https://dx.doi.org/10.1002/qua.22895)
56. A. K. Theophilou, J. Phys., C: Solid State Phys. 12 (1979) 5419
(https://dx.doi.org/10.1088/0022-3719/12/24/013)
57. A. K. Theophilou, P. G. Papaconstantinou, Phys. Rev., A 61 (2000) 022502
(https://dx.doi.org/10.1103/PhysRevA.61.022502)
58. M. Filatov, J. Chem. Theory Comput. 9 (2013) 4526
(https://dx.doi.org/10.1021/ct400598b)
59. R. Requist, E. K. U. Gross, Phys. Rev. Lett. 117 (2016) 193001
(https://dx.doi.org/10.1103/PhysRevLett.117.193001)
60. R. Baer, Phys. Rev. Lett. 104 (2010) 073001
(https://dx.doi.org/10.1103/PhysRevLett.104.073001)
61. J. M. García-Lastra, M. T. Barriuso, J. A. Aramburu, M. Moreno, Chem. Phys. 317
(2005) 103 (https://dx.doi.org/10.1016/j.chemphys.2005.06.004)
62. W. L. Clinton, B. Rice, J. Chem. Phys. 30 (1959) 542
(https://dx.doi.org/10.1063/1.1729984)
63. H. Nakatsuji, J. Am. Chem. Soc. 96 (1974) 30 (https://dx.doi.org/10.1021/ja00808a005)
64. R. Bruyndonckx, C. Daul, P. T. Manoharan, E. Deiss, R. Bruyndockx, C. Daul, P. T.
Manoharan, E. Deiss, Inorg. Chem. 36 (1997) 4251
(https://dx.doi.org/10.1021/ic961220+)
65. T. K. Kundu, R. Bruyndonckx, C. Daul, P. T. Manoharan, Inorg. Chem. 38 (1999) 3931
(https://dx.doi.org/10.1021/ic981111q)
66. M. Gruden-Pavlović, P. García-Fernández, L. Andjelković, C. Daul, M. Zlatar, J. Phys.
Chem., A 115 (2011) 10801 (https://dx.doi.org/10.1021/jp206083j)
67. M. Zlatar, M. Gruden-Pavlović, C. W. Schläpfer, C. Daul, J. Mol. Struct. THEOCHEM
954 (2010) 86 (https://dx.doi.org/10.1016/j.theochem.2010.04.020)
68. L. Andjelković, M. Gruden-Pavlović, M. Zlatar, Chem. Phys. 460 (2015) 64
(https://dx.doi.org/10.1016/j.chemphys.2015.05.007)
69. M. Zlatar, M. Gruden-Pavlović, C.-W. Schläpfer, C. Daul, Chimia 64 (2010) 161
(https://dx.doi.org/10.2533/chimia.2010.161)
70. M. Perić, L. Andjelković, M. Zlatar, C. Daul, M. Gruden-Pavlović, Polyhedron 80 (2014)
69 (https://dx.doi.org/10.1016/j.poly.2014.02.005)
71. L. Andjelkovic, S. Stepanovic, F. Vlahovic, M. Zlatar, M. Gruden, L. Andjelković, S.
Stepanović, F. Vlahović, M. Zlatar, M. Gruden, Phys. Chem. Chem. Phys. 18 (2016)
29122 (https://dx.doi.org/10.1039/C6CP03859J)
72. M. Zlatar, C.-W. Schläpfer, E. P. Fowe, C. Daul, Pure Appl. Chem. 81 (2009) 1397
(https://dx.doi.org/10.1351/PAC-CON-08-06-04)
73. L. Andjelković, M. Gruden-Pavlović, C. Daul, M. Zlatar, Int. J. Quantum Chem. 113
(2013) 859 (https://dx.doi.org/10.1002/qua.24245)
74. M. Atanasov, P. Comba, C. a Daul, A. Hauser, J. Phys. Chem., A 111 (2007) 9145
(https://dx.doi.org/10.1021/jp0731912)
Available on line at www.shd.org.rs/JSCS/
________________________________________________________________________________________________________________________
(CC) 2019 SCS.
DFT AND JT EFFECT
799
75. M. Gruden-Pavlović, M. Zlatar, C.-W. Schläpfer, C. Daul, J. Mol. Struct. THEOCHEM
954 (2010) 80 (https://dx.doi.org/10.1016/j.theochem.2010.03.031)
76. M. Swart, M. Gruden, Acc. Chem. Res. 49 (2016) 2690
(https://dx.doi.org/10.1021/acs.accounts.6b00271)
77. H. Ramanantoanina, M. Zlatar, P. García-Fernández, C. Daul, M. Gruden-Pavlović, Phys.
Chem. Chem. Phys. 15 (2013) 1252 (https://dx.doi.org/10.1039/c2cp43591h)
78. V. P. Khlopin, V. Z. Polinger, I. B. Bersuker, Theor. Chim. Acta 48 (1978) 87
(https://dx.doi.org/10.1007/BF02399020)
79. I. B. Bersuker, V. Z. Polinger, Vibronic interactions in Molecules and Crystals, Springer-Verlag, Berlin, 1989 (https://www.springer.com/gp/book/9783642834813)
80. J. H. Ammeter, L. Zoller, J. Bachmann, P. Baltzer, E. Gamp, R. Bucher, E. Deiss, Helv.
Chim. Acta 64 (1981) 1063 (https://dx.doi.org/10.1002/chin.198141061)
81. F. A. Blankenship, R. L. Belford, J. Chem. Phys. 36 (1962) 633
(https://dx.doi.org/10.1063/1.1732585)
82. R. B. Johannesen, G. A. Candela, T. Tsang, J. Chem. Phys. 48 (1968) 5544
(https://doi.org/10.1063/1.1668254)
83. Y. Morino, H. Uehara, J. Chem. Phys. 45 (1966) 4543
(https://dx.doi.org/10.1063/1.1727535)
84. H. von Busch, V. Dev, H.-A. Eckel, S. Kasahara, J. Wang, W. Demtröder, P. Sebald, W.
Meyer, Phys. Rev. Lett. 81 (1998) 4584
(https://dx.doi.org/10.1103/PhysRevLett.81.4584)
85. J. Gaus, K. Kobe, V. Bonacic-Koutecky, H. Kuehling, J. Manz, B. Reischl, S. Rutz, E.
Schreiber, L. Woeste, J. Phys. Chem. 97 (1993) 12509
(https://dx.doi.org/10.1021/j100150a011)
86. B. E. Applegate, T. A. Miller, J. Chem. Phys. 114 (2001) 4855
(https://dx.doi.org/10.1063/1.1348275)
87. B. E. Applegate, T. A. Miller, J. Chem. Phys. 117 (2002) 10654
(https://dx.doi.org/10.1063/1.1520531)
88. V. Perebeinos, P. B. Allen, M. Pederson, Phys. Rev., A 72 (2005) 12501
(https://dx.doi.org/10.1103/PhysRevA.72.012501)
89. K. Tokunaga, T. Sato, K. Tanaka, J. Chem. Phys. 124 (2006) 154303
(https://dx.doi.org/10.1063/1.2184317)
90. I. Sioutis, V. L. Stakhursky, G. Tarczay, T. A. Miller, J. Chem. Phys. 128 (2008) 084311
(https://dx.doi.org/10.1063/1.2829471)
91. N. Manini, A. Dal Corso, M. Fabrizio, E. Tosatti, Philos. Mag., B 81 (2001) 793
(https://dx.doi.org/10.1080/13642810110062663)
92. H. Ramanantoanina, M. Gruden-Pavlovic, M. Zlatar, C. Daul, Int. J. Quantum Chem. 113
(2013) 802 (https://dx.doi.org/10.1002/qua.24080)
93. T. Yamabe, K. Yahara, T. Kato, K. Yoshizawa, J. Phys. Chem., A 104 (2000) 589
(https://dx.doi.org/10.1021/jp992496g)
94. T. Kato, K. Yoshizawa, J. Chem. Phys. 110 (1999) 249
(https://dx.doi.org/10.1063/1.478100)
95. R. Bucher, ESR-Untersuchungen an Jahn-Teller-Aktiven Sandwitchkomplexen, ETH
Zuerich, 1977 (https://dx.doi.org/10.3929/ethz-a-000150322)
96. S. Stepanović, M. Zlatar, M. Swart, M. Gruden, J. Chem. Inf. Model. 59 (2019) 1806
(https://doi.org/10.1021/acs.jcim.8b00870)
97. J. Tóbik, E. Tosatti, J. Mol. Struct. 838 (2007) 112
(https://doi.org/10.1016/J.MOLSTRUC.2006.12.051)
Available on line at www.shd.org.rs/JSCS/
________________________________________________________________________________________________________________________
(CC) 2019 SCS.
800
ZLATAR and GRUDEN
98. M. Zlatar, M. Gruden-Pavlovic, M. Guell, M. Swart, M. Gruden-Pavlović, M. Güell, M.
Swart, Phys. Chem. Chem. Phys. 15 (2013) 6631
(https://dx.doi.org/10.1039/C2CP43735J)
99. P. Chaudhuri, K. Oder, K. Wieghardt, J. Weiss, J. Reedijk, W. Hinrichs, J. Wood, A.
Ozarowski, H. Stratemaier, D. Reinen, Inorg. Chem. 25 (1986) 2951
(https://dx.doi.org/10.1021/ic00237a007)
100. K. Wieghardt, W. Walz, B. Nuber, J. Weiss, A. Ozarowski, H. Stratemeier, D. Reinen,
Inorg. Chem. 25 (1986) 1650 (https://dx.doi.org/10.1021/ic00230a025)
101. E. Gamp, ESR-Untersuchungen uber den Jahn-Teller-Effekt in oktaedrischen Kupfer(II)Komplexen mit trigonalen dreizahnigen Liganden, ETH, Zurich, 1980
(https://dx.doi.org/10.3929/ethz-a-000215808)
102. A. D. Liehr, Z. Phys. Chem. 9 (1956) 338
(https://dx.doi.org/10.1524/zpch.1956.9.5_6.338)
103. L. C. Snyder, J. Chem. Phys. 33 (1960) 619 (https://dx.doi.org/10.1063/1.1731211)
104. W. D. Hobey, A. D. McLachlan, J. Chem. Phys. 33 (1960) 1703
(https://dx.doi.org/10.1063/1.1731485)
105. R. Meyer, F. Graf, T. Ha, H. H. Gunthard, Chem. Phys. Lett. 66 (1979) 65
(https://dx.doi.org/10.1016/0009-2614(79)80370-X)
106. W. T. Borden, E. R. Davidson, J. Am. Chem. Soc. 101 (1979) 3771
(https://dx.doi.org/10.1021/ja00508a012)
107. C. Cunha, S. Canuto, J. Molec. Struct THEOCHEM 464 (1999) 73
(https://dx.doi.org/10.1016/S0166-1280(98)00536-3)
108. M. J. Bearpark, M. A. Robb, N. Yamamoto, Spectrochim. Acta, A 55 (1999) 639
(https://dx.doi.org/10.1016/S1386-1425(98)00267-4)
109. J. H. Kiefer, R. S. Tranter, H. Wang, A. F. Wagner, Int. J. Chem. Kin. 33 (2001) 834
(https://dx.doi.org/10.1002/kin.10006)
110. S. Zilberg, Y. Haas, J. Am. Chem. Soc. 124 (2002) 10683
(https://dx.doi.org/10.1021/ja026304y)
111. T. Ichino, S. W. Wren, K. M. Vogelhuber, A. J. Gianola, W. C. Lineberger, J. F. Stanton,
J. Chem. Phys. 129 (2008) 084310 (https://dx.doi.org/10.1063/1.2973631)
112. D. Leicht, M. Kaufmann, G. Schwaab, M. Havenith, J. Chem. Phys. 145 (2016) 074304
(https://dx.doi.org/10.1063/1.4960781)
113. A. D. Liehr, J. Phys. Chem. 67 (1963) 389 (https://dx.doi.org/10.1021/j100796a043)
114. N. Iwahara, T. Sato, K. Tanaka, L. F. Chibotaru, Phys. Rev., B 82 (2010) 245409
(https://dx.doi.org/10.1103/PhysRevB.82.245409)
115. J. H. Ammeter, R. Bucher, N. Oswald, J. Am. Chem. Soc. 96 (1974) 7833
(https://dx.doi.org/10.1021/ja00832a049)
116. I. Utke, A. Gölzhäuser, Angew. Chem. Int. Ed. 49 (2010) 9328
(https://dx.doi.org/10.1002/anie.201002677)
117. M. Zlatar, M. Allan, J. Fedor, J. Phys. Chem., C 120 (2016) 10667
(https://dx.doi.org/10.1021/acs.jpcc.6b02660)
118. K. E. R. Marriott, L. Bhaskaran, C. Wilson, M. Medarde, S. T. Ochsenbein, S. Hill, M.
Murrie, Chem. Sci. 6 (2015) 6823 (https://dx.doi.org/10.1039/C5SC02854J).
Available on line at www.shd.org.rs/JSCS/
________________________________________________________________________________________________________________________
(CC) 2019 SCS.