Photophysics and Photochemistry of A BODIPY-Based Photosensitizer
Photophysics and Photochemistry of A BODIPY-Based Photosensitizer
Photophysics and
Photochemistry
of a BODIPY-Based
Photosensitizer
Quantumchemical Simulations
BestMasters
Mit „BestMasters“ zeichnet Springer die besten Masterarbeiten aus, die an renom
mierten Hochschulen in Deutschland, Österreich und der Schweiz entstanden sind.
Die mit Höchstnote ausgezeichneten Arbeiten wurden durch Gutachter zur Veröf
fentlichung empfohlen und behandeln aktuelle Themen aus unterschiedlichen
Fachgebieten der Naturwissenschaften, Psychologie, Technik und Wirtschaftswis
senschaften. Die Reihe wendet sich an Praktiker und Wissenschaftler gleicherma
ßen und soll insbesondere auch Nachwuchswissenschaftlern Orientierung geben.
Springer awards “BestMasters” to the best master’s theses which have been com
pleted at renowned Universities in Germany, Austria, and Switzerland. The studies
received highest marks and were recommended for publication by supervisors.
They address current issues from various fields of research in natural sciences,
psychology, technology, and economics. The series addresses practitioners as well
as scientists and, in particular, offers guidance for early stage researchers.
Photophysics and
Photochemistry
of a BODIPY‐Based
Photosensitizer
Quantumchemical Simulations
Karl Michael Ziems
Selb, Germany
Springer Spektrum
© Springer Fachmedien Wiesbaden GmbH, part of Springer Nature 2019
This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part
of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations,
recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission
or information storage and retrieval, electronic adaptation, computer software, or by similar or
dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this
publication does not imply, even in the absence of a specific statement, that such names are exempt
from the relevant protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this
book are believed to be true and accurate at the date of publication. Neither the publisher nor the
authors or the editors give a warranty, expressed or implied, with respect to the material contained
herein or for any errors or omissions that may have been made. The publisher remains neutral with
regard to jurisdictional claims in published maps and institutional affiliations.
This Springer Spektrum imprint is published by the registered company Springer Fachmedien Wiesbaden
GmbH part of Springer Nature
The registered company address is: Abraham-Lincoln-Str. 46, 65189 Wiesbaden, Germany
Contents
List of Abbreviations vii
Abstract xi
1 Introduction 1
2 Theory 9
2.1 Hamiltonian operator . . . . . . . . . . . . . . . . . . . . . . 9
2.2 Born-Oppenheimer Approximation . . . . . . . . . . . . . . . 10
2.3 Beyond Born-Oppenheimer . . . . . . . . . . . . . . . . . . . 12
2.4 Hartree-Fock theory . . . . . . . . . . . . . . . . . . . . . . . 14
2.4.1 Ansatz of the Hartree-Fock theory . . . . . . . . . . . 14
2.4.2 The Hartree-Fock equations . . . . . . . . . . . . . . . 15
2.4.3 Restricted and Unrestricted Hartree-Fock . . . . . . . 18
2.4.4 The Basis Set Approximation . . . . . . . . . . . . . . 18
2.4.5 The Roothaan-Hall equations . . . . . . . . . . . . . . 19
2.4.6 The SCF Procedure . . . . . . . . . . . . . . . . . . . 20
2.5 Electron Correlation Methods . . . . . . . . . . . . . . . . . . 21
2.5.1 Configuration Interaction . . . . . . . . . . . . . . . . 22
2.5.2 Multi-Configuration Self-Consistent Field . . . . . . . 24
2.6 Density Functional Theory . . . . . . . . . . . . . . . . . . . 32
2.6.1 Hohenberg-Kohn Theorems . . . . . . . . . . . . . . . 33
2.6.2 Kohn-Sham Theory . . . . . . . . . . . . . . . . . . . 33
2.6.3 Exchange-Correlation Functionals . . . . . . . . . . . 35
2.6.4 Time-Dependent Density Functional Theory . . . . . . 37
2.6.5 (TD)DFT Problems . . . . . . . . . . . . . . . . . . . 41
2.7 Spin-Orbit Coupling . . . . . . . . . . . . . . . . . . . . . . . 41
3 Computational Details 45
3.1 Used Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
3.2 Active Spaces for RASSCF/RASPT2 calculations . . . . . . . 48
3.2.1 Active spaces for dissociation and charge transfer . . . 49
vi Contents
6 Conclusion 85
7 Summary 89
8 Zusammenfassung 93
Bibliography 97
Danksagung 139
List of Abbreviations
AMFI Atomic-Mean-Field Integral
AS Active Space
BP Breit-Pauli
CI Configuration Interaction
CT Charge Transfer
EC Electron Correlation
ER Electron Relay
GS Ground State
GTO Gaussian Type Orbital
viii List of Abbreviations
HF Hartree-Fock
HOMO Highest Occupied Molecular Orbital
H-L HOMO → LUMO
H-1-L HOMO-1 → LUMO
IC Internal Conversion
ISC Intersystem Crossing
LCAO Linear Combination of Atomic Orbitals
LDA Local Density Approximation
LICC Linear-Interpolated Cartesian Coordinate
LSDA Local Spin Density Approximation
LUMO Lowest Unoccupied Molecular Orbital
MCSCF Multi-Configuration Self-Consistent Field
MEP Minimum Energy Path
MO Molecular Orbital
MP Møller-Plesset Perturbation Theory
MS-CASPT2 Multi-State Complete Active Space Perturbation Theory
through Second-Order
MS-RASPT2 Multi-State Restricted Active Space Perturbation Theory
through Second-Order
PCM Polarizable Continuum Model
PES Potential Energy Surface
PMD Photochemical Molecular Devices
PS Photosensitizer
RASPT2 Restricted Active Space Perturbation Theory through
Second-Order
RASSCF Restricted Active Space Self-Consistent Field
List of Abbreviations ix
SD Slater Determinant
SOC Spin-Orbit Coupling
SOMF Spin-Orbit Mean Field
2 H2 O −−→ 2 H2 + O2 (1.1)
+ −
2 H2 O −−→ 4 H + 4 e + O2 (1.2)
+ −
4 H + 4 e −−→ 2 H2 (1.3)
functions as electron mediator and donates the electron to the catalyst. Here,
the water splitting to hydrogen occurs. The two-component system omits
the electron relay and the excited electron from the PS is transferred directly
to the catalyst. [29–33] A major problem with these systems is controlling and
monitoring the electron transfer processes, which depend on many factors in
heterogeneous systems. Thus, supramolecular, homogeneous, one-component
approaches have been the focus of research in recent years. They couple a PS
as light-harvesting unit with a catalytic center in a supramolecular approach
- commonly referred to as photochemical molecular devices (PMD). [34–39]
One prominent example is [(tbbpy)2 Ru(tpphz)PdCl2 ](PF6 )2 (tbbpy = 4,4’-
Di-tert-butyl-2,2’-bipyridin, tpphz = Tetrapyrido-[3,2-a:2’,3’-c:3”,2”’,3”’-
j]phenanzin) shown in Figure 1.1, [40,41] where the Ruthenium moiety acts as
PS and transfers one of several excess charges via MLCT excitation to the
tpphz bridging ligand and from there to the catalytic active Palladium cen-
ter. [34,41] The design concept underlying all heterogeneous systems discussed
so far is depicted in Figure 1.2. [5]
The preliminary step, the energy supply for these processes, is achieved
via light harvesting by the PS through population of an electronic excited
state. Such a PS is defined as a molecule with a chromophor absorbing
in the ultraviolet (UV) and visible (Vis) range of the solar spectrum, and
capable of transferring electrons to an electronic acceptor. An efficient
PS is distinguished by a high absorbency in a broad range of the UV/vis
part of the solar light, long-term stability, its electron-transfer efficiency
and excited state lifetime, all with respect to cost, productivity, toxicity
and feasibility of hydrogen generation. [42] The most prominent example
of a PS in nature is chlorophyll, used for photosynthesis in photosystem I
and II. [34,35] Thus, man-made light-induced hydrogen-producing systems
N
N N N N Cl
Ru2+ Pd
N N N N Cl
N
SD H2
(1) PS + ER + Cat
SD+ H2O
SD H2
(2) PS + Cat
SD+ H2O
SD H2
(3) PS Cat
SD+ H2O
R
Cl PPh3
Pd
X1 X2 Cl Cl
N N N
Pd
B
Ph3P Cl
F F
PS Pd source TEA
The aim of this thesis is to investigate the BODIPY dye shown in Fig-
ure 1.5. The structure was slightly modified compared to the Reference [86] by
replacing the mesityl group with a phenyl group to reduce the computational
costs of the quantum chemical simulations. Several tasks are outlined, based
on the experimental observations and mechanistic thoughts described above.
First of all, a comparison of the results obtained by different levels of theory
for the BODIPY has to be conducted. This is especially important since the
calculations from the previous work were mainly done by TDDFT, which is
known to perform poorly for boron-containing systems2 . As state-of-the-art
computational reference method Multi-State Restricted Active Space Per-
turbation Theory through Second-Order (MS-RASPT2) has been chosen.
The main goal is to unravel the photo-induced mechanism (photoactivation)
leading to hydrogen production, i.e. heavy atom effect (non-reduced system)
vs. chemical reduction (singly reduced system), and to study the C-I cleavage
(photodegradation) in order to prevent such cleavage and to enhance the
dye’s photostability. Therefore, pathways and kinetics for dissociative and
CT processes have to be identified and an investigation of the role of the
Iodines and a possible relativistic heavy atom effect has to be incorporated.
Figure 1.5: Structure of T. Beweries et al. most promising BODIPY dye re-
garding hydrogen production in a two-component system. The
main heteroaromatic ring and the side ring are perpendicular to
each other and the structure has been slightly modified substitut-
ing the mesityl with a phenyl side ring to reduce computational
costs.
2 Theory
The following chapter presents an overview of the used methods, as well as
the underlying physical principles and approximations.
here T̂n is the nuclear kinetic energy operator, T̂e the electronic kinetic energy
operator and V̂ = V̂ne + V̂ee + V̂nn the corresponding potential energy operator
of the interactions of electron and nucleus, V̂ne , electron and electron, V̂ee ,
and nucleus and nucleus, V̂nn . The parameter me and mα describe the
electron- and nuclear-mass, respectively, e corresponds to the elementary
charge, Zα to the nuclear charge and 0 is the vacuum permittivity. The
distances between the elementary particles are indicated by r and R. [89,90]
In this case, Ĥe is the electronic Hermitian Hamiltonian and Ĥmp the
so-called mass-polarization. The latter denotes that it is impossible to
strictly separate the center of mass motion from the internal motion. This
contribution to the total Hamiltonian will be neglected from now on.1
The Born-Oppenheimer Ansatz (without introducing any approximations
yet) is an expansion of the total, exact wave function in a complete set of
electronic wave functions |ψi (r, R), with the expansion coefficient being
functions of the nuclear coordinates |ψni (R).
∞
=
|Ψ(r, R) |ψi (r, R)
|ψni (R) (2.6)
i=1
This means that the total wave function is separated in a term, |ψni (R),
depending on the nuclear coordinates (nuclear wave function), and a term,
|ψi (r, R), depending on the electronic coordinates and the current nuclear
position R (electronic wave function).2
Equation 2.5 and 2.6 allow to define an electronic Schrödinger equation:
= Ei (R)
Ĥe |ψi (r, R) |ψi (r, R)
; i = 1, 2, ..., ∞ (2.7)
1 The adiabatic and BOA both neglect the Ĥmp term. For the sake of clarity, it will
already be neglected in the derivation of these approximations.
2 Thus, the index i sums over the electronic eigenfunctions |ψ (
i r , R), whereas n indicates
the nuclear eigenfunctions |ψni (R) for a given electronic eigenfunction |ψi (
r, R).
2.2 Born-Oppenheimer Approximation 11
Combining the TISE in Equation 2.2 with the separation of the total wave
function in Equation 2.6 gives:
∞
∞
|ψi (r, R)
(Ĥe + T̂n ) |ψni (R) = Etot |ψi (r, R)
|ψni (R) . (2.9)
i=1 i=1
For reasons of clarity, from now on the dependencies of the wave functions
will be left out and atomic units will be used for the operators. Moreover,
the electronic wave function is written as ψi and the nuclear wave function
as ψni .3
K
h̄2 2
∇2n = T̂n = − ∇
α=1
2mα α
ψni = |ψni (R)
ψi = |ψi (r, R) (2.10)
Inserting the exact expression for the nuclear operator and using Equation 2.7,
results after some conversions in
∞
∞
ψi (∇2n ψni ) + 2(∇n ψi )(∇n ψni )+
= E tot ψni ψi . (2.11)
i=1
ψni (∇2n ψi ) + ψni Ei ψi i=1
Multiplying the equation from the left by a specific electronic wave function
ψj∗ and integrating over the electronic coordinates yields, with the use of the
orthonormality of the electronic wave function (Equation 2.8), the following
equation:
∞
∞
2 ψj |∇n |ψi (∇n ψni )+
2
∇n ψnj + Ej ψnj + = E tot ψnj . (2.12)
i=1
ψj |∇2n |ψi ψni i=1
3 The notation in bras and kets will also be neglected. The braket notation will only
be used to indicate scalar products (integration over electronic coordinates) in this
section.
12 2 Theory
The terms in the curly bracket couple different electronic states and are called
first- and second-order non-adiabatic coupling elements. These become very
important at the (near-)degeneracy of electronic states and for describing
interactions between them, e.g. for photochemical reactions.
In the adiabatic approximation, the total wave function is restricted to
one electronic surface. This means that in Equation 2.12 only the coupling
elements i = j survive and the first-order non-adiabatic elements are zero
(within the assumption of no spatial degeneracy).
ψnj (R)
T̂n + Ej (R) = Etot ψnj (R)
(2.14)
Also, because two (or more) states with different electronic distributions
become energetically equal, the electronic wave function cannot adjust
instantly on any change in the nuclear motion and therefore, an adiabatic
approach is not reasonable. As a result of this, concepts such as molecular
geometries become blurred and energy surfaces no longer exist.
Calculating such PES with crossing states in the framework of the BOA
would result in adiabatic potentials which are described poorly in the crossing
region and does not allow any switch of state. This is based on the break-
down of the main approximation within the BOA, the neglection of the
non-adiabatic coupling elements:
(1)
Tij = ψj |∇n |ψi
(2)
Tij = ψj |∇2n |ψi , (2.15)
(1)
T T12 E1 0 ψn1 ψn1
(1) + = Etot (2.17)
T12 T 0 E2 ψn2 ψn2
(d) (d)
T 0 E1 E12 ψn1 ψn1
+ (d) (d) = Etot (2.18)
0 T E12 E2 ψn2 ψn2
14 2 Theory
Ψ|Ĥe |Ψ
E= . (2.19)
Ψ|Ψ
In the HF method, the trial wave function is built by one-electron functions
- Molecular Orbitals (MO). Since the electronic wave function must be
antisymmetric with respect to interchange of any two electron coordinates
(Pauli principle), these orbitals are organized in a Slater Determinant (SD).
As another Ansatz in the HF theory, the electron spin is introduced as
an ad hoc quantum effect. This results in orthonormal spin orbitals χa (xi )
as product of a spatial orbital ϕa (ri ) and a spin function s(ωi ) with two
possible arrangements (α(ωi ) or β(ωi )).
α(ωi )
χa (xi ) = ϕa (ri ) · (2.20)
β(ωi )
Here, the index a specifies the MO and i the electron. The spin functions
are orthogonal to each other:
α|α = β|β = 1
α|β = β|α = 0 . (2.21)
2.4 Hartree-Fock theory 15
1 Zα N
with ĥ(xi ) = − ∇2i − (2.24)
2 α
riα
1
and ĝ(xi , xj ) = (2.25)
rij
The one-electron operator ĥ(xi ) describes the kinetic energy of the electron
i in the field of all nuclei, and ĝ(xi , xj ), the two-electron operator, gives the
electron-electron repulsion. Applying these operators to the HF-Ansatz for
the electronic energy as expectation value of a N electron, N orbital SD,
results in Equation 2.26
16 2 Theory
N
E= χa (x1 )|ĥ(x1 )|χa (x1 )
a
1
N N
+ χa (x1 )|Jˆb (x1 )|χa (x1 ) − χa (x1 )|K̂b (x1 )|χa (x1 ) + V̂nn
2 a
b=a
N
1
N N
= haa + (Jab − Kab ) (2.26)
2 a
a b=a
with the Coulomb operator Jˆb (x1 ) and the exchange operator K̂b (x1 )
Jˆb (x1 )χa (x1 ) = χb (x2 )|ĝ(x1 , x2 )|χb (x2 ) χa (x1 ) (2.27)
K̂b (x1 )χa (x1 ) = χb (x2 )|ĝ(x1 , x2 )|χa (x2 ) χb (x1 ) (2.28)
N
δL = δE − λab (δχa |χb − χa |δχb ) = 0 (2.31)
a,b
4 The integrations are over a single or two electrons. The dummy variables of integration
are, by convention, chosen to be the coordinates of electron one, and electron one and
two, respectively.
2.4 Hartree-Fock theory 17
N
fˆ(x1 )χa (x1 ) = λab χb (x1 ) (2.32)
b
N
fˆ(x1 ) = ĥ(x1 ) + Jˆb (x1 ) − K̂b (x1 ) , (2.33)
b
describing the kinetic energy of an electron and the attraction to all nuclei
via the one-electron operator ĥ(x1 ), as well as the repulsion to a (average)
sum of all other electrons via the Coulomb and exchange operator.5 After a
unitary transformation of the Hartree-Fock equation that diagonalizes the
matrix of Lagrange multipliers λab , a new HF equation with a special set of
so-called canonical orbitals χ and their orbital energies a is obtained:
N
1
N
E= a − (Jab − Kab ) + V̂nn (2.35)
a
2
a,b
5 Note that the Fock operator is an operator for the variational HF method and therefore,
associated with the variation of the energy. It is not associated with the energy itself.
Thus, the Hamilton operator is not a sum of Fock operators.
6 The Coulomb and exchange operator sum over all orbitals.
7 The prime in the HF solution (Equation 2.34) will be neglected from now on.
18 2 Theory
Inserting Equation 2.37 into the general spin orbital Hartree-Fock equa-
tion (Equation 2.34) and integrating out the spin functions gives the RHF
equation:
fˆ(r1 )ϕa (r1 ) = a ϕa (r1 ) (2.39)
and operators not depending on the spin coordinate. Furthermore, one
contribution of the exchange operator is lost due to the fact that it only
occurs between electrons of the same spin:
N/2
fˆ(r1 ) = ĥ(r1 ) + 2Jˆb (r1 ) − K̂b (r1 ) . (2.40)
b
K
K
fˆ(r1 ) caμ φμ (r1 ) = a caμ φμ (r1 ). (2.44)
μ μ
20 2 Theory
Multiplying from the left by a specific basis function φ∗ν and integrating over
the electronic coordinates results in the Roothaan-Hall equations, being the
Hartree-Fock equations in the atomic orbital basis:
K
K
Fνμ Cμa = a Sνμ Cμa . (2.45)
μ μ
Equation 2.45 reduces the problem of calculating the HF MOs to the problem
of calculating a set of expansion coefficients Cμa , which can be derived by
diagonalizing8 the K × K Hermitian Fock- and overlap-matrix, Fνμ and Sνμ ,
in the set of basis functions.
− φν (r1 )φλ (r2 )|ĝ(r1 , r2 )|φσ (r1 )φμ (r2 )] (2.48)
4. Obtain a guess for the density matrix P , e.g. through extended Hückel
method. [102]
5. Calculate the two-electron matrix G from the density matrix and the two-
electron integrals and the entire Fock matrix by adding the one-electron
term (Equation 2.48).
6. Calculate the transformed Fock matrix F = X † F X.
10. Determine if the convergence criteria are fulfilled. If not, start at step (5)
with the new density matrix P .
Solving the SCF algorithm (for a closed shell system) produces a total of
K MOs with N/2 occupied MOs and K − (N/2) unoccupied or virtual
MOs. [89,90]
ar a<b a<b<c
r<s r<s<t
= c0 |ΨHF + cS |S + cD |D + cT |T + · · · (2.51)
S D T
The wave function |Ψra , or |S in short form, describes a SD where one
electron in an occupied orbital a has been excited to a virtual orbital r with
respect to the |ΨHF reference. The other terms in Equation 2.51 describe
doubly and triply excited determinants.
Similar to the HF derivation, the energy of the CI wave function is mini-
mized under the constraint that the total wave function |ΨCI is normalized:
(Hij − EIij ) ci = 0
Hij ci = Eci (2.54)
⎛ ⎞⎛ ⎞ ⎛ ⎞
H00 − E H01 ··· H0j ··· c0 0
⎜ H11 − E ··· · · ·⎟ ⎟ ⎜0⎟
⎜ H10 H1j ⎟⎜ c ⎟ ⎜ ⎟
⎟⎜
1
⎜ .. .. .. .. ⎜ .. ⎟ ⎜ .. ⎟
⎜ · · ·⎟
⎜ . . . . ⎟⎜ ⎟
⎜ . =⎜
⎜.⎟
⎟ (2.55)
⎜ .. ⎟⎜ ⎟ ⎟ ⎜ ⎟
⎜ · · · Hjj − E · · ·⎟ c 0
Hj0 . ⎠ ⎝ . ⎠ ⎝.⎠
j
⎝
.. .. .. ..
··· ··· ··· . .
If two SDs differ by more than two (spatial) MOs, the matrix elements
are zero, as well; this connection is known as Slater Condon rule. [106,107]
Therefore, the main contribution for evaluating the ground state with CI are
⎛ ⎞
ΨHF |Ĥ|ΨHF 0 ΨHF |Ĥ|D 0 0 ···
⎜ S|Ĥ|S S|Ĥ|D S|Ĥ|T · · ·⎟
⎜ 0 ⎟
⎜ ⎟
⎜ D|Ĥ|D D|Ĥ|T D|Ĥ|Q · · ·⎟
⎜ ⎟
⎜ T |Ĥ|T T |Ĥ|Q · · ·⎟
⎜ ⎟
⎜ Q|Ĥ|Q · · ·⎟
⎝ ⎠
..
.
(2.57)
If all possible excited SDs are considered, the wave function ΨCI is called
full CI wave function. Though, the factorial growth of the number of
determinants with the size of the basis set and the size of the molecule makes
the full CI method unfeasible for all but the very smallest systems.
Therefore, truncated CI methods are preferable. These methods stop the
linear combination of excited SDs after a specific excitation level. Whereas
CIS (truncating after singlet excited SDs) does not introduce any improve-
ment to the HF wave function due to the Brillouin theorem, CISD recovers
80-90% of the correlation energy. Incorporating more excitation levels, re-
sults in new methods, like CISDT, CISDTQ, etc. The more levels are used,
the higher the computational costs are. The main problem of all truncated
methods is the size inconsistency, which is rather disadvantageous especially
for the description of dissociation or reaction pathways. [89,90,103,108]
11 The Hermitian character is expressed in form of only one depicted half of the symmetric
CI matrix.
2.5 Electron Correlation Methods 25
|ΨMCSCF = ci |Ψi (2.58)
i
|Ψi = |χ1 χ2 · · · χN −1 χN = |ϕ1ϕ1 · · · ϕaϕa · · · ϕN/2 ϕN/2 (2.59)
K
|ϕa = bμa φμ (2.60)
μ
1
Ĥ = hij Êij + gijkl Êij Êkl − δjk Êil . (2.62)
ij
2
i,j,k,l
The MCSCF matrix consists of one-electron hij and two-electron gijkl ma-
trices containing all information about the MOs and their coefficients. The
density matrices Dij and Pijkl contain the CI coefficients and, thus, are
coupling the individual SDs.12
MCSCF methods are - contrary to the CI methods discussed above - not
designed for obtaining a large fraction of dynamic correlation. They are
rather used for systems with a large contribution of static electron correlation,
which in general includes every system in which more than one non-equivalent
resonance structure is important.13 The different structures/configurations
|Ψi , represented by optimized SDs14 , generate the multiconfigurational wave
function |ΨMCSCF . Examples for systems with such a multiconfigurational
12 This describes a solution for a more-determinant system like seen in the section of
Configuration Interaction, but in this case through a second quantification approach.
13 This means that even the ground state wave function cannot be described qualitatively
N
Eav = N −1 Ei . (2.63)
i
State average calculations are useful to describe near degeneracy states (of
the same symmetry) or even conical intersections. [103]
(0)
By multiplication from the left with a specific wave function ψi |, the
energy and wave function corrections of specific order n are obtained. A
general solution for the energy is:
(n) (0) (n−1)
Ei = ψi |Ĥ |ψi . (2.74)
the explicit first, second- and third-order energy correction, as well as the
first-order wave function coefficients can be obtained:
(1) (0) (0)
Ei = ψi |Ĥ |ψi (2.76)
(0) (0)
(1) ψj |Ĥ |ψi
cj = (0) (0)
(2.77)
Ei − Ej
(2)
(1) (0) (0)
ψi(0) |Ĥ |ψj(0) ψj(0) |Ĥ |ψi(0)
Ei = cj ψi |Ĥ |ψj = (0) (0)
. (2.78)
j j Ei − Ej
Applying this on Equation 2.78 results in MP2, which mixes excited de-
terminants with the HF reference. Because of the Slater Condon rule and
Brillouin theorem, only doubly excited SDs have to be taken into account.
This yields a dynamic energy correlation correction. [89,90,113]
occ. virt.
ψHF |Ĥ |ψ rs ψ rs |Ĥ |ψHF
E (2) = ab ab
(2.81)
EHF − Eab
rs
a<b r<s
30 2 Theory
CASPT2 formalism For CASPT2 the same principle applies as for MP2.
The energy correction in second-order relies on the first-order wave function
and is constructed using doubly excited determinants generated from the
CASSCF reference wave function. But only doubly excited SDs that are not
already in the CASSCF reference are used.15
Therefore, the first-order wave function is written as:
(K) (K)
|Ψ(1) = tP X̂P |Ψ(0) (2.82)
KP
(K)
X̂P ∈ Êvjtu , Êvjti , Êatvx , Êaivx , Êvjai , Êbvat , Êbjat , Êbjai , (2.83)
15 In general, it is requested that the n-th order wave function lies in the n-th order
interacting space. This space is defined by the residual wave function
ρ = Ĥ − E0 |Ψ(0) ,
which measures the error of the reference to the exact wave function Ψ and is spanned
by all SDs that are not in the reference, but differ by at most two-electron excitation
from at least on reference SD.
2.5 Electron Correlation Methods 31
Table 2.1: Labeling and contributions for the configurations in CASPT2 from
the different excitation operators Epqrs . P and M describes the
positive and negative linear combination.
Config. Excitation 1
VJTU Inactive (J) → Active (V)
VJTIP Inactive (J) → Active (V)
VJTIM Inactive (J) → Active (V)
ATVX Active (T) → Secondary (A)
AIVX Inactive (I) → Secondary (A)
or: Active (X) → Secondary (A)
VJAIP Inactive (J) → Active (V)
VJAIM Inactive (J) → Active (V)
BVATP Active (V) → Secondary (B)
BVATM Active (V) → Secondary (B)
BJATP Inactive (J) → Secondary (B)
BJATM Inactive (J) → Secondary (B)
BJAIP Inactive (J) → Secondary (B)
BJAIM Inactive (J) → Secondary (B)
Config. Excitation 2
VJTU Active (U) → Active (T)
VJTIP Inactive (I) → Active (T)
VJTIM Inactive (I) → Active (T)
ATVX Active (X) → Active (V)
AIVX Active (X) → Active (V)
or: Inactive (I) → Active (V)
VJAIP Inactive (I) → Secondary (A)
VJAIM Inactive (I) → Secondary (A)
BVATP Active (T) → Secondary (A)
BVATM Active (T) → Secondary (A)
BJATP Active (T) → Secondary (A)
BJATM Active (T) → Secondary (A)
BJAIP Inactive (I) → Secondary (A)
BJAIM Inactive (I) → Secondary (A)
The two states α and β are coupled through the perturbation δ and are
separated by a level shift γ. In the whole, exact Hamiltonian H = H0 + H
the shift γ is gone and, therefore, does not influence the energy. However,
benchmark calculations have shown that the level shift cannot be removed
32 2 Theory
ρ(
r1 ) = N ··· |Ψ (
r1 , ω1 , rN , ωN ) |2 dω1 , d
r2 , ω2 , ..., r2 , dω2 , ..., d
rN , dωN
(2.89)
Equation 2.89 gives the probability of finding an electron with arbitrary spin
in the volume element r1 . The main advantage is the dependency on only
three coordinates regardless of the size of the system.
2.6 Density Functional Theory 33
Theorem I
The first theorem proofs that the ground state electronic energy is determined
completely by the electron density ρ0 :
In the definition of E.B. Wilson [120] this means that the integral of the
density defines the number of electrons, the cusps in the density the position
of the nuclei and the height of the cusps the corresponding nuclear charges.
Theorem II
The expectation value of the Hamiltonian (Equation 2.90) follows the varia-
tional principle.
E[ρ0 ] ≤ E[ρ ] (2.91)
An arbitrary test density ρ can never result in a smaller energy than the
real electron density of the ground state ρ0 .
The Ene [ρ] and J[ρ] functionals can be expressed easily in terms of the
electron density.
K α )ρ(r)
Zα (R
Ene [ρ] = − dr (2.94)
α − r|
|R
α
1 ρ(r)ρ(r )
J[ρ] = drdr (2.95)
2 |r − r |
However, the kinetic functional with a second derivative of the wave function
and the exchange functional cannot be expressed in dependence on the
electron density. [121]
Kohn-Sham Ansatz
A solution to this is given by Kohn and Sham, who suggest a division of the
electron kinetic energy into two parts. The first part TS [ρ] calculates an exact
expression for the energy with the assumption of non-interacting electrons,
whereas the second part is a small correction term TC [ρ] = T [ρ] − TS [ρ]. The
exact expression is gained through the reintroduction of orbitals in DFT.
Therefore, a Slater determinant composed of molecular orbitals ϕa (ri ) is
used to give the exact kinetic energy functional:
N
1
TS = ϕa (r1 )| − ∇2 (r1 )|ϕa (r1 ) , (2.96)
a
2
(KS)
N
ρ0 (r) = |ϕa (ri )|2 . (2.97)
a
The remaining kinetic energy TC [ρ] is combined with the exchange functional
and the missing electron correlation to an exchange-correlation term Exc [ρ]:
the various DFT methods (see. subsection 2.6.3) aim to find an approximate
solution.
Even though, the reintroduction of orbitals increases the coordinates from
3 to 3N , the computational costs are similar to HF theory and the results
are much better. And once the exchange-correlation functional has been
selected, the computational problem is very similar to HF. The energy is
optimized with orbitals ϕa (ri ), constrained to orthogonality by the Lagrange
method. Using the variational principle on the Lagrange function, like in
HF, gives an effective one-electron operator:
1 2
ĥKS (r1 ) = ∇ (r1 ) + V̂eff (r1 ) (2.100)
2
V̂eff (r1 ) = V̂ne (r1 ) + V̂J (r1 ) + V̂xc (r1 ) (2.101)
δExc [ρ] δxc (r2 )
V̂xc (r1 ) = = xc [ρ] + ρ(r2 ) dρ(r2 ), (2.102)
δρ(r1 ) δρ(r1 )
17 Every quantum chemical program has a default grid setting, which can be adjusted.
2.6 Density Functional Theory 37
Hybrid Methods
Hybrid methods use an exact expression for the exchange functional via HF
exchange potential. This is done by using the already present Kohn-Sham
orbitals, which results in very accurate functionals, applicable to a lot of
chemical systems.
In general, the exact exchange is added to expressions from other func-
tionals. E.g. the PBE0 functional consists of the GGA PBE functional and
exact HF exchange: [132]
PBE0 1 HF 3 PBE
Exc = E + Ex + EcPBE . (2.104)
4 x 4
The exact amount of HF exchange potential is different for each functional.
Examples for hybrid functionals are, besides PBE0 (25% HF exchange),
B3LYP (20% HF exchange) and TPSSh (10% HF exchange). [132–136]
N
ρ(KS) (r, t) = |ϕa (r, t)|2 . (2.109)
a
δSxc [ρ]
Vxc [ρ, Ψ0 , ΨKS ](r, t) = (2.113)
δρ(r, t)
with Ωn = En − Egs . The equation above shows that the resonant part of the
response function has poles at the exact excitation energy of state n of the
system. The remaining numerator at the poles represents transition moments
between the ground and excited state n. Simply put, if a perturbation
potential is applied and its frequency matches one of the excitation energies,
the response of the system in the frequency domain is very large, seen as a
peak in the spectrum.
In TDDFT, the linear density response is calculated as the response of the
(KS)
time-independent, non-interacting Kohn-Sham system ρ0 (Equation 2.97)
to an effective, time-dependent, perturbation potential (see Equation 2.112)
with a monochromatic dipole field along z direction as external potential:
20 The Boron atom has the electron configuration 1s2 2s2 2p1 with a single p-electron
that can be distributed in three different, but energetic degenerated p-orbitals. This
multiconfigurational property leads to problems in (TD)DFT.
42 2 Theory
and also for light elements and impacts photophysical and photochemical
properties. [151–153]
The exact, but also most expensive way, to implement relativistic ef-
fects would be the four-component Dirac equation [154,155] instead of the
Schrödinger equation. A common kind of simplification, including the preser-
vation of the Schrödinger equation as basis of all calculations, is given by
the Douglas-Kroll-Hess (DKH) transformation. [156] It transforms the Dirac
equation in a spin-independent scalar-Hamiltonian and a spin-dependent
SOC-Hamiltonian.21 [151,152,157]
The spin-independent scalar-relativistic effect is usually implemented
in form of a correction to the one-electron integrals and/or as Effective
Core Potential (ECP). For the spin-dependent SOC-Hamiltonian different
approaches are available. The most common one is the Breit-Pauli (BP)
Ansatz with the corresponding BP operator defined as follows: [151,152,158,159]
SOC (1) (2)
ĤBP = ĤBP + ĤBP (2.120)
(1)
α
ĤBP = ĥ1el−SOC
i = −3 ˆ
Zα riα liα ŝi (2.121)
i
2 i α
(2) (2) (2)
ĤBP = ĤSSO + ĤSOO (2.122)
α
2el−SOC −3 ˆ 2 −3 ˆ
= gij =− Zα rij lij ŝi − α rij lij ŝj .
2 i
i j=i j=i i j=i
repulsion in HF. The SOC between two states is calculated by expressing the
two states as two orthonormal SDs with biorthonormal MOs, which differ
only by an excitation χa → χr .22 This offers an expression of the SOC energy
correction in terms of one-electron and two-electron integrals. [151,152,163–165]
Ψ1 |ĤSOC |Ψ2 (2.123)
= χa |ĥ1el−SOC |χr (2.124)
2el−SOC 2el−SOC
+ (χa χj |g12 |χr χj + χj χa |g12 |χj χr (2.125)
j
2el−SOC 2el−SOC
− χa χj |g12 |χj χr − χj χa |g12 |χr χj ) (2.126)
The two electron integrals can be divided into a Coulomb (SOO) part
(Equation 2.125) and an exchange (SOO) part (Equation 2.126).
The aim is now to generate an effective one-electron SOC operator, com-
parable to the Fock operator in HF. This means a separation of the spin
and orbital angular momentum part:
ĤSOC = ˆli ŝi . (2.127)
i
The one-electron part (Equation 2.124) is already in the desired form. The
two-electron parts can be separated by the following mechanism:
2el−SOC
χa χb |ĝ12 |χr χs = (ϕa ϕb |ĝ SOC |ϕr ϕs ){δsb ss sa |ŝ|sr +2δsa sr sb |ŝ|ss } .
(2.128)
Here, s(ω) represents the spin function with spin α or β, whereas the orbital
angular momentum two-electron operator is given as:
α2 ˆ −3
ĝ SOC (i, j) = −
lij rij . (2.129)
2
The outcome of these conversions allows a treatment of SOC as effective one-
electron operator. [163,164] Qualitatively speaking, this results in a notation
of:
Ψ1 |ĤSOC |Ψ2 (2.130)
= ϕa sa (ω)|ĤSOC |ϕr sr (ω)
= ϕa |ˆli |ϕr sa |ŝi |sr .
i
22 Thisis the Restricted Active Space State Interaction (RASSI) implemented in e.g.
MOLCAS. [162] There is a variety of operator that can be chosen for the interaction
between the states.
44 2 Theory
Equation 2.130 shows that depending on the symmetry of the system, SOC
does not occur between all states of a given spatial symmetry.23 [166]
Another implementation with two additional but reasonable approxi-
mations is Schimmelpfennig’s Atomic-Mean-Field Integral (AMFI) pro-
gram. [165,167] The approximations are that: i) only one-center terms are
retained, and ii) the "mean-field-orbitals" are replaced by atomic SCF or-
bitals obtained from spherically averaged atomic SCF calculations with pre-
determined valence shell occupations, instead of the complete non-spherical
molecular density.
gets stabilized by 0.22 eV and loses two imaginary frequencies. The remaining
two vibrations are shown in Figure 3.2.
In conclusion, the negative frequencies were tolerated due to the stated
reasons above and a C2v model system introduced to lower the computational
costs and make the system applicable for state-of-the-art multiconfigurational
methods. In later sections, this phenomenon is revisited and will be discussed
in detail.
For the solvent calculation, the PCM was applied, which utilizes the
integral equation formalism (IEFPCM) and considers a solute cavity out of
MCSCF
All MCSCF calculations, namely (MS-)RASPT2 and (SA-)RASSCF, were
calculated with Molcas 8.0 service pack 1 [161] and the ano-rcc-vdzp [175–180]
basis set.
Relativistic ano-type basis sets include a priori scalar-relativistic effects;
therefore, the integrals are calculated with a DKH transformation of one-
electron integrals and correlation of semi-core orbitals. Furthermore, this
basis set was studied and optimized for SOC and the use in the AMFI ap-
proximation. Thus, ano-rcc-type basis sets are ideal for incorporating scalar-
and SO-relativistic effects. All these concepts were thoroughly discussed in
section 2.7. [175–180]
For speed-up, all calculations were subjected to the Cholesky decompo-
sition of the two-electron integrals with the default threshold of 10−4 a.u. [181]
48 3 Computational Details
MEP
The Minimum Energy Path (MEP) was obtained by MOLCAS 8.2 [161] on
the (single state) RASSCF level of theory with the ano-rcc-vdzp basis set, the
Cholesky-decomposition, analytical gradients and applying C2v symmetry.
A MEP is the steepest descendant minimum energy reaction path on the
PES starting from a given geometry in an initial state. The path is built
through a series of geometry optimizations, each requiring the minimization
of the potential energy on a hyperspherical cross-section of the PES and a
predefined hyperspherical radius from the previous optimization step (or the
initial structure for the first step). [184,185]
Table 3.1: Different RAS2 distributions and their CSFs for different Active
Spaces (AS)
AS RAS 2 CSFs
1 σ(a1 ), σ(b1 ), σ ∗ (a1 ), σ ∗ (b1 ), π3 (a2 ), π4∗ (b2 ) 222869
2 σ ∗ (a1 ), σ ∗ (b1 ), π2 (a2 ), π3 (a2 ), π4∗ (b2 ) 71140
3 σ ∗ (a1 ), σ ∗ (b1 ), π2 (a2 ), π3 (a2 ), π4∗ (b2 ), π5∗ (b2 ) 122948
4 π2 (a2 ), π3 (a2 ), π4∗ (b2 ), πph,2 (b1 ), πph,3 (a2 ), πph,4
∗ ∗
(a2 ), πph,5 (b1 ) 534332
50 3 Computational Details
HF
Figure 3.3: All orbitals included in the RAS. The RAS distribution differs
in AS1-AS4, recall Table 3.1. Dashed line (HF) labels the MOs
occupied in the electronic singlet ground state. Orbital symmetry
is labeled by a1 , b1 , a2 , b2 .
3.2 Active Spaces for RASSCF/RASPT2 calculations 51
AS1 includes the bonding and anti-bonding σ-orbitals σ(a1 ), σ(b1 ), σ ∗ (a1 ),
σ (b1 ), as well as the HOMO and LUMO (π3 (a2 ) and π4∗ (b2 )) in RAS2. Since
∗
the bonding σ-orbitals are very low in energy with a occupation number of
2.0, they are attributed to RAS1 in AS2. Instead, the HOMO-1 (π2 (a2 ))
is moved to RAS2 for a better treatment of the excitations starting from
the occupied HOMO (and HOMO-1) orbital(s). In the third considered
active space (AS3) the LUMO+1 (π5∗ (b2 )) is taken into RAS2, additionally.
However, this does not result in any improvement in comparison with AS2.
The final active space, AS4, consists of HOMO (π3 (a2 )), HOMO-1 (π2 (a2 )),
LUMO (π4∗ (b2 )) and the four π-orbitals of the side ring, πph,2 (b1 ), πph,3 (a2 ),
∗ ∗
πph,4 (a2 ), πph,5 (b1 ). In summary, AS1-AS3 are suitable for dissociative
calculations, and AS4 for CT.
After a comparison of the excited state at GS (A1 ) geometry in all four
ASs, AS2 and AS4 were chosen to be utilized in further investigations on
the RASPT2 level of theory. AS1 and AS3 were neglected in favor of AS2,
because they have more CSFs, but do not present any improvement in
excitation energies. From now on, AS2 is also referred to as the dissociative
AS , since this AS involves excitations into the antibonding C-I orbitals,
σ ∗ (a1 ) and σ ∗ (b1 ), within RAS2 and, thus, is used to describe dissociative
processes. AS4 is the charge transfer AS because it includes the important
side-ring π/π ∗ orbitals in RAS2, as well as the HOMO and LUMO of the
boron moiety.
All calculated roots with each AS (and different level shifts) are collected
in section A.2.
of 0.3 a.u. More detail on the impact of the level shift is summarized in
Appendix A. However, SOCs were influenced by up to a factor of two (see
Appendix C). Thus, for calculations whose goal it was to obtain SOCs, the
smallest possible level shift with respect to a fitting reference weight was
used.
From now on, unless noted otherwise, all RASPT2 calculations include a
level shift of 0.3 a.u.
4 Results for non-reduced singlet
system
The preliminary step to asses the photochemistry of the present dye is to
obtain insight into its initial photo-excitation in the Franck-Condon region,
followed by possible relaxation mechanisms. This will be done for the
non-reduced singlet and the singly reduced doublet system of the BODIPY.
The first point of interest is the non-reduced singlet system. Here, the
postulated photoactivation mechanism is an excitation of the BODIPY in its
singlet ground state, followed by a relaxation in a dissociative or CT state
upon ISC into the triplet manifold, recall Figure 1.4.
In this chapter, a uniform color-code is used to label different excited
states of interest. Black represents the S0 ground state, blue excited singlet
states in general, red dissociative singlet states, green triplet states in general,
orange dissociative triplet states and cyan triplet CT states. Dissociative
state implicates an excitation into an anti-bonding σ ∗ C-I orbital, while CT
state features excitation into the phenyl ring from the main π-system or vice
versa.
Table 4.1: Comparison of TDDFT and RASPT2 results, namely excitation energies, oscillator strengths and excited
state characters obtained in the fully optimized GS structure
TDDFT RASPT2
Transition name
state E/eV f state E/eV f AS
π3 (a2 ) → π4∗ (b2 ) S1 (B1 ) 2.88 0.518 S1 (B1 ) 2.71 0.938 2 HOMO → LUMO
π2 (a2 ) → π4∗ (b2 ) S3 (B1 ) 3.45 0.311 S2 (B1 ) 3.65 0.094 2 HOMO-1 → LUMO
S5(B2) T11(A1)
5.5 T10(B2)
T6(B2)
T8(B1)
5.0 S3(A2)
T13(A1) T7(A2)
S11(B2) T5(A2)
4.5 S8(A2) T12(B2)
T8(B2) T11(B1)
4.0 T7(A2) T6(A2)
S2(B1) S2(B1)
3.5 S3(B1) S3(B1)
Energy in eV
T2(B1) T2(B1)
3.0 S1(B1) f = 0.518 S1(B1) f = 0.518
S1(B1) S1(B1)
T2(B1) f = 0.938 T2(B1) f = 0.906
2.5
1.0
0.5
The excitation energy of the bright S1 (B1 ) state is stabilized from 2.88 to
2.71 eV going from TDDFT to RASPT2, whereas an opposite trend is
observed for dissociative and CT states. TDDFT underestimates these
states by up to 1.1 eV. This shows clearly the problem of (TD)DFT to
describe this kind of systems.
Moreover, a population of the desired CT and dissociative states in the
GS geometry within the visible light spectrum can be excluded.
56 4 Results for non-reduced singlet system
Table 4.2: Results for SOCs from the bright S1 (B1 ) in AS2
Table 4.3: Results for SOCs from the bright S1 (B1 ) in AS4 with level shift of
0.3
1 Like discussed in subsection 3.2.2 the level shift influences the SOCs; thus, the minimal
possible level shift regarding a reasonable reference weight has been chosen. This is
0.0 a.u. for AS2 and 0.3 a.u. for AS4.
4.3 TDDFT geometry-states correlation diagram 57
T7 (A2 ) T8 (B2 )
Figure 4.3: Optimized TDDFT structures for dissociative states with R(C-I)
bond length. At GS (A1 ) geometry R(C-I) = 2.082 Å
4.4 Dissociation
The next step was a more thorough investigation of the dissociation starting
from the GS (A1 ) equilibrium, since the TDDFT geometry-states correlation
diagram (Figure 4.2) showed already some promising results concerning the
population of the dissociative states.
In the following section, several methods and approaches have been used to
describe the dissociation of the non-reduced BODIPY species. A MEP with
RASSCF, diabatic coordinates for dissociation with TDDFT and RASPT2
and spin-orbit kinetics were applied.
60 4 Results for non-reduced singlet system
0.00
Energy
rel. Energy in Ha
−0.01
−0.02
−0.03
−0.04
1 2 3 4 5 6 7 8 9
MEP Step
Figure 4.4: Pathway of the energy (left) and the final structure at MEP step 9
(right).
2 The state S5 (B2 ) in RASSCF/RASPT2 notation is the same like state S11 (B2 ) in
TDDFT notation, defined through their excitation pattern. Therefore, the results ob-
tained in the section before regarding S11 (B2 ) and its relaxation into a di-dissociation
also hold for S5 (B2 ). Recall Table 4.1 for all important states.
4.4 Dissociation 61
Figure 4.7: LICC between S0 (A1 ) Figure 4.8: LICC between S0 (A1 )
and S11 (B2 ) geometry and S8 (A2 ) geometry
Figure 4.9: Elongation of the C-I bond to 7 Ångström for the symmetric vi-
bration at GS (left) and the LICC between S0 (A1 ) and S11 (B2 )
geometry (right)
4.4 Dissociation 63
S0(A1) S1(B1) S8(A2) S11(B2) T7(A2) T8(B2) T6(A2) T11(B1) T12(B2) T13(A1)
13
12
11
10
9
Energy in eV
8
7
6
5
4
3
2
1
0
2.0 2.2 2.4 2.6 2.8 3.0 3.2 3.4 3.6 3.8 4.0 4.2 4.4 4.6 4.8 5.0
R(C-I) in Ångström
Figure 4.10: TDDFT potentials along the frozen symmetric C-I elongation.
Colorcode: ground state, singlet state, dissociative singlet state,
triplet state, dissociative triplet state, triplet CT state. A smaller
detail of this coordinate until an elongation of 3.0 Å can be seen
in Figure D.1
the potentials of the dissociative states are not dissociative. They rather
feature a minimum at approximately 2.32 Å after crossing the S1 (B1 ) state.
Furthermore, the degeneracy of the (now pseudo-)dissociative state at long
elongation can be seen. This is a reason why bigger elongations are not
accessible with (TD)DFT. It ends in a degenerated ground state that cannot
be described with this single determinant method. The potential energy
landscape of CT states is hardly influenced by this coordinate and follows
the S0 ground state potential.
The idea that this might be due to the poor description by TDDFT was
excluded after performing RASPT2 simulations along the frozen coordinate
led to a qualitatively similar picture; see PESs in Figure 4.11. Several
observations have to be mentioned in regard to the RASPT2 PES. First
64 4 Results for non-reduced singlet system
of all, like seen in Figure 4.1, the dissociative3 excited states are higher
in energy and the S1 (B1 ) state is lower in energy compared to TDDFT,
which, in consequence, leads to a crossing of the dissociative states with S1
at longer distances of approximately 2.6 Å. Secondly, the minimum of the
dissociative states is found at around 2.3 Å, which is in accordance with the
final MEP structure (see Figure 4.4). Thirdly, the SOCs were calculated
along the coordinate.4 As visualized in Figure 4.12 for S1 (B1 )/T5 (A2 ) and
S1 (B1 )/T6 (B2 ) the obtained SOCs show a small dependency with respect to
the C-I distance. Between the GS structure and the crossing at approximately
2.6 Å the coupling slightly rises and decreases again until values of 0.6 cm−1
and 37.0 cm−1 are obtained at the crossing. Not surprisingly, the SOCs
decrease for longer distances. But overall no significant high value can be
observed, which indicates a low probability for ISC.
A comparison of the frozen potentials with the states at the optimized,
dissociative S11 (B2 ) geometry on TDDFT and RASPT2 level of theory,
revealed a lowering of the dissociative (and CT) states by preservation of the
GS and S1 (B1 ) state energy, accompanied by an increase of the SOCs. For
the graphical display of these results, see section D.1 and also Figure 4.13.
This comparison indicates that a relaxation of the system results in an
energetic lowering of dissociative and CT states. Thus, the potentials were
calculated as relaxed scan along both C-I bond elongations, which is shown
for TDDFT in Figure 4.13 and reveals no significant improvement regarding
the conceptual question for dissociative states.
To summarize, a symmetric di-dissociation as result of an ISC (SOC)
between S1 (B1 ) and a dissociative state is impossible. This result is in
agreement with the fact that it was possible to find a minimum during the
optimization of these very dissociative states with TDDFT. The position
of the excited states of the optimized geometry S11 (B2 ) is in accordance
with the minima found in the diabatic potentials at around 2.32 Å, which
can be seen for the potentials along the symmetric elongation in Figure 4.13
and section D.1. Another interesting observation is that the minimum is
described equally, independent of the method or degree of relaxation. Merely
the excitation energy levels change and, thus, the position of the crossing
points.
3 Even though these states are not really dissociative, they will be called so for notation
purposes.
4 It is important to mention that following the results of subsection 3.2.2 the smallest
possible level shift was chosen to get only marginally influenced SOCs. This was at a
shift of 0.1 for the RASPT2 calculations.
4.4 Dissociation 65
5
4
3
2
1
0
2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 3.0
R(C-I) in Å
Figure 4.11: RASPT2 potentials along the frozen symmetric C-I elongation.
Colorcode: ground state, singlet state, dissociative singlet state,
dissociative triplet state
1.25
SOC in cm−1
S1(B1)-T5(A2)
1.00
0.75
0.50
0.25
60
SOC in cm−1
S1(B1)-T6(B2)
40
20
0
2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 3.0
R(C-I) in Å
Figure 4.12: SOCs for the diabatic potentials along the frozen symmetric C-I
elongation between the S1 (B1 ) state and the two dissociative
triplet states, T5 (A2 ) (above) and T6 (B2 ) (below); calculated
with RASPT2.
66 4 Results for non-reduced singlet system
7
6
5
4
3
2
1
0
2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 3.0 3.1 3.2 3.3 3.4 3.5 3.6 3.7 3.8 3.9 4.0
R(C-I) in Å
S0(A1) S1(B1) S8(A2) S11(B2) T7(A2) T8(B2) T6(A2) T11(B1) T12(B2) T13(A1)
9
8
7
6
Energy in eV
5
4
3
2
1
0
2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 3.0 3.1 3.2 3.3 3.4 3.5 3.6 3.7 3.8 3.9 4.0
R(C-I) in Å
Figure 4.14: TDDFT PESs along the frozen mono C-I elongation. Colorcode:
ground state, singlet state, dissociative singlet state, triplet state,
dissociative triplet state, triplet CT state
S0(A1) S1(B1) S8(A2) S11(B2) T7(A2) T8(B2) T6(A2) T11(B1) T12(B2) T13(A1)
9
8
7
6
Energy in eV
5
4
3
2
1
0
2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 3.0 3.1 3.2 3.3 3.4 3.5 3.6 3.7 3.8 3.9 4.0
R(C-I) in Å
Figure 4.15: TDDFT PESs along the relaxed mono C-I elongation. Colorcode:
ground state, singlet state, dissociative singlet state, triplet state,
dissociative triplet state, triplet CT state
68 4 Results for non-reduced singlet system
5.00
6.0 S1(B1) S1(B1)
Potential S1(B1) 4.75 Potential S1(B1)
5.5
T7(A2) 4.50 T8(B2)
5.0 Potential T7(A2) Potential T8(B2)
4.25
Energy in eV
Energy in eV
4.5 λi 4.00 λf
4.0 3.75
λf
3.5 3.50 λi
3.25
3.0 ΔE
ΔE 3.00
2.5
2.75
S0(A1) T7(A2) S0(A1) T8(B2)
Geometry Geometry
For both cases, the results show a very slow ISC of 1.69 × 104 s−1 for
S1 (B1 ) to T7 (A2 ) and 1.51 × 103 s−1 for S1 (B1 ) to T8 (B2 ), but still within
the range of literature values. [189] This means that the population of the
dissociative states occurs very slowly, which might be an explanation for the
slow cleavage after several hours of irradiation observed in experiments. [86]
The final and initial reorganization energies (λf and λi ) should ideally
be the same or at least vary little, which is not the case, especially for the
mono-dissociation. According to Marcus theory, both potentials for a given
radiationless transition should have the same curvature, which would result
5 This is the energy, that is required to get from one optimized geometry to another on
the diabatic surface of one state without crossing.
70 4 Results for non-reduced singlet system
Table 4.4: Results for ISC based on Marcus theory and Equation 4.1. f = final
and i = initial, representing the geometry from which the reorgani-
zation energy was taken
in consistent reorganization energies. Since only two points on the PESs were
used here, a fit to obtain potentials, which describe the system sufficiently,
was not possible, resulting in varying reorganization energies.
The crude approximation in the calculations is the use of the SOCs at
the GS (A1 ) geometry. The justification for this are the results of the SOCs
seen for the scan coordinates (Figure 4.12), which showed no pronounced
dependency on the C-I distance between ground state and crossing point.
Nevertheless, the SOCs were calculated with RASPT2 exemplarily at the
geometry derived from the crossing point of the two quadratic potentials of
S1 (B1 ) and T8 (B2 ) (recall Figure 4.16); the result is shown in Table 4.5. Two
conclusions can be derived from it. First, the big difference of the methods
TDDFT and RASPT2, which shifts the states energetically, consequences in
not (near) degeneracy for RASPT2 at the TDDFT crossing point. Secondly,
even though the energy gap between the bright S1 (B1 ) and the dissociative
T5 (A2 ) state gets halved compared to the GS (A1 ) geometry (Table 4.1),
the SOCs do not change significantly.
To obtain the SOC at the RASPT2 crossing point a scan along this
S1 (B1 )/T8 (B2 ) LICC would have to be performed with RASPT2; the same
holds for S1 (B1 )/T7 (A2 ). Moreover, a complete scan along these LICCs
would be needed to verify the use of Marcus theory.
4.5 Charge-Transfer 71
Table 4.5: RASPT2 results with SOCs at the crossing point of the quadratic
Marcus theory potentials S1 (B1 ) and T8 (B2 ), being S1 (B1 )/T5 (A2 )
in RASPT2 state notation (see Table 4.1)
4.5 Charge-Transfer
A further investigation of the photoactivation via charge transfer states is the
next step. Figure 4.2 already showed that CT states do not get energetically
stabilized at any optimized geometry. Furthermore, the optimized CT
geometries themselves do not expose any major change in their structure
compared to the GS (A1 ) geometry. Thus, an easily traceable coordinate
does not exist.
A reason for the high lying CT states might be the fact that the two
rings are orthogonal to each other and, thus, their orbitals do not overlap.
Therefore, a torsion around the dihedral of the main- and side-ring, illustrated
in Figure 4.17, might cause an energetic stabilization of the CT states.
Figure 4.17: Dihedral torsion around the two aromatic ring systems
For a first look, a frozen scan in the interval [90◦ , 10◦ ] and with a step size
of 5◦ around the dihedral was performed using TDDFT. The eclipsed, as
well as the staggered frozen scan, showed a small stabilization of the T6 (A2 )
72 4 Results for non-reduced singlet system
excitation energy of 0.55 eV and 0.65 eV, respectively (see Figure D.5).6
Consequently, a relaxed scan was conducted. The calculated PESs are
displayed in Figure 4.18, which reveal a significant maximum energetic
stabilization of 0.98 eV at 45◦ for the T6 (A2 ) CT state.
T13(A1)
5.5
T12(B2)
5.0
4.5 T11(B1)
4.0
Energy in eV
S1(B1)
3.5
T6(A2)
3.0
2.5
2.0
1.5
1.0 S0(A1)
0.5
0.0
15 20 25 30 35 40 45 50 55 60 65 70 75 80 85 90
Dihedral (◦)
Figure 4.18: TDDFT PESs along the relaxed scan around the dihedral be-
tween the two aromatic ring systems. Colorcode: ground state (o)
singlet state (o), triplet CT state (x)
Table 4.6: Results for SOC with RASPT2 AS4 with a level shift of 0.3 in the
fully optimized non-reduced geometry for a torsion of 55◦ around
the main/side ring dihedral.
The results exhibit the same problems as were seen with the Marcus
theory (Table 4.5). Due to the overestimation of the S1 (B1 ) state and the
underestimation of the CT states with TDDFT, 55◦ does not represent a
7 The CT state T6 (A2 ) with TDDFT represents the T7 (A2 ) state with RASPT2-AS2.
74 4 Results for non-reduced singlet system
Table 5.1: Comparison of TDDFT and RASPT2 results, namely excitation energies, oscillator strengths, and excited
state characters obtained in the fully optimized singly reduced doublet GS structure.
TDDFT RASPT2
Transition name
state E/eV f state E/eV f AS
∗
π4∗ (b2 ) → πph,5 (b1 ) D1 (B1 ) 2.04 0.000 D1 (B1 ) 2.11 0.000 4 CT
∗
π4∗ (b2 ) → πph,4 (a2 ) D2 (A2 ) 2.11 0.000 D5 (A2 ) 3.54 0.150 4 CT
π4∗ (b2 ) → σ ∗ (a1 ) D3 (A1 ) 2.12 0.000 D1 (A1 ) 2.58 0.000 2 Diss.
π4∗ (b2 ) → σ ∗ (b1 ) D4 (B1 ) 2.26 0.000 D3 (B1 ) 2.83 0.000 2 Diss.
π3 (a2 ) → π4∗ (b2 ) D5 (A2 ) 2.69 0.194 D2 (A2 ) 2.49 0.183 4 HOMO → LUMO
π2 (a2 ) → π4∗ (b2 ) D6 (A2 ) 3.38 0.003 D4 (A2 ) 3.35 0.010 4 HOMO-1 → LUMO
π3 (a2 ) → σ ∗ (a1 ) Q1 (B1 ) 3.55 - Q1 (B1 ) 4.26 - 2 Diss.
π3 (a2 ) → σ ∗ (b1 ) Q3 (A1 ) 3.64 - Q2 (A1 ) 4.43 - 2 Diss.
∗
π3 (a2 ) → πph,5 (b1 ) - - - Q2 (A1 ) 4.72 - 4 CT
5 Results for reduced doublet system
5.1 Comparison of TDDFT and RASPT2 results 77
5.0
Q2(A1)
Q2(A1)
4.5
Q1(B1)
4.0
Q3(A1)
Q1(B1) D5(A2)
3.5 D4(A2)
D6(A2) D4(A2)
3.0 D3(B1)
Energy in eV
f = 0.2630 f = 0.1942
D2(A2) D5(A2) f = 0.1831
D2(A2)
2.5 D1(A1) D4(B1)
D3(A1) D1(B1)
2.0 D2(A2)
D1(B1)
1.5
1.0
0.5
D0(B2) D0(B2) D0(B2)
0.0
RASPT2-AS2 TDDFT RASPT2-AS4
Figure 5.1: Comparison of TDDFT and RASPT2 (AS2 and AS4) states at
singly reduced GS doublet (B2 ) geometry. Colorcode: ground
state, doublet state, dissociative doublet state, CT doublet state,
dissociative quartet state, CT quartet state
1 Be aware of the TDDFT and RASPT2 state notation, seen in Table 5.1
78 5 Results for reduced doublet system
Beside the already mentioned differences, the RASPT2 state D5 (A2 ) has
to be discussed in detail. It can be seen in Table 5.1 that this state has an
oscillator strength of 0.15 a.u. Furthermore, the difference of the excitation
energy compared to the corresponding TDDFT state D2 (A2 ) is with 1.4 eV
very large. The problem, and presumably the underlying cause, is that the
weight of the important CT character is only 29 %. The main contribution
of this state is the transition π4∗ (b2 ) to π4∗ (a2 ) with 38 %, which might be
the transition contributing to the high transition dipole momentum. A
separation of these two transitions into separated electronic states could not
be achieved, e.g. by changing the level shift, input orbitals, convergence
criteria etc. (see Table A.13). To summarize, the conclusion of a population
of the RASPT2 state D5 (A2 ) at GS geometry via irradiation should not be
made.
changed to the dissociative state D3 (A1 ). For the diabatic follow up of the
bright state within this new ground state D3 (A1 ) a double excitation into
the π4∗ (b2 ) from the HOMO π3 (a2 ), as well as from the Singly Occupied
Molecular Orbital (SOMO) σ ∗ (a1 ) would be needed.
3.0
2.5
Energy (eV)
2.0
1.5
1.0
0.5
0.0
D0(B2) D1(B1) D2(A2) D3(A1) D4(B1)
Geometry
Figure 5.2: TDDFT geometry-states correlation diagram for the first five
doublet states. Colorcode: ground state, bright doublet state,
dissociative doublet state, CT doublet state
D3 (A1 ) D4 (B1 )
Figure 5.3: Optimized TDDFT structures for the dissociative states with
their R(C-I) bond length. At doublet GS (B2 ) geometry R(C-I)
= 2.095 Å
80 5 Results for reduced doublet system
5.3 Dissociation
Based on the previous results for the non-reduced system and the TDDFT
geometry-states correlation diagram (recall Figure 5.2), only the antisym-
metric mono-dissociation into the symmetric σ ∗ (a1 ) orbital was taken into
account. However, this means a loss in symmetry compared to the C2v
symmetry at the ground state. Hence, all calculations were done using
TDDFT exclusively.
Again, a scan coordinate (frozen and relaxed) describing the simple elonga-
tion of one C-I bond was chosen over the LICC between D0 (B2 ) and D3 (A1 )
and an antisymmetric vibrational mode at the GS, both displayed in Figure
Figure 5.4, for the sake of simplicity. The mono-dissociative vibrational
mode exhibits predominantly displacement vectors not associated with the
cleavage. The LICC differs only marginally from the used elongation scan
coordinate.
Figure 5.4: Left: vibrational mode at ν = 201 cm−1 . Right: LICC between
D0 (B2 ) and D3 (A1 ). Both coordinates of the reduced dye were not
investigated, in favor of an elongation scan.
The aim of the calculations was to clarify whether or not the population
of the mono-dissociative state D3 (A1 ) is leading to a C-I cleavage. The
result for the relaxed scan can be seen in Figure 5.5,2 for the frozen scan see
Figure D.6.
The state D3 (A1 ) is clearly dissociative until the crossing with the ground
state. In the region of the crossing, the results are poorly described due to
2 This coordinate shows the same problem of TDDFT regarding double excitations, as
seen in section 5.2. The diabatic follow up of the bright state would be a double
excitation into the π4∗ (b2 ) from the HOMO π3 (a2 ), as well as from the SOMO σ ∗ (a1 ),
and this is not possible with TDDFT.
5.3 Dissociation 81
4.0
3.5
3.0 D5(A2)
Energy (eV)
D4(B1)
2.5
D3(A1)
2.0 D2(A2)
D1(B1)
1.5
1.0
0.5
0.0 D0(B2)
2.0 2.2 2.4 2.6 2.8 3.0 3.2 3.4 3.6 3.8 4.0
R(C-I) in Å
Figure 5.5: TDDFT PESs along the relaxed mono C-I elongation. The arrows
indicate a possible relaxation mechanism. Colorcode: ground state,
bright doublet state, dissociative doublet state, CT doublet state
3.5
3.0
D5(A2)
2.5 D4(B1)
D3(A1)
Energy (eV)
2.0 D2(A2)
D1(B1)
1.5
1.0
0.5
0.0 D0(B2)
10 20 30 40 50 60 70 80 90
Dihedral (◦)
Figure 5.6: TDDFT PESs along the relaxed scan around the two aromatic
ring systems. The arrows indicate a possible relaxation mechanism
starting from the bright D5 (A2 ) state. Colorcode: ground state,
bright doublet state, dissociative doublet state, CT doublet state
3.5
4.0
3.0
D5(A2) 3.5
2.5 D4(B1)
D3(A1) 3.0 D5(A2)
Energy (eV)
Energy (eV)
Figure 6.1: Two possible relaxation mechanisms for the reduced system. Left:
Relaxation starting from the bright D5 towards CT and dissocia-
tive states along the relaxed scan around the ring dihedral. Right:
Dissociation of the C-I bond with crossing of the GS along the
relaxed C-I elongation. Colorcode: ground state, bright doublet
state, dissociative doublet state, CT doublet state
energetically less favorable and, secondly, that the SOC showed very small
values for the population of these states (recall Table 4.3 and 4.6).
The reduced doublet system allows a possible population of both, dissocia-
tive and charge transfer states, even at the GS geometry (see Figure 5.1). The
symmetric di-dissociation is again not favorable, but the mono-dissociation
yields a cleavage of the C-I bond with a crossing of the GS along its mono-
dissociative coordinate. Thus, a population of the dissociative state does not
result in an immediate photodegradation, coinciding with the reported 20 h
of photostability (recall chapter 1). Two possible mechanisms for photoacti-
vation were found. First of all, only small vibrations in the ring systems are
needed to populate the low lying CT states. Secondly, a scan around the
main/side ring dihedral revealed promising results for populating CT, but,
unfortunately, also for dissociative states. The excitation into the bright
doublet state and its relaxation towards a minimum on its PES results in
a crossing with the CT states, which cross the dissociative potentials by
relaxing towards their minima. The mono-dissociative state can then relax
according to the potential discussed before, which is depicted in Figure 6.1.
A major advantage of the reduced system is that all states involved have
the same spin symmetry and, therefore, do not depend on high SOCs. On
the other hand, no indications for the need of the Iodines due to a relativistic
6 Conclusion 87
First of all, calculations with different RASPT2 active spaces were con-
ducted - one to treat dissociative states and one for CT transitions - and
subsequently compared to TDDFT results. This was done for several excited
states at the Franck-Condon region in the fully optimized non-reduced singlet
geometry (see Table 4.1), as well as in the fully optimized reduced doublet
geometry (see Table 5.1). Generally, TDDFT overestimated the bright
© Springer Fachmedien Wiesbaden GmbH, part of Springer Nature 2019
K. M. Ziems, Photophysics and Photochemistry of a BODIPY-Based
Photosensitizer, BestMasters, https://doi.org/10.1007/978-3-658-26188-7_7
90 7 Summary
which prevents any approach to increase the dye’s photostability, and iii)
the poor quantitative description of boron-containing systems with TDDFT
as level of theory was confirmed. iv) A significant heavy atom effect of the
Iodines could not be observed.
[2] Kim, D.; Sakimoto, K. K.; Hong, D.; Yang, P. Angewandte Chemie
International Edition 2015, 54, 3259–3266.
[5] Yuan, Y.-J.; Yu, Z.-T.; Chen, D.-Q.; Zou, Z.-G. Chemical Society
Reviews 2017, 46, 603–631.
[6] Liao, W.-M.; Zhang, J.-H.; Hou, Y.-J.; Wang, H.-P.; Pan, M. Inorganic
Chemistry Communications 2016, 73, 80–89.
[11] Xu, Y.; Kraft, M.; Xu, R. Chemical Society Reviews 2016, 45, 3039–
3052.
[12] Oshima, T.; Lu, D.; Ishitani, O.; Maeda, K. Angewandte Chemie
International Edition 2015, 54, 2698–2702.
[13] Joya, K. S.; Joya, Y. F.; Ocakoglu, K.; van de Krol, R. Angewandte
Chemie International Edition 2013, 52, 10426–10437.
[14] Wang, X.; Maeda, K.; Chen, X.; Takanabe, K.; Domen, K.; Hou, Y.;
Fu, X.; Antonietti, M. Journal of the American Chemical Society 2009,
131, 1680–1681.
[15] Zhang, P.; Zhang, J.; Gong, J. Chemical Society Reviews 2014, 43,
4395–4422.
[16] Yang, J.; Wang, D.; Han, H.; Li, C. Accounts of chemical research
2013, 46, 1900–1909.
[17] Chang, K.; Mei, Z.; Wang, T.; Kang, Q.; Ouyang, S.; Ye, J. ACS nano
2014, 8, 7078–7087.
[18] Wu, B.; Liu, D.; Mubeen, S.; Chuong, T. T.; Moskovits, M.;
Stucky, G. D. J. Am. Chem. Soc 2016, 138, 1114–1117.
[19] Zhang, X.; Jin, Z.; Li, Y.; Li, S.; Lu, G. The Journal of Physical
Chemistry C 2009, 113, 2630–2635.
[20] Gong, L.; Wang, J.; Li, H.; Wang, L.; Zhao, J.; Zhu, Z. Catalysis
Communications 2011, 12, 1099–1103.
[21] Zhang, W.; Hong, J.; Zheng, J.; Huang, Z.; Zhou, J.; Xu, R. Journal
of the American Chemical Society 2011, 133, 20680–20683.
[22] Lazarides, T.; McCormick, T.; Du, P.; Luo, G.; Lindley, B.; Eisen-
berg, R. Journal of the American Chemical Society 2009, 131, 9192–
9194.
[23] McCormick, T. M.; Calitree, B. D.; Orchard, A.; Kraut, N. D.;
Bright, F. V.; Detty, M. R.; Eisenberg, R. Journal of the American
Chemical Society 2010, 132, 15480–15483.
[27] Yuan, Y.-J.; Yu, Z.-T.; Chen, D.-Q.; Zou, Z.-G. Chemical Society
Reviews 2017, 46, 603–631.
[28] Chan, S.-F.; Chou, M.; Creutz, C.; Matsubara, T.; Sutin, N. Journal
of the American Chemical Society 1981, 103, 369–379.
Bibliography 99
[44] Du, P.; Knowles, K.; Eisenberg, R. Journal of the American Chemical
Society 2008, 130, 12576–12577.
[45] Metz, S.; Bernhard, S. Chemical Communications 2010, 46, 7551–
7553.
[46] DiSalle, B. F.; Bernhard, S. Journal of the American Chemical Society
2011, 133, 11819–11821.
[47] Whang, D. R.; Sakai, K.; Park, S. Y. Angewandte Chemie International
Edition 2013, 52, 11612–11615.
[48] Mak, C. S.; Wong, H. L.; Leung, Q. Y.; Tam, W. Y.; Chan, W. K.;
Djurišić, A. B. Journal of Organometallic Chemistry 2009, 694, 2770–
2776.
[49] Esswein, A. J.; Nocera, D. G. Chemical reviews 2007, 107, 4022–4047.
[50] Eckenhoff, W. T.; Eisenberg, R. Dalton Transactions 2012, 41, 13004–
13021.
[51] Cossi, M.; Rega, N.; Scalmani, G.; Barone, V. Journal of computational
chemistry 2003, 24, 669–681.
[52] Lakadamyali, F.; Kato, M.; Muresan, N. M.; Reisner, E. Angewandte
Chemie International Edition 2012, 51, 9381–9384.
[53] Sakai, T.; Mersch, D.; Reisner, E. Angewandte Chemie International
Edition 2013, 52, 12313–12316.
[54] Cheng, M.; Yang, X.; Li, J.; Zhang, F.; Sun, L. ChemSusChem 2013,
6, 70–77.
[55] Zhang, P.; Wang, M.; Li, C.; Li, X.; Dong, J.; Sun, L. Chemical
Communications 2010, 46, 8806–8808.
[56] Han, Z.; McNamara, W. R.; Eum, M.-S.; Holland, P. L.; Eisenberg, R.
Angewandte Chemie International Edition 2012, 51, 1667–1670.
[57] Lazarides, T.; McCormick, T.; Du, P.; Luo, G.; Lindley, B.; Eisen-
berg, R. Journal of the American Chemical Society 2009, 131, 9192–
9194.
[58] Zhang, W.; Hong, J.; Zheng, J.; Huang, Z.; Zhou, J.; Xu, R. Journal
of the American Chemical Society 2011, 133, 20680–20683.
Bibliography 101
[59] Hartley, C. L.; DiRisio, R. J.; Screen, M. E.; Mayer, K. J.; McNa-
mara, W. R. Inorganic chemistry 2016, 55, 8865–8870.
[60] Sabatini, R. P.; McCormick, T. M.; Lazarides, T.; Wilson, K. C.;
Eisenberg, R.; McCamant, D. W. The Journal of Physical Chemistry
Letters 2011, 2, 223–227.
[61] Bartelmess, J.; Francis, A. J.; El Roz, K. A.; Castellano, F. N.;
Weare, W. W.; Sommer, R. D. Inorganic chemistry 2014, 53, 4527–
4534.
[62] Manton, J. C.; Long, C.; Vos, J. G.; Pryce, M. T. Physical Chemistry
Chemical Physics 2014, 16, 5229–5236.
[63] Luo, G.-G.; Fang, K.; Wu, J.-H.; Dai, J.-C.; Zhao, Q.-H. Physical
Chemistry Chemical Physics 2014, 16, 23884–23894.
[64] Luo, G.-G.; Fang, K.; Wu, J.-H.; Mo, J. Chemical Communications
2015, 51, 12361–12364.
[65] Luo, G.-G.; Lu, H.; Zhang, X.-L.; Dai, J.-C.; Wu, J.-H.; Wu, J.-J.
Physical Chemistry Chemical Physics 2015, 17, 9716–9729.
[66] Loudet, A.; Burgess, K. Chemical reviews 2007, 107, 4891–4932.
[67] Ulrich, G.; Ziessel, R.; Harriman, A. Angewandte Chemie International
Edition 2008, 47, 1184–1201.
[68] Momeni, M. R.; Brown, A. Journal of chemical theory and computation
2015, 11, 2619–2632.
[69] Ni, Y.; Zeng, W.; Huang, K.-W.; Wu, J. Chemical Communications
2013, 49, 1217–1219.
[70] Gresser, R.; Hartmann, H.; Wrackmeyer, M.; Leo, K.; Riede, M.
Tetrahedron 2011, 67, 7148–7155.
[71] Hayashi, Y.; Obata, N.; Tamaru, M.; Yamaguchi, S.; Matsuo, Y.;
Saeki, A.; Seki, S.; Kureishi, Y.; Saito, S.; Yamaguchi, S. Organic
letters 2012, 14, 866–869.
[72] Jiang, X.-D.; Fu, Y.; Zhang, T.; Zhao, W. Tetrahedron Letters 2012,
53, 5703–5706.
102 Bibliography
[73] Yu, C.; Xu, Y.; Jiao, L.; Zhou, J.; Wang, Z.; Hao, E. Chemistry-A
European Journal 2012, 18, 6437–6442.
[74] Berhe, S. A.; Rodriguez, M. T.; Park, E.; Nesterov, V. N.; Pan, H.;
Youngblood, W. J. Inorganic chemistry 2014, 53, 2346–2348.
[75] Sarkar, S. K.; Mukherjee, S.; Thilagar, P. Inorganic chemistry 2014,
53, 2343–2345.
[76] Kamkaew, A.; Lim, S. H.; Lee, H. B.; Kiew, L. V.; Chung, L. Y.;
Burgess, K. Chemical Society Reviews 2013, 42, 77–88.
[77] Awuah, S. G.; You, Y. Rsc Advances 2012, 2, 11169–11183.
[78] Nepomnyashchii, A. B.; Bard, A. J. Accounts of chemical research
2012, 45, 1844–1853.
[79] López Arbeloa, F.; Banuelos, J.; Martínez, V.; Arbeloa, T.; López Arbe-
loa, I. International Reviews in Physical Chemistry 2005, 24, 339–374.
[80] Duran-Sampedro, G.; Agarrabeitia, A. R.; Garcia-Moreno, I.;
Costela, A.; Bañuelos, J.; Arbeloa, T.; López Arbeloa, I.; Chiara, J. L.;
Ortiz, M. J. European Journal of Organic Chemistry 2012, 2012,
6335–6350.
[81] Boens, N.; Leen, V.; Dehaen, W. Chemical Society Reviews 2012, 41,
1130–1172.
[82] Gonçalves, M. S. T. Chemical reviews 2008, 109, 190–212.
[83] Yuan, L.; Lin, W.; Zheng, K.; He, L.; Huang, W. Chemical Society
Reviews 2013, 42, 622–661.
[84] Vendrell, M.; Zhai, D.; Er, J. C.; Chang, Y.-T. Chemical reviews 2012,
112, 4391–4420.
[85] Batat, P.; Cantuel, M.; Jonusauskas, G.; Scarpantonio, L.; Palma, A.;
O’Shea, D. F.; McClenaghan, N. D. The Journal of Physical Chemistry
A 2011, 115, 14034–14039.
[86] Dura, L.; Wächtler, M.; Kupfer, S.; Kübel, J.; Ahrens, J.; Höfler, S.;
Bröring, M.; Dietzek, B.; Beweries, T. Inorganics 2017, 5, 21.
[87] Dura, L.; Ahrens, J.; Pohl, M.-M.; Höfler, S.; Bröring, M.; Beweries, T.
Chemistry-A European Journal 2015, 21, 13549–13552.
Bibliography 103
[92] Born, M.; Oppenheimer, R. Annalen der Physik 1927, 389, 457–484.
[93] Sutcliffe, B. Advances in Chemical Physics 2000, 114, 1–122.
[94] Bochevarov, A. D.; Valeev, E. F.; David SheRrill, C. Molecular Physics
2004, 102, 111–123.
[95] Hartree, D. R. The wave mechanics of an atom with a non-Coulomb
central field. Part I. Theory and methods. Mathematical Proceedings
of the Cambridge Philosophical Society. 1928; pp 89–110.
[96] Pople, J.; Nesbet, R. The Journal of Chemical Physics 1954, 22,
571–572.
[97] Kobus, J. Advances in Quantum Chemistry 1997, 28, 1–14.
[100] Weigend, F.; Furche, F.; Ahlrichs, R. The Journal of chemical physics
2003, 119, 12753–12762.
[101] Roothaan, C. Hall GG (1951) Proc R Soc (London) A 1951, 205, 541.
[102] Hoffmann, R. The Journal of Chemical Physics 1963, 39, 1397–1412.
[103] Roos, B. O. Multiconfigurational quantum chemistry. 2005.
[114] Andersson, K.; Malmqvist, P. A.; Roos, B. O.; Sadlej, A. J.; Wolin-
ski, K. Journal of Physical Chemistry 1990, 94, 5483–5488.
[115] Schrödinger, E. Annalen der physik 1926, 385, 437–490.
[118] Roos, B. O.; Andersson, K. Chemical physics letters 1995, 245, 215–
223.
[119] Hohenberg, P.; Kohn, W. Physical review 1964, 136, B864.
[120] Löwdin, P.-O. International Journal of Quantum Chemistry 1985, 28,
19–37.
[121] Kohn, W.; Sham, L. J. Physical review 1965, 140, A1133.
[122] Ullrich, C. A.; Yang, Z.-h. Brazilian Journal of Physics 2014, 44,
154–188.
Bibliography 105
[137] Runge, E.; Gross, E. K. Physical Review Letters 1984, 52, 997.
[138] van Leeuwen, R. Time-Dependent Density Functional Theory 2006,
17–31.
[139] Marques, M. A.; Maitra, N. T.; Nogueira, F. M.; Gross, E. K.; Rubio, A.
Fundamentals of time-dependent density functional theory; Springer
Science & Business Media, 2012; Vol. 837.
106 Bibliography
[140] Giuliani, G.; Vignale, G. Quantum theory of the electron liquid; Cam-
bridge university press, 2005.
[141] Ullrich, C. A. Time-dependent density-functional theory: concepts and
applications; OUP Oxford, 2011.
[142] Tsuneda, T. Density functional theory in quantum chemistry; Springer
Science & Business Media, 2014.
[143] Meijer, E. J.; Sprik, M. The Journal of chemical physics 1996, 105,
8684–8689.
[144] Cybulski, S. M.; Seversen, C. E. The Journal of chemical physics 2005,
122, 014117.
[145] Dreuw, A.; Weisman, J. L.; Head-Gordon, M. The Journal of chemical
physics 2003, 119, 2943–2946.
[146] Gritsenko, O.; Baerends, E. J. The Journal of chemical physics 2004,
121, 655–660.
[147] Becke, A. D. The Journal of chemical physics 2002, 117, 6935–6938.
[148] Spiegel, J. D.; Kleinschmidt, M.; Larbig, A.; Tatchen, J.; Marian, C. M.
Journal of chemical theory and computation 2015, 11, 4316–4327.
[149] Chibani, S.; Laurent, A. D.; Le Guennic, B.; Jacquemin, D. Journal
of chemical theory and computation 2014, 10, 4574–4582.
[150] Ji, S.; Ge, J.; Escudero, D.; Wang, Z.; Zhao, J.; Jacquemin, D. The
Journal of organic chemistry 2015, 80, 5958–5963.
[151] Neese, F.; Petrenko, T.; Ganyushin, D.; Olbrich, G. Coord. Chem. Rev.
2007, 251, 288.
[171] Fu, L.; Jiang, F.-L.; Fortin, D.; Harvey, P. D.; Liu, Y. Chemical
Communications 2011, 47, 5503–5505.
[172] Tomasi, J.; Mennucci, B.; Cammi, R. Chemical reviews 2005, 105,
2999–3094.
[173] Mennucci, B.; Tomasi, J.; Cammi, R.; Cheeseman, J.; Frisch, M.;
Devlin, F.; Gabriel, S.; Stephens, P. The Journal of Physical Chemistry
A 2002, 106, 6102–6113.
[174] Kudin, K. N.; Scuseria, G. E.; Cancès, E. The Journal of chemical
physics 2002, 116, 8255–8261.
[175] Roos, B. O.; Veryazov, V.; Widmark, P.-O. Theoretical Chemistry
Accounts 2004, 111, 345–351.
[176] Roos, B. O.; Lindh, R.; Malmqvist, P.-Å.; Veryazov, V.; Widmark, P.-
O. The Journal of Physical Chemistry A 2004, 108, 2851–2858.
[177] Roos, B. O.; Lindh, R.; Malmqvist, P.-Å.; Veryazov, V.; Widmark, P.-
O. The Journal of Physical Chemistry A 2005, 109, 6575–6579.
[178] Roos, B. O.; Lindh, R.; Malmqvist, P.-Å.; Veryazov, V.; Widmark, P.-
O. Chemical physics letters 2005, 409, 295–299.
[179] Roos, B. O.; Lindh, R.; Malmqvist, P.-Å.; Veryazov, V.; Widmark, P.-
O.; Borin, A. C. The Journal of Physical Chemistry A 2008, 112,
11431–11435.
[180] Roos, B.; Malmqvist, P.-Å.; Gagliardi, L. Fundamental World of
Quantum Chemistry 2003,
[181] Aquilante, F.; Lindh, R.; Bondo Pedersen, T. The Journal of chemical
physics 2007, 127, 114107.
[182] Malmqvist, P. A.; Rendell, A.; Roos, B. O. Journal of Physical Chem-
istry 1990, 94, 5477–5482.
[183] Roos, B. O. International Journal of Quantum Chemistry 1980, 18,
175–189.
[184] Fukui, K. Accounts of chemical research 1981, 14, 363–368.
[185] Quapp, W.; Heidrich, D. Theoretical Chemistry Accounts: Theory,
Computation, and Modeling (Theoretica Chimica Acta) 1984, 66, 245–
260.
Bibliography 109
Table A.1: Different RAS distributions and their CSFs for different ASs
AS RAS 1
1 p(a2 ), p(b2 ), π1 (a2 ), π1 (b2 ), πph,2 (b1 ), πph,3 (a2 ), π2 (a2 ), π2 (b2 ), π3 (b2 )
2 σ(a1 ), σ(b1 ), p(a2 ), p(b2 ), π1 (a2 ), π1 (b2 ), πph,2 (b1 ), πph,3 (a2 ), π2 (b2 ), π3 (b2 )
3 σ(a1 ), σ(b1 ), p(a2 ), p(b2 ), π1 (a2 ), π1 (b2 ), πph,2 (b1 ), πph,3 (a2 ), π2 (b2 ), π3 (b2
4 σ(a1 ), σ(b1 ), p(a2 ), p(b2 ), π1 (a2 ), π1 (b2 ), π2 (b2 ), π3 (b2
AS RAS 2
1 σ(a1 ), σ(b1 ), σ ∗ (a ∗ ∗
1 ), σ (b1 ), π3 (a2 ), π4 (b2 )
2 σ ∗ (a1 ), σ ∗ (b1 ), π2 (a2 ), π3 (a2 ), π4∗ (b2 )
3 σ ∗ (a1 ), σ ∗ (b1 ), π2 (a2 ), π3 (a2 ), π4∗ (b2 ), π5∗ (b2 )
4 π2 (a2 ), π3 (a2 ), π4∗ (b2 ), πph,2 (b1 ), πph,3 (a2 ), πph,4
∗ ∗
(a2 ), πph,5 (b1 )
AS RAS 3
1 π4∗ (a2 ), πph,4
∗ ∗
(a2 ), πph,5 (b1 ), π5∗ (a2 ), π5∗ (b2 ), π6∗ (b2 )
2 π4∗ (a2 ), πph,4
∗ ∗
(a2 ), πph,5 (b1 ), π5∗ (a2 ), π5∗ (b2 ), π6∗ (b2 )
3 ∗ ∗
π4 (a2 ), πph,4 (a2 ), πph,5∗ (b1 ), π5∗ (a2 ), π6∗ (b2 )
4 π4∗ (a2 ), σ ∗ (a1 ), σ ∗ (b1 ), π5∗ (a2 ), π5∗ (b2 ), π6∗ (b2 )
AS CSFs
1 222869
2 71140
3 122948
4 534332
Table A.2: TDDFT results for the fully optimized non-reduced GS structure
weight E λ
Transition f name
% eV nm
S0 (A1 ) - - - - - HF
S1 (B1 ) π3 (a2 ) → π4∗ (b2 ) 89 2.88 430 0.518 H-L
π2 (a2 ) → π4∗ (b2 ) 11 H-1-L
S2 (A1 ) π2 (b2 ) → π4∗ (b2 ) 98 3.44 360 0.045
S3 (B1 ) π2 (a2 ) → π4∗ (b2 ) 88 3.45 359 0.311 H-1-L
π3 (a2 ) → π4∗ (b2 ) 12 H-L
S8 (A2 ) π3 (a2 ) → σ ∗ (a1 ) 93 4.41 281 0.000 Diss.
S11 (B2 ) π3 (a2 ) → σ ∗ (b1 ) 87 4.55 272 0.000 Diss.
Table A.3: RASPT2 results for the fully optimized non-reduced GS structure
with AS1 and no level shift
weight E λ
Transition f name
% eV nm
S0 (A1 ) HF 81 - - - HF
S1 (B1 ) π3 (a2 ) → π4∗ (b2 ) 80 2.75 451 1.059 H-L
S2 (B1 ) π2 (a2 ) → π4∗ (b2 ) 70 4.50 276 0.002 H-1-L
DE 10
S3 (A2 ) πph,2 (b1 ) → π4∗ (b2 ) 84 4.77 260 0.000 CT
S4 (A1 ) DE: π3 (a2 ) → π4∗ (b2 ) 48 4.93 251 0.001 H-L
π2 (b2 ) → π4∗ (b2 ) 25
S5 (B2 ) π3 (a2 ) → σ ∗ (b1 ) 78 5.39 230 0.001 Diss.
S6 (B2 ) ∗
π3 (a2 ) → πph,5 (b1 ) 81 6.29 197 0.000 CT
S7 (A2 ) π3 (a2 ) → π4∗ (b2 ) 8.04 154 0.000 H-L
∗
πph,2 (b1 ) → πph,4 (a2 ) 62 intra Ph
DE 21
Table A.4: RASPT2 results for the fully optimized non-reduced GS structure
with AS2 and no level shift
weight E λ
Transition f name
% eV nm
S0 (A1 ) HF 83 - - - HF
S1 (B1 ) π3 (a2 ) → π4∗ (b2 ) 78 2.66 467 0.954 H-L
S2 (B1 ) π2 (a2 ) → π4∗ (b2 ) 75 3.62 342 0.058 H-1-L
S3 (A2 ) π3 (a2 ) → σ ∗ (a 1) 74 5.01 247 0.000 Diss.
S4 (A1 ) DE: π3 (a2 ) → π4∗ (b2 ) 59 5.05 245 0.003 H-L
π3 (b2 ) → π4∗ (b2 ) 13
S5 (B2 ) π3 (a2 ) → σ ∗ (b1 ) 77 5.37 231 0.001 Diss.
S6 (A2 ) πph,2 (b1 ) → π4∗ (b2 ) 86 5.80 214 0.000 CT
S7 (A2 ) π2 (a2 ) → σ ∗ (a 1) 65 5.84 212 0.000 Diss.
S8 (B2 ) ∗
π3 (a2 ) → πph,5 (b1 ) 84 6.29 197 0.000 CT
S9 (A2 ) DE 51 7.14 174 0.000 Diss.
DE 11
Table A.5: RASPT2 results for the fully optimized non-reduced GS structure
with AS2 and a level shift of 0.3 a.u.
weight E λ
Transition f name
% eV nm
S0 (A1 ) HF 83 - - - HF
S1 (B1 ) π3 (a2 ) → π4∗ (b2 ) 76 2.71 457 0.938 H-L
S2 (B1 ) π2 (a2 ) → π4∗ (b2 ) 75 3.65 340 0.094 H-1-L
S3 (A2 ) π3 (a2 ) → σ ∗ (a 1) 74 5.06 245 0.000 Diss.
S4 (A1 ) DE: π3 (a2 ) → π4∗ (b2 ) 59 5.06 245 0.003 H-L
π3 (b2 ) → π4∗ (b2 ) 13
S5 (B2 ) π3 (a2 ) → σ ∗ (b1 ) 77 5.42 229 0.001 Diss.
S6 (A2 ) πph,2 (b1 ) → π4∗ (b2 ) 86 5.91 210 0.000 CT
S7 (A2 ) π2 (a2 ) → σ ∗ (a 1) 64 5.93 209 0.000 Diss.
S8 (B2 ) ∗
π3 (a2 ) → πph,5 (b1 ) 84 6.33 196 0.000 CT
S9 (A2 ) DE 51 7.22 172 0.000 Diss.
DE 11
Table A.6: RASPT2 results for the fully optimized non-reduced GS structure
with AS3 and no level shift
weight E λ
Transition f name
% eV nm
S0 (A1 ) HF 82 HF
S1 (B1 ) π3 (a2 ) → π4∗ (b2 ) 77 2.63 472 0.921 H-L
S2 (B1 ) π2 (a2 ) → π4∗ (b2 ) 74 3.62 343 0.060 H-1-L
S3 (A2 ) π3 (a2 ) → σ ∗ (a 1) 73 5.00 248 0.000 Diss.
S4 (A1 ) DE: π3 (a2 ) → π4∗ (b2 ) 57 5.03 246 0.000 H-L
π3 (b2 ) → π4∗ (b2 ) 13
π3 (a2 ) → π4∗ (a2 ) 11
S5 (B2 ) π3 (a2 ) → σ ∗ (b1 ) 77 5.34 232 0.001 Diss.
S6 (A2 ) πph,2 (b1 ) → π4∗ (b2 ) 85 5.76 215 0.000 CT
S7 (A2 ) π2 (a2 ) → σ ∗ (a 1) 64 5.84 212 0.000 Diss.
S8 (B2 ) ∗
π3 (a2 ) → πph,5 (b1 ) 83 6.02 206 0.000 CT
S9 (A2 ) DE 51 Diss.
DE 11
Table A.7: RASPT2 results for the fully optimized non-reduced GS structure
with AS4 and no level shift
weight E λ
Transition f name
% eV nm
S0 (A1 ) HF 79 - - - HF
S1 (B1 ) π3 (a2 ) → π4∗ (b2 ) 75 2.64 471 0.915 H-L
S2 (B1 ) π2 (a2 ) → π4∗ (b2 ) 70 3.60 345 0.059 H-1-L
S3 (A1 ) DE: π3 (a2 ) → π4∗ (b2 ) 17 4.49 276 0.056 H-L
π2 (b2 ) → π4∗ (b2 ) 53
Table A.8: RASPT2 results for the fully optimized non-reduced GS structure
with AS4 and a level shift of 0.3 a.u.
weight E λ
Transition f name
% eV nm
S0 (A1 ) HF 79 - - - HF
S1 (B1 ) π3 (a2 ) → π4∗ (b2 ) 73 2.69 460 0.906 H-L
S2 (B1 ) π2 (a2 ) → π4∗ (b2 ) 68 3.62 342 0.099 H-1-L
S3 (A1 ) DE: π3 (a2 ) → π4∗ (b2 ) 17 4.52 275 0.056 H-L
π2 (b2 ) → π4∗ (b2 ) 53
a) b) c)
Figure A.1: Spin Densities for optimized singly reduced ground state geometry
with DFT for different multiplicities (doublet and quartet) and
optimization step size. a) doublet optimization with 0.3 Bohr step
size, b) quartet optimization with 0.3 Bohr step size, c) quartet
optimization with 0.1 Bohr step size
Table A.9: TDDFT results for the fully optimized singly reduced GS structure
Table A.10: RASPT2 results for the fully optimized singly reduced GS struc-
ture with AS2 and no level shift
weight E λ
Transition f name
% eV nm
D0 (B2 ) - - - - - -
D1 (A1 ) π4∗ (b2 ) → σ ∗ (a1 ) 70 2.53 490 0.001 Diss.
D2 (A2 ) π3 (a2 ) → π4∗ (b2 ) 49 2.59 478 0.259 H-L
π2 (a2 ) → π4∗ (b2 ) 29 H-1-L
D3 (B1 ) π4∗ (b2 ) → σ ∗ (b1 ) 73 2.76 449 0.000 Diss.
D4 (A2 ) π2 (a2 ) → π4∗ (b2 ) 48 3.40 365 0.001 H-1-L
π3 (a2 ) → π4∗ (b2 ) 31 H-L
D5 (B1 ) π4∗ (b2 ) → πph,5
∗ (b1 ) 86 3.59 346 0.000 CT
D6 (A1 ) π3 (a2 ) → σ ∗ (b1 ) 53 4.56 272 0.001 Diss.
DE 21
Table A.11: RASPT2 results for the fully optimized singly reduced GS struc-
ture with AS2 with a level shift of 0.3 a.u.
weight E λ
Transition f name
% eV nm
D0 (B2 ) - 83 - - - -
D1 (A1 ) π4∗ (b2 ) → σ ∗ (a1 ) 70 2.58 480 0.000 Diss.
D2 (A2 ) π3 (a2 ) → π4∗ (b2 ) 47 2.63 471 0.263 H-L
π2 (a2 ) → π4∗ (b2 ) 30 H-1-L
D3 (B1 ) π4∗ (b2 ) → σ ∗ (b1 ) 73 2.83 438 0.000 Diss.
D4 (A2 ) π2 (a2 ) → π4∗ (b2 ) 46 3.44 360 0.001 H-1-L
π3 (a2 ) → π4∗ (b2 ) 32 H-L
D5 (B1 ) π4∗ (b2 ) → πph,5
∗ (b1 ) 86 3.64 341 0.000 CT
D6 (A1 ) π3 (a2 ) → σ ∗ (b1 ) 54 4.61 269 0.001 Diss.
DE 21
Table A.12: RASPT2 results for the fully optimized singly reduced GS struc-
ture with AS4 and no level shift
weight E λ
Transition f name
% eV nm
D0 (B2 ) - - - - - -
D1 (A2 ) π3 (a2 ) → π4∗ (b2 ) 40 2.48 500 0.125 H-L
π2 (a2 ) → π4∗ (b2 ) 26 H-1-L
π4∗ (b2 ) → π4∗ (A2 ) 12
D2 (B1 ) π4∗ (b2 ) → πph,5
∗ (b1 ) 83 2.58 480 0.000 CT
D3 (A2 ) π2 (a2 ) → π4∗ (b2 ) 50 3.33 372 0.002 H-1-L
π3 (a2 ) → π4∗ (b2 ) 23 H-L
D4 (B1 ) π4∗ (b2 ) → σ ∗ (b1 ) 67 3.59 346 0.000 Diss.
D5 (B2 ) DE 72 4.58 271 0.000 CT
DE 11
Q1 (B2 ) ∗
πph,3 (a2 ) → πph,4 (a2 ) 56 4.29 289 - intra Ph
∗
πph,2 (b1 ) → πph,5 (b1 ) 27
Q2 (A1 ) ∗
π3 (a2 ) → πph,5 (b1 ) 80 4.64 267 - CT
Q3 (B1 ) π3 (a2 ) → σ ∗ (a
1) 37 4.85 256 - Diss.
π2 (a2 ) → σ ∗ (a1 ) 22
π2 (b2 ) → σ ∗ (b1 ) 13
Q4 (A1 ) ∗
πph,3 (a2 ) → πph,5 (b1 ) 40 4.98 249 - intra Ph
∗
πph,2 (b1 ) → πph,4 (a2 ) 38
Q5 (B2 ) ∗
πph,2 (b1 ) → πph,5 (b1 ) 53 5.11 243 - intra Ph
∗
πph,3 (a2 ) → πph,4 (a2 ) 24
124 A TDDFT and RASPT2 states
Table A.13: RASPT2 results for the fully optimized singly reduced GS struc-
ture with AS4 with a level shift of 0.3 a.u.
weight E λ
Transition f name
% eV nm
D0 (B2 ) - 81 - - - -
D1 (B1 ) π4∗ (b2 ) → πph,5
∗ (b1 ) 83 2.11 587 0.000 CT
D2 (A2 ) π3 (a2 ) → π4∗ (b2 ) 72 2.49 497 0.183 H-L
D3 (B1 ) π4∗ (b2 ) → σ ∗ (b1 ) 67 3.03 409 0.000 Diss.
D4 (A2 ) π2 (a2 ) → π4∗ (b2 ) 66 3.35 370 0.010 H-1-L
D5 (A2 ) π4∗ (b2 ) → π4∗ (a2 ) 38 3.54 350 0.150
π4∗ (b2 ) → πph,4
∗ (a2 ) 29 CT
D6 (A2 ) π4∗ (b2 ) → πph,4
∗ (a2 ) 50 4.36 284 0.150 CT
π4∗ (b2 ) → π4∗ (a2 ) 23
Q1 (B2 ) ∗
πph,3 (a2 ) → πph,4 (a2 ) 60 4.38 283 - intra Ph
∗
πph,2 (b1 ) → πph,5 (b1 ) 23
Q2 (A1 ) ∗
π3 (a2 ) → πph,5 (b1 ) 82 4.72 262 - CT
Q3 (B1 ) π3 (a2 ) → σ ∗ (a
1) 34 4.89 254 - Diss.
π2 (a2 ) → σ ∗ (a1 ) 23
π2 (b2 ) → σ ∗ (b1 ) 13
Q4 (A1 ) ∗
πph,3 (a2 ) → πph,5 (b1 ) 41 5.06 245 - intra Ph
∗
πph,2 (b1 ) → πph,4 (a2 ) 39
Q5 (B2 ) ∗
πph,2 (b1 ) → πph,5 (b1 ) 57 5.23 237 - intra Ph
∗
πph,3 (a2 ) → πph,4 (a2 ) 20
B Reference weight
Table B.1: Reference weight for AS2 of the non-reduced singlet system. Big
deviations are shown in italic and red
Table B.2: Reference weight AS4 of the non-reduced singlet system. Big devia-
tions are shown in italic and red
Table B.3: Reference weight for AS2 of the reduced doublet system
Table B.4: Reference weight for AS4 of the reduced doublet system. Big devia-
tions are shown in italic and red
Table C.1: SOCs with RASPT2 AS2 within the optimized non-reduced GS C2v
model geometry with no level shift
Table C.2: SOCs with RASPT2 AS2 within the optimized non-reduced GS C2v
model geometry with a level shift of 0.3 a.u.
Table C.3: SOCs with RASPT2 AS4 within the optimized non-reduced GS C2v
model geometry with no level shift
Table C.4: SOCs with RASPT2 AS4 in the optimized non-reduced gGS C2v
model geometry with a level shift of 0.3 a.u.
S0(A1) S1(B1) S8(A2) S11(B2) T7(A2) T8(B2) T6(A2) T11(B1) T12(B2) T13(A1)
8
7
6
Energy in eV
5
4
3
2
1
0
2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 3.0
R(C-I) in Å
Figure D.1: TDDFT potentials along the frozen symmetric C-I elongation for
the non-reduced system. Colorcode: ground state, singlet state,
dissociative singlet state, triplet state, dissociative triplet state,
triplet CT state
S0(A1) S1(B1) S8(A2) S11(B2) T7(A2) T8(B2) T6(A2) T11(B1) T12(B2) T13(A1)
8
at optimized S11(B2) geometry
7
6
Energy in eV
5
4
3
2
1
0
2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 3.0
R(C-I) in Å
Figure D.2: TDDFT potentials along the frozen symmetric C-I elongation
for the non-reduced system including the states at the optimized,
dissociative S11 (B2 ) geometry. Colorcode: ground state, singlet
state, dissociative singlet state, triplet state, dissociative triplet
state, triplet CT state
D Further diabatic potentials 135
5
4
3
2
1
0
2.0 2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 3.0
R(C-I) in Å
Figure D.3: RASPT2 potentials along the frozen symmetric C-I elongation
for the non-reduced system including the states at the optimized,
dissociative S11 (B2 ) geometry. Colorcode: ground state, singlet
state, dissociative singlet state, dissociative triplet state
2.00
1.75
SOC in cm−1
S1(B1)-T5(A2)
1.50
1.25
1.00
0.75
0.50
0.25
60
SOC in cm−1
S1(B1)-T6(B2)
40
20
0
2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 3.0
R(C-I) in Å
Figure D.4: SOCs for the diabatic potentials along the frozen symmetric
C-I elongation and at optimized S11 (B2 ) geometry between
the S1 (B1 ) state and the two dissociative triplet states, T5 (A2 )
(above) and T6 (B2 ) (below). Calculated with RASPT2 for the
non-reduced system.
136 D Further diabatic potentials
30
28 T13(A1) T12(B2)
26 T11(B1)
24 T6(A2)
22
20 S0(A1)
Energy in eV
18
16
14
12
10
8
6
4
2
0
10 20 30 40 50 60 70 80 90
Dihedral (◦)
20
18 T (A ) T12(B2)
13 1
16
T11(B1)
14 T6(A2)
Energy in eV
12 S0(A1)
10
8
6
4
2
0
10 20 30 40 50 60 70 80 90
Dihedral (◦)
Figure D.5: TDDFT PESs along the frozen scan around the dihedral be-
tween the two aromatic ring systems for the non-reduced system.
Eclipsed above and staggered below. Colorcode: ground state (o),
triplet CT state (x)
D Further diabatic potentials 137
2.5
2.0
1.5
1.0
0.5
0.0
2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 3.0 3.1 3.2 3.3 3.4 3.5 3.6 3.7 3.8 3.9 4.0
R(C-I) in Å
Figure D.6: TDDFT PESs along the frozen mono C-I elongation for the re-
duced system. Colorcode: ground state, bright doublet state,
dissociative doublet state, CT doublet state
D0(B2) D1(B1) D2(A2) D3(A1) D4(B1) D5(A2)
22
20
18
16
Energy in eV
14
12
10
8
6
4
2
0
10 20 30 40 50 60 70 80 90
Dihedral (◦)
Figure D.7: TDDFT PESs along the frozen scan around the two aromatic ring
systems for the reduced system. Colorcode: ground state, bright
doublet state, dissociative doublet state, CT doublet state
Danksagung
Zunächst möchte ich mich bei Frau Prof. Dr. Stefanie Gräfe für die heraus-
fordernde und interessante Aufgabenstellung, sowie die unterstützenden
Besprechungen bedanken.
Weiterhin gilt mein großer Dank Herrn Dr. Stephan Kupfer ohne dessen
sehr gute Betreuung diese Arbeit nicht Zustande gekommen wäre. Ein
weiterer Dank geht an die gesamte Arbeitsgruppe für die Unterstützung,
die hilfreichen Gespräch und die angenehme Arbeitsatmosphäre. Besonders
möchte ich Johannes Steinmetzer herausstellen, dessen Wissen im Bereich der
Quantenchemie im Allgemeinen und v.a. im Bereich von Python-gestützter
Auswertung und Verarbeitung von Daten meine Arbeit um ein vielfaches
erleichtert und auch abwechslungsreicher gemacht hat. Herrn Matthias
Paul danke ich für die anregenden Gespräche und (fast) ausgeglichenen
Tischtennispartien.
Für die Betreuung der Computer- und Server-Systeme, die für diese Arbeit
verwendet wurden, bedanke ich mich bei Dr. Dirk Bender und Frau Sabine
Irmer.
Prof. Dr. Michael Schmitt möchte ich für die Anfertigung des Zweitgutacht-
ens danken.
Mein herzlicher Dank gilt besonders meiner Familie, die mich bei all
meinen Vorhaben und meinem Studium im Allgemeinen stets unterstützt
und motiviert hat. Schließlich danke ich meinen Freunden und insbesondere
Nina Hagmeyer für die große Hilfe v.a. bezüglich des Korrekturlesens dieser
Arbeit.
Ein großer Dank geht an alle Stiftungen, die mich während meines Stu-
diums nicht nur finanziell, sondern auch ideell gefördert haben. Ich danke
daher der August-Wilhelm-von-Hofmann-Stiftung, der Konrad-Adenauer-
Stiftung und der Studienstiftung des deutschen Volkes.