169
Theoretical Investigations of the Photophysical Properties of Chromophoric Metal-Organic Frameworks D I S S E R T A T I O N zur Erlangung des akademischen Grades Doctor rerum naturalium (Dr. rer. nat.) vorgelegt dem Bereich Mathematik und Naturwissenschaften der Technische Universität Dresden von M.Sc. Kamal Batra geboren am 12.01.1991 in New Delhi / India Verteidigt am 17. Februar 2021 Die Dissertation wurde in der Zeit von Mai 2017 bis Dezember 2020 am Institut für Theoretische Chemie angefertigt.

Theoretical Investigations of the Photophysical Properties

  • Upload
    others

  • View
    7

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Theoretical Investigations of the Photophysical Properties

Theoretical Investigations of the

Photophysical Properties of Chromophoric

Metal-Organic Frameworks

D I S S E R T A T I O N

zur Erlangung des akademischen Grades

Doctor rerum naturalium

(Dr. rer. nat.)

vorgelegt

dem Bereich Mathematik und Naturwissenschaften

der Technische Universität Dresden

von

M.Sc. Kamal Batra

geboren am 12.01.1991 in New Delhi / India

Verteidigt am 17. Februar 2021

Die Dissertation wurde in der Zeit von Mai 2017 bis Dezember 2020

am Institut für Theoretische Chemie angefertigt.

Page 2: Theoretical Investigations of the Photophysical Properties

Dekan: Prof. Dr. Thomas Henle

1. Gutachter: Prof. Dr. Thomas Heine

2. Gutachter: Dr. Renhao Dong

Page 3: Theoretical Investigations of the Photophysical Properties

Die Dissertation wurde in der Zeit von Mai 2017 bis Dezember 2020 an der Technischen

Universität Dresden unter der Betreuung von Prof. Dr. Thomas Heine durchgeführt.

Ich danke Prof. Dr. Thomas Heine für seine wissenschaftliche und persönliche Unterstützung

sowie für seine ständige Diskussionsbereitschaft.

Page 4: Theoretical Investigations of the Photophysical Properties
Page 5: Theoretical Investigations of the Photophysical Properties

Contents

Abstract……………………………………………………………………………………...01

1. Introduction………………………………………………………………………………03

1.1 Energy Sources……………………………………………………………………………………03

1.2 Photovoltaics (PVs)……………………………………………………………………………….03

1.3 Porphyrins (PPs) and Porphyrinoids……………………………………………………………....05

1.4 Metal-Organic Frameworks (MOFs)……………………………………………………………...08

1.5 Surface-mounted Metal-Organic Frameworks (SURMOFs)……………………………………...10

1.6 Porphyrin based Metal-Organic Frameworks……………………………………………………..11

Motivation, Objective, and Outline of Thesis…………………………………………………………13

2. Theory and Methods……………………………………………………………………...15

2.1 Quantum Chemistry…………………………………………………………………….................16

2.1.1 The Schrödinger Equation (SE)………………………………………………………...16

2.1.2 The Born-Oppenheimer (BO) Approximation………………………………………….17

2.2 Density Functional Theory (DFT)………………………………………………………………...18

2.2.1 The Hohenberg-Kohn (HK) Theorems………………………………………................19

2.2.2 The Kohn-Sham (KS) Formalism……………………………………………................19

2.2.3 Exchange-Correlation Functionals……………………………………………………...21

2.2.3.1 Local Density Approximation (LDA)………………………………………..22

2.2.3.2 Generalized Gradient Approximations (GGAs)……………………..............23

2.2.3.3 Global Hybrids (GHs)………………………………………………………..23

2.3 Time Dependent Density Functional Theory (TD-DFT)……………………………….................24

2.3.1 The Runge-Gross (RG) Theorem and TD KS Formalism……………………………...25

2.3.2 The Linear Response (LR) Formalism ………………………………………...............27

2.3.2.1 The LR TD-DFT and Tamm-Dancoff Approximation (TDA) ……………...27

2.3.2.2 The simplified TDA and TD-DFT approaches………………………………29

2.4 Density Functional Based Tight Binding (DFTB) ………………………………………………..30

2.5 Periodic Treatment of Crystalline Materials………………………………………………………33

2.6 Applied Computational Chemistry Packages……………………………………………………..37

3. Benchmarking the Performance of Time Dependent Density Functional Theory for

Predicting the UV-Vis Spectral Properties of Porphyrinoids……………………………39

3.1 Introduction………………………………………………………………………………………..41

Page 6: Theoretical Investigations of the Photophysical Properties

3.2 Computational Methodologies………………………………………………………….................45

3.3 Results and Discussions…………………………………………………………………………...48

3.3.1 GGA and meta-GGA Functionals……………………………………………................49

3.3.2 Global Hybrid Functionals……………………………………………………………...51

3.3.3 Range Separated Hybrid (RSH) Functionals…………………………………...............53

3.3.4 Double Hybrid Functionals and post-Hartree Fock (HF) Approaches…………………57

3.3.5 Influence of Diffuse Basis Set and Ground State Structure…………………………….59

3.4 Conclusions………………………………………………………………………………………..60

4. Computational Screening of Surface-mounted Metal Organic Frameworks

Assembled from Porphyrins………………………………………………………………..63

4.1 Introduction………………………………………………………………………………………..66

4.2 Computational Methodologies………………………………………………………….................69

4.3 Results and Discussions…………………………………………………………………………...70

4.4 Conclusions………………………………………………………………………………………..74

5. The Proximity effect in Porphyrin-based Surface-mounted Metal-Organic

Frameworks…………………………………………………………………………………77

5.1 Introduction………………………………………………………………………………………..79

5.2 Computational Methodologies………………………………………………………….................83

5.3 Results and Discussions…………………………………………………………………………...84

5.4 Conclusions………………………………………………………………………………………..86

6. Computational Screening of Phthalocyanine-based Surface-mounted Metal-Organic

Framework…………………………………………………………………………………..89

5.1 Introduction………………………………………………………………………………………..91

5.2 Computational Methodologies………………………………………………………….................94

5.3 Results and Discussion………………………………………………………………….................95

5.4 Conclusions………………………………………………………………………………………..99

7. Summary………………………………………………………………………………...101

A. Acronyms………………………………………………………………………………..109

B. Appendices………………………………………………………………………………111

B.1 Supporting Information of Chapter 3…………………………………………………...111

B.2 Supporting Information of Chapter 4…………………………………………………...129

C. Bibliography…………………………………………………………………………….145

D. Acknowledgement………………………………………………………………………161

Page 7: Theoretical Investigations of the Photophysical Properties

1

Abstract

For inorganic semiconductors such as silicon, crystalline order leads to bands in the electronic structure

which give rise to drastic differences with respect to disordered materials. Distinct band features lead

to photo-effect, and the band structure can be tuned to optimize the performance of the photovoltaic

(PV) device. An example is the presence of an indirect band gap. For organic semiconductors, such

effects are typically precluded, since most organic materials employed are disordered, which hampers

their characterization and theoretical analysis.

The inspiration for this thesis came from the very first evidence of an indirect band gap exhibited by

highly ordered and crystalline porphyrin-based surface-mounted metal-organic framework (PP-based

SURMOF) material [J. Liu et al. Angew. Chem. Int. Ed. 2015, 54, 7441]. The presence of an indirect

band gap should in principle result in suppressed charge recombination and efficient charge separations

which would significantly enhance the PV device performance. However, the energy gain from the

electronic band dispersion in the reported Pd-PP-Zn-SURMOF is far too low (≈5 meV) and results in a

very low photocurrent generation (efficiency 0.2%), which is certainly not sufficient for the application.

Another noticeable shortcoming is the weakly absorbing Q-bands of the employed PP chromophore

(Pd-metal containing porphyrinoid, Pd-PP) in the visible region of the solar spectrum. Nevertheless,

this novel research has highlighted the potential to improve the photophysical properties of PP-based

SURMOFs by (i) introducing various functional groups or metal ions to the PP-core and (ii) controlling

the PP-stacking behavior in layered materials.

To overcome the posed shortcomings of the PP-MOF prototype PV material and to exploit the potential

of PP-based SURMOFs, we have employed the following approach to increase the light absorption and

the electronic band dispersion. Firstly, we proposed a computationally feasible simplified time-

dependent approach to investigate the light absorption properties of PP derivatives or related PP-

containing materials. Secondly, we predicted the light absorption properties of multi-functionalized PPs

(i.e. tuning the weakly absorbing Q-bands), thus allowing us to identify different PP linkers with

different light absorption properties, allowing to bridge the so-called green gap. Finally, we

incorporated the most promising PP linkers for the construction of SURMOFs and applied state-of-the-

art DFT methods in various approximations to optimize the PP-stacking behavior to achieve the desired

photophysical properties. Besides PPs, we have extended our investigations to phthalocyanines (PCs)

as alternative individual SURMOF building blocks, because they do not only exhibit structural

robustness and stability but also possess enhanced absorption in the visible and the near IR spectral

regions in comparison to PPs. Hence, the exploitation of PCs could enrich the library of SURMOFs

with the desired optical quality.

Page 8: Theoretical Investigations of the Photophysical Properties

2

Page 9: Theoretical Investigations of the Photophysical Properties

3

Chapter 1

Introduction

"I’d put my money on the sun and solar energy. What a source of power!

I hope we don’t have to wait until oil and coal run out before we tackle that." – Thomas Edison

1.1 Energy Sources

Energy has been one of the fundamental prerequisites for human activity and has played a

crucial role in human history. The development and success of human industrialization have

mainly been driven by using fossil fuels as an energy source. However, the continued use of

fossil energy is linked to harmful impacts on the global climate and has forced us to examine

alternative forms of energy sources that are eco-friendly, sustainable, and economical.

Prominent and possible resources that are alternative to fossil fuels include biomass, solar,

wind, geothermal, and nuclear energy. All these possible energy sources can certainly assist us

in reducing our dependence on fossil fuels. However, among all, solar energy has the unique

potential as it is readily available and renewable at the same time to meet a broad scope of

current global energy demand and remains an attractive source of energy.

1.2. Photovoltaics

Harnessing solar energy, to create an artificial source of energy in the form of electricity or as

fuel is one of the top priorities in the 21st century. The direct capture of solar energy by green-

clean technology is a valuable option to reduce human dependence on fossil fuels as an energy

source. Therefore, there is a rapidly growing demand for devices that convert solar energy into

Page 10: Theoretical Investigations of the Photophysical Properties

4

electrical energy. Today the highest conversion efficiencies are achieved with photovoltaic

(PV) devices based on inorganic semiconductors like silicon.1 Apart from that, there are many

other promising PV technologies in their rising stage. These mainly include dye-sensitized

solar cells,2 organic solar cells,3 and perovskite solar cells.4 Organic solar cells, also known as

organic photovoltaics (OPVs), are very promising as there is huge interest in the use of organic

materials because they offer a less expensive and non-toxic alternative to silicon and other thin-

film PV technologies.

Although the first successful OPV device was reported in 1986 by Tang et al.5, progress in

improving their efficiency, stability, and scalability is slow, since the search for organic

molecules that are suited as active material in OPV devices is to a large extent still controlled

by empirical approaches. Most organic materials employed in OPV devices are disordered,

which hampers their characterization and theoretical analysis. To overcome these problems,

the employment of suitable crystalline systems with well-defined structure is sought after, as

this would enable the OPV device characteristics to be understood on the grounds of high-level

Figure 1.1 Schematic representation of recombination process of free electrons and holes in a

solid with an indirect and direct band gap. The hole-electron recombination is suppressed in an

indirect band gap solid.

Page 11: Theoretical Investigations of the Photophysical Properties

5

electronic structure calculations. Finding such well-defined structures will unlock a promising

new path to the design of optoelectronic materials. The most relevant earlier work6 has

suggested that the formation of an indirect band gap (as schematically shown in Figure 1.1)

might be indeed possible for highly ordered organic materials and represents a condition which

is anticipated to significantly enhance the device performance. It is important to note that in

the case of an indirect band gap, there is fast and highly efficient charge separation and the

charge carrier recombination is strongly suppressed. The advantage of having a regular

arrangement of organic molecules for light-harvesting is demonstrated by exploiting

porphyrins, a class of organic molecules, which are present in many biological systems and

have attracted much attention in light-harvesting or PV applications.7-10

1.3 Porphyrins and Porphyrinoids

Porphyrins (PPs) and their derivatives (also called as Porphyrinoids) are a group of heterocyclic

macrocycle organic compounds derived from the simplest free base PP named “porphin”, to

which a variant of substituents can be connected. Porphin consists of four modified pyrroles

(5-membered organic ring) subunits connected by methine bridges (=CH-) as shown in Figure

1.2. The structure of PP has a stable configuration of single and double bonds with aromatic

Figure 1.2 The unsubstituted porphyrin macrocycle (porphin).

Page 12: Theoretical Investigations of the Photophysical Properties

6

character, resulting in a highly conjugated molecule containing 18π-electrons. The PP core is

a tetradentate ligand in which the space available for metal coordination has a maximum

diameter of approximately 3.7 Å. During the metalation of the PP macrocycle, two protons are

removed from the pyrrole nitrogen atoms, leaving two negative charges,11 which could bind

with transition metal ions to form metal-porphyrins (M-PPs). In the PP macrocycle, there are

two distinct sites, the so-called meso- and β-pyrrole positions (Figure 1.2), where electrophilic

substitution can take place with different reactivity.12

The light absorbing properties of PPs constitute one of their most fascinating attributes, which

can be examined by Ultraviolet-Visible (UV-Vis) spectroscopy. The typical absorption

spectrum of PPs (e.g. tetraphenyl PP as shown in Figure 1.3a) consists of a sharp, strongly

intense Soret-band, which is typically located in the near UV region (350 to 450 nm), and four

weakly intense Q-bands located in the visible region (450 to 700 nm). In principle, all the

transitions between the frontier orbitals are allowed based on symmetry rules. However, as the

both highest occupied molecular orbitals (HOMO-1 and HOMO) as well as both lowest

unoccupied molecular orbitals (LUMO and LUMO+1) are nearly degenerated, the electronic

Figure 1.3 a) Typical spectrum of a porphyrinoid consists of a strong Soret-band and four weak Q-

bands in which two of them are transition between the ground states Qx,y(00) and two from the

ground state to the single-excited vibrational state Qx,y(01). Blue lines in the Lewis structure

highlight the π-electrons on which a 18π cyclic polyene model is reasonable; b) Comparison of the

frontier orbitals obtained from DFT calculations of porphyrin and the 18π cyclic polyene model.

Page 13: Theoretical Investigations of the Photophysical Properties

7

structure can be approximated by a 18π cyclic polyene model, as suggested by Gouterman,13

where two transitions are allowed between the degenerated frontier orbitals, while two are

forbidden (Figure 1.3b). In Gouterman’s model, the Q-band intensity is almost negligible. Any,

Q-band intensity is due to the distortion of Gouterman’s perfect 18π-electron system. Thus,

two bands arise due to the transition between the ground states Q(0,0) and from the ground

state to the single-excited vibrational state Q(0,1). The presence of the NH protons breaks the

symmetry and further split into two bands each in the no longer degenerated x and y

components: Qx(0,0), Qy(0,0), Qx(0,1), and Qy(0,1) as indicated in Figure 1.3a. The Q-bands

arise from the vibrational coupling are not visible in a standard calculation. Furthermore, a

strong mixing of the transitions was observed for Soret and Q-bands by quantum chemical

studies.14-15 In order to adjust the positions and intensities of characteristic Q-bands, the energy

levels and shapes of frontier orbitals must be tuned. The larger the energy gap between HOMO-

1 and HOMO as well as LUMO and LUMO+1, the stronger will be the absorption intensity of

the Q-bands and vice-versa. Hence, the tuning of energy level can be performed through the

rational design of PP skeleton by modifying the π-conjugation, ring functionalization as well

as inclusion/change of a central metal atom.

Compared to many other photoactive molecules, PPs are not only very effective in

transforming photons into electronic excitations but are also rather stable. In the natural

photosystems, the PPs are usually stacked to yield extended columns along which the excitons

can be transported and the most important operation takes place at a target chlorophyll center,16

where dissociation of the exciton yields an electron and a hole through a process known as

charge separation. Given all that, PPs are among the best-performing organic molecules

regarding photon absorption and charge separation. Numerous previous investigations with

variant approaches17-19 (e.g. vapor phase deposition or self-assembly) have been performed

with the goal to develop well-defined, thin PP layers on various substrates, but the resulting

Page 14: Theoretical Investigations of the Photophysical Properties

8

systems do not exhibit a high degree of ordering. In the present thesis, we propose to use a

rather unconventional approach based on metal-organic frameworks (MOFs). The basic idea is

to use PP-containing organic linkers, which are coordinated to metal-or-metal/oxo nodes to

yield crystalline, porous MOFs. In addition to PPs, we would like to extend our studies and

introduce phthalocyanine (PC) as an alternative organic linker. PCs are consisting of four iso-

indole units linked together through nitrogen atoms. PCs possess a 18π-electron aromatic cloud

delocalized over an arrangement of alternating carbon and nitrogen atoms. The replacement of

the meso carbons in PPs by nitrogen linkages, combined with four fused benzo rings

significantly breaks the degeneracy of the frontier orbitals. PCs are not only owing structural

robustness and stability but also possess red-shift of the Q-bands with a significantly increased

absorption intensity as compared to PPs (more details are described in Chapter 6). Also, it is

worth mentioning that the synthesis of PCs is more exhausting and challenging than that of

PPs. This is reflected by the rough numbers of successful PC syntheses, which amounts to over

100,000 for PPs (according to Sci-Finder), but only to 20,000 for PCs. Given all the facts and

figures, we propose to extend the idea of PP-MOF to PC-MOF. This will enable the extension

of MOF application areas to photo-electrochemistry, where the conditions are much harsher.

In the following section, the MOF approach will be described in more detail.

1.4 Metal-Organic Frameworks

Metal-Organic Frameworks (MOFs), also known as porous coordination polymers (PCPs), are

a class of two- or three-dimensional (2D or 3D) materials assembled through coordination

bonds between organic linkers and inorganic connectors of metal ions or clusters (Figure

1.4).20-22 Although the first 3D coordination polymer was published by Saito and co-workers

in 1959,23 the first MOF constructed by organic linkers and metal-ions resulting in a structure

with potential voids was introduced by Yaghi et al.24-26 in the late 1990s. Since then, thousands

Page 15: Theoretical Investigations of the Photophysical Properties

9

of MOF structures have been reported with a variety of constituents, geometry, size, and

functionality.27

MOFs show an extraordinarily high porosity with up to 9.000 m2/g combined with a high

crystallinity has not been surpassed by any other porous material till date.28 These unique

properties have opened the door for the MOF research to grow significantly, and consequently,

MOFs have attracted considerable attention in many applications, including gas storage,29-30

gas separation,31-33 sensing,34-35 catalysis,36-38 and drug-delivery.39 Besides conventional

applications of 3D (bulk) MOFs as materials for gas storage or separation, recently their

optical,40-42 electrical,43-45 and magnetic properties46-48 have attracted increasing attention.

However, the bulk MOF materials are not well suited for investigating the optical properties

since the light scattering by the powder particles/pellets hamper a thorough optical

characterization. For these purposes, MOF thin films are much better suited and rapidly

growing area. Recently, in a review article Fischer et al.49 distinguished the MOF thin films

into two classes: polycrystalline and surface-mounted MOFs (SURMOFs). The polycrystalline

films are generally regarded as an assembly of randomly oriented MOF crystals on a surface.

The thickness of these films is usually related to the size of MOF crystals/particles and ranging

up to micrometers. The second class of MOF thin film is emerging, referred to as SURMOFs.

These films consist of ultrathin (in the nanometer range) MOF multilayers that are perfectly

Figure 1.4 Schematic of the construction of a 3D MOF via the reticular chemistry approach:

stitching of organic linkers and metal connectors to build metal-organic frameworks.

Page 16: Theoretical Investigations of the Photophysical Properties

10

oriented and homogeneously deposited on the surface. In the following section, the SURMOF

approach will be described in some more details.

1.5 Surface-mounted Metal-Organic Frameworks

SURMOFs are crystalline, highly oriented MOF thin films which are deposited on the surface

of a given substrate by employing Liquid Phase Epitaxial (LPE) growth scheme in layer-by-

layer (LbL) fashion, introduced by Wöll and Fischer50-51 in 2007. The most notable difference

of LPE-approach compared to the conventional solvothermal method52 yielding MOF powder

is the stepwise growth of the MOFs on a –COOH-functionalized surface, by repeated

immersion cycles, first in a solution of the metal precursor and subsequently in a solution of an

organic linker, as shown in Figure 1.4. Another notable difference, it is mandatory to deposit

the MOFs on conducting and, if possible, transparent substrates for being able to investigate

photophysical properties. Also, it is worth mentioning that not all MOF types are well-suited

for the LPE-approach. For example, attempts to grow ZIF-8 thin films were unsuccessful unless

a special synthetic route had to be formulated (replacing the ethanol solvent, used in most other

MOF thin film syntheses by methanol).53 Nonetheless, in the case of PP-based SURMOFs, the

LPE approach has already been successfully applied to fabricate thin films, first by J. Hupp

and co-workers, and later by C. Wöll and co-workers. To extend the idea of PP-MOF thin films,

only very few papers concerning PC-MOFs have been published to date54-56 and as per our

knowledge, no previous literature was found for PC-MOF thin films. To highlight the fact of

insufficient results is substantial experimental effort requires to synthesize PC compounds, and

especially functionalized PCs (more details are described in Chapter 6).

In contrast to other MOF thin films synthesized by different techniques (e.g. mother-solution,

dip coating, electrochemical growth, etc.), SURMOFs exhibit several unique properties,57-59

Page 17: Theoretical Investigations of the Photophysical Properties

11

namely: (a) monolithic, crystalline, and highly-oriented MOF layers; (c) smooth and

homogeneous morphologies; (b) controllable thickness and growth orientations; (d) low

defects densities; (e) allows well-defined MOF-on-MOF multilayer systems. It is also worth to

mention that the templating effect gives typically higher-symmetry structures compared to the

solvothermal approach (e.g. P4 for SURMOF-2).60 These properties of SURMOFs make them

perfectly suitable candidates for membrane separations, sensor techniques, electronic devices

etc.61-63 More importantly, recent works in the groups of Wöll and Heine have demonstrated

the successful fabrication of crystalline, highly orientated PP-MOF thin films and their use as

a platform in the application of light-harvesting and conversion of solar energy. In the

following section, a state of art for PP-MOF thin films will be discussed in a detailed manner.

1.6 Porphyrin based Metal Organic Frameworks

In the past decades, PPs and their derivatives have attracted much attention as an interesting

platform in the construction of MOFs. The first PP-based coordination polymer was reported

in 1991,64 however the first PP-MOF was successfully synthesized in 2006 by self-assembly

of a 2D coordination network. The characteristics of N2 adsorption has indicated that this

coordination network has uniform micropores and gas adsorption cavities.65 Since then, an

exponential increase in the synthesis of various kinds of PP-MOF structures has been

Figure 1.4 Schematic of the layer-by-layer (LbL) growth of MOFs on a functionalized substrate by

repeated cycles of immersion in solution of metal precursor and subsequently solution of an organic

linker. Between steps, the material is rinsed with solvent and typical thickness range is 100-200 nm.

Page 18: Theoretical Investigations of the Photophysical Properties

12

observed.66 As discussed earlier, we are interested in a particular class of MOF thin films,

known as SURMOFs, with regard to their electrical and optical properties and applications. An

excellent description of the present status regarding applications exploiting electrical and

optical properties of MOFs is provided by the review from Falcaro, Allendorf, and Ameloot,67

while in the review by Liu and Wöll, the focus is on MOF thin films.68

H. Kitagawa and co-workers reported the fabrication of a perfect preferentially oriented

nanofilm of MOFs on surface-no.1 (NAFS-1), consisting of Co2+ substituted PP building units.

This was achieved by applying the layer-by-layer growth technique coupled with the Langmuir

Blodgett (LB) method.69 Subsequently, the authors have applied the same technique to prepare

NAFS-2 consisting of free base substituted PP building units coupled with Cu2 paddlewheel

dimers on a surface. The interlayer spacing in NAFS-2 varied while retaining the same in-plane

molecular arrangement by employing different molecular building units than for the previously

reported NAFS-1, and maintains its highly crystalline order above 200 °C.70

J. Hupp and co-workers have fabricated pillar-layered paddlewheel type PP-based MOFs on

functionalized surfaces using the LbL growth technique. The obtained PP-MOF thin films are

preferentially oriented, highly porous, and have controlled thickness. Long-range energy

transfer has been demonstrated for MOF films and the reported findings offer useful insights

for the subsequent fabrication of MOF-based solar energy conversion devices.71

In 2015, a study co-authored by C. Wöll and T. Heine72 reported the first evidence of an indirect

band gap formation in an epitaxial MOF/SURMOF. The investigated SURMOF constructed

by Zn-paddlewheel units and Pd-porphyrinoid linkers using LbL growth technique, and exhibit

photocarrier generation efficiency of 9.510-2. The electronic structure calculation for such a

system displays that layers are stacked in AAA fashion within a square lattice. Furthermore,

this MOF-system reveals a small but distinct dispersion (≈5 meV) in both conduction and

Page 19: Theoretical Investigations of the Photophysical Properties

13

valence band leading to the formation of an indirect band gap. This novel finding demonstrates

that the solid-state properties of PP-based SURMOFs offer a huge potential for OPVs.

Subsequently, C. Wöll and co-workers73 have demonstrated the fabrication of a new class of

epitaxial PP-based SURMOF incorporating electron donor diphenylamine (DPA) groups into

the PP-macrocycle, which exhibits the highest photocarrier generation efficiency of up to

3.010-1 reported so far. Although this value is still rather low, the study has highlighted the

potential for improving the performance of these systems by adding electron donor or acceptor

groups to the PP linkers and by optimizing the device architecture.

Motivation, Objective, and Outline of the Thesis

The motivation of the thesis is led by the very first evidence of an indirect band gap exhibited

by highly ordered and crystalline PP-SURMOF. The photovoltaic efficiency of the reported72

PP-SURMOF amounts to only 0.2% and is thus far too low for realizing a competitive device.

However, the study has highlighted the potential to improve the photophysical properties of

PP-SURMOFs by functionalizing PP-core and optimizing their layered stacks within material.

The overarching focus of the thesis to proceed with an interesting class of photoactive organic

molecules, PPs and their derivatives as organic linkers and to utilize the SURMOF approach

for designing crystalline, highly ordered, and well-defined organic thin films. Considering the

light-harvesting properties of PPs, they can be tuned to maximize the light-to-electricity

conversion (i.e. tuning the positions and intensities of weakly Q-bands). However, tuning of

PPs depends on the side groups (e.g. meso and β-positions) at the PP-core or the presence of

central metal ions, different intermolecular interactions can be either supported or blocked, and

by that the photoelectric and photophysical properties. Undoubtedly, there is a huge range of

combinations possible to substitute the variant of PPs and building SURMOFs out of them.

Here, the computational modelling plays a pivotal role, not only to screen a large library of PPs

Page 20: Theoretical Investigations of the Photophysical Properties

14

and their corresponding SURMOFs, but also circumvents the huge experimental synthesis

effort and resources. Among the large number of possibilities, three particularly interesting PPs

have been identified by combining rational design using computer simulation methods.

Moreover, the computational modelling and screenings of the anticipated PP-SURMOF

structures will also provide an insight for tailoring their photoelectric and photophysical to

achieve the remarkable efficiency of this novel OPV device.

The thesis is organized as follows: The current Chapter (Chapter 1) has introduced the main

phenomena and systems of interest. Chapter 2 introduces the theoretical background and

methods, as well as the computational protocols that have been applied to investigate the posed

problems and achieve the respective goals. In Chapter 3, we investigate the capability of

various variants of time-dependent density functional theory (TD-DFT) for predicting the UV-

Vis spectra of porphyrinoids having a diverse extent of conjugation, ring functionalization, as

well as inclusion/modification of a central metal atom. Chapter 4 lays out how the molecular

and electronic structure of variously substituted PPs can be tuned for increasing their absorption

efficiency with help of the identified TD-DFT approach in Chapter 3. Subsequently, we

demonstrate the power of computational screening methods for the construction of selected

PP-based SURMOF structures with the desired photophysical properties. Chapter 5 introduces

a promising PP-SURMOF modeled in Chapter 4 to exploit its band dispersion characteristic as

a function of the rotation of the employed functional groups with respect to the PP core (in a

bulk framework). In Chapter 6, we extend our studies from PPs to PCs as alternative SURMOF

building blocks and introduce some of the preliminary calculated results for the same. Finally,

the summary, acronyms, appendices, bibliography, and acknowledgement are presented.

Page 21: Theoretical Investigations of the Photophysical Properties

15

Chapter 2

Theory and Methods

"If quantum mechanics hasn't profoundly shocked you, you haven't understood it yet."

– Niels Bohr

In this Chapter, we give a brief survey over quantum chemistry and we review the various

theoretical methods that are utilized in the doctoral studies. In Section 2.1, we introduce the

fundamental Schrödinger equation, which is the basis for nearly all further theories and

methods. Section 2.2 elaborates on the theory behind the quantum chemical ground state

methods. Most importantly, time-independent Kohn-Sham based density-functional theory

(DFT), which is the workhorse for proper description of electronic ground state. Section 2.3

lays out flavors of time-dependent density functional theory (TD-DFT), which can efficiently

and reliably compute the electronic excited states of large molecular or extended biological

systems. In Section 2.4, we discuss the density functional tight binding (DFTB) method which

allows us to determine the electronic structure properties of systems with increasing complexity

such as larger molecules, clusters, and solids. Section 2.5 discusses the treatment of periodic

crystalline solids by exploiting their lattice periodicity in order to calculate their electronic

structure properties efficiently. Finally, in Section 2.6, applied computational chemistry

toolbox are briefly discussed.

Page 22: Theoretical Investigations of the Photophysical Properties

16

2.1 Quantum Chemistry

Quantum chemistry applies quantum mechanics to address numerous aspects and phenomena

associated with the behavior of electrons and thus of chemistry. It aims, in principle, to solve

the Schrödinger equation, postulated by the Nobel Laureate Erwin Schrödinger in 1925, later

published in 1926.74 It represents the most significant landmark in the given field. However,

its complexity for all but the simplest of atoms or molecules requires simplifying assumptions

and approximations, establishing a balance between accuracy and computational effort.

2.1.1 The Schrödinger Equation

In general, the time-independent Schrödinger Equation (SE) is the basis for almost all quantum

chemical methods and used for the estimation of the electronic structural properties of the

molecular and solid-state systems. It can be expressed in equation 2.1 as follows:

𝐻𝛹 = 𝐸𝛹 (2.1)

Here, H is the Hamiltonian operator, which defines all the properties of the system, 𝛹 is the

wave function of the system comprising of n electrons and N nuclei in terms of atomic units,

and E is the total energy of the system. The typical form of the Hamiltonian takes into account

five contributions to the total energy of the system: the kinetic energy of the electrons (𝑇𝑛) and

the nuclei (𝑇𝑁), the attractive potential between the electrons and the nuclei (𝑉𝑛𝑁), and the inter-

electronic (𝑉𝑛𝑛) and the inter-nuclear (𝑉𝑁𝑁) repulsive potentials. Thus, for a system of n+N

interacting particles (electrons and nuclei), the Hamiltonian can be expressed as:

𝐻 = 𝑇𝑛 + 𝑇𝑁 + 𝑉𝑛𝑁 + 𝑉𝑛𝑛 + 𝑉𝑁𝑁 (2.2a)

or

Page 23: Theoretical Investigations of the Photophysical Properties

17

𝐻 = −∑1

2∇𝑖

2𝑛𝑖 − ∑

1

2𝑀𝐼∇𝐼

2𝑁𝐼 − ∑ ∑

𝑍𝐼

ǀ𝑟𝑖 −𝑅𝐼 ǀ 𝑁

𝐼𝑛𝑖 + ∑

1

ǀ𝑟𝑖 −𝑟𝑗 ǀ

𝑛𝑖≠𝑗 + ∑

𝑍𝐼𝑍𝐽

ǀ𝑅𝐼 −𝑅𝐽 ǀ𝑁𝐼≠𝐽 (2.2b)

where ∇2 is the Laplacian acting on particles, 𝑟 stands for the particle position, whereas M, Z

and R stand for the nuclear mass, charge, and positions, respectively. Indices i and j denote

electrons with different spatial positions, while I and J are used for nuclei of different type and

spatial positions.

For any kind of property, the SE (Eq. 2.1) must be solved. However, the analytical solutions

for the SE are only possible for a few simple one-particle systems. While in the case of many-

particle systems, the number of nuclei and electrons are very high, leading to a complex

Hamiltonian operator. In fact, accurate wave function description of such systems is then

practically impossible because of the correlated motion of particles. That is, the Hamiltonian

in equation 2.2 includes pairwise interactions (attraction and repulsion) which means no

particle is moving independently. To simplify the problem, it is important to employ some

approximations, and in this context, the Born-Oppenheimer approximation is quite useful.

2.1.2 The Born-Oppenheimer Approximation

In Born-Oppenheimer (BO) approximation, we consider the speed of the nuclei and electrons.75

The nuclei speed is very less compared to the electrons as the nucleus is ~1800 times much

heavier than the mass of the electron for a hydrogen-like atom. Hence, the nucleus is presumed

to be stationary (or in rest) with respect to electrons. As a result, the total wave function can be

separated into two components: one is electronic and other nuclear wave function.

Consequently, we focus on the electronic wave function in which the electronic Hamiltonian

operates can be expressed as:

𝐻 = 𝑇𝑛 + 𝑉𝑛𝑛 + 𝑉𝑛𝑁 (2.3a)

Page 24: Theoretical Investigations of the Photophysical Properties

18

or

𝐻 = −∑1

2∇𝑖

2𝑛𝑖 + ∑

1

ǀ𝑟𝑖 −𝑟𝑗 ǀ

𝑛𝑖≠𝑗 − ∑ ∑

𝑍𝐼

ǀ𝑟𝑖 −𝑅𝐼 ǀ 𝑁

𝐼𝑛𝑖 (2.3b)

In general, the BO approximation is an extremely mild one and not fully sufficient to solve of

the SE for larger, more realistic, but also computationally demanding many-electron systems.

Hence, the equation 2.3 requires even further modifications. In this context, other

approximations76-77 such as Hartree-Fock theory (HF theory), density functional theory (DFT)

etc. are extremely useful. In HF theory, the wave function for an n-electron system is generally

approximated by the linear combination of Hartree products of orbitals within a Slater

determinant78 which is essential for fulfilling the anti-symmetric behavior of particles.

However, the major drawback of HF theory is the absence of electron-electron correlation in

its single determinant wave function, which results in large deviations when it comes to

comparing with experimental data. To address this drawback, more accurate wave function-

based methods such as configuration interaction (CI) and other post-HF methods are available

which explicitly include electron-electron correlation. However, the high computational cost

to accuracy ratio limits their further usage. In this case, the DFT is an alternative method that

not only recovers the electron-electron correlation but quite efficient as well.

2.2 Density Functional Theory

DFT is a well-established quantum-mechanical method for efficiently solving the SE of

complex many-electron systems. It has gained huge popularity among theoreticians, to

calculate the electronic structure and properties of the many-electron system by consideration

of its electron density. Moreover, it has proven based on the pragmatic observation that it is

less computationally demanding than other methods with similar accuracy for various

applications in physics and chemistry. This section provides a detailed introduction to the

Page 25: Theoretical Investigations of the Photophysical Properties

19

basics of DFT as it is used extensively in this thesis to calculate the electronic structure

properties of molecular and extended systems with respect to their light-harvesting properties

and solar energy conversion-based applications.

2.2.1 The Hohenberg-Kohn Theorems

The theoretical background of DFT was first established in 1964 by two major Hohenberg-

Kohn (HK) based theorems.79 The 1st HK theorem states that the ground state properties of

many-electron system are uniquely determined by the ground state electron density ρ(r ) that

relies on only three spatial coordinates. It sets down the basis for reducing the many-body

problem of n electrons with 3n spatial coordinates to three spatial coordinates, through the use

of functionals of the electron density. The 2nd HK theorem introduces a term universal total

energy functional 𝐸[𝜌(𝑟 )] with respect to electron density 𝜌(𝑟 ) under the external potential of

𝑉𝑒𝑥𝑡(𝑟 ). The term 'universal' means that this functional can be applied to any electronic system

independently of 𝑉𝑒𝑥𝑡(𝑟 ). The functional is written in equation 2.4 as follows:

𝐸[𝜌(𝑟 )] = 𝐹𝐻𝐾[𝜌(𝑟 )] + ∫𝑉𝑒𝑥𝑡(𝑟 )𝜌(𝑟 )𝑑𝑟 (2.4)

where, 𝐹𝐻𝐾[𝜌(𝑟 )] contains the kinetic energy and the electron-electron interactions. To sum-

up, from the above-equation, the exact ground-state energy of the many-body system can be

determined variationally by minimizing the total energy as a functional of the electron density.

However, the explicit forms of the two terms which compose the functional 𝐹𝐻𝐾[𝜌] are

unknown. Therefore, further improvement of the theory is needed.

2.2.2 The Kohn-Sham Formalism

In 1965, Kohn and Sham proposed a new formalism80 by introducing the concept of a fictitious

system of non-interacting particles. The scheme was to map a system of interacting electrons

onto a system of non-interacting electrons in an effective potential 𝑉𝑒𝑓𝑓(𝑟 ), known as the Kohn-

Page 26: Theoretical Investigations of the Photophysical Properties

20

Sham (KS) potential, which is constructed such that the density of the fictitious system equals

the density of the real, interacting system. Thus, in KS-DFT, the electronic wave function of

the fictitious system is then represented by a single Slater determinant, consisting of the single

electron wave functions or KS orbitals 𝜓𝑖(𝑟 ). The electron density can thus be expressed in

equation 2.5 as follows:

𝜌(𝑟 ) = ∑ |𝜓𝑖(𝑟 )|2

𝑖 (2.5)

According to the KS-DFT scheme, the total energy functional can then be calculated as follows:

𝐸𝐾𝑆−𝐷𝐹𝑇 = 𝑇𝐾𝑆 + 𝐸𝐻 + 𝐸𝑥𝑐 + 𝐸𝑒𝑥𝑡 (2.6)

Figure 2.1 Flow chart of a typical time-independent KS-DFT calculation

Initial Guess

Calculate Effective Potential

Solve Kohn Sham Equation

Calculate Electron Density

Convergence ?No

Yes

Output

Time-Independent Properties

Page 27: Theoretical Investigations of the Photophysical Properties

21

Here, 𝑇𝐾𝑆 is the kinetic energy of the non-interacting system of electrons, 𝐸𝐻 is the classical

Coulomb (or Hartree) energy corresponding to electron-electron interaction, 𝐸𝑥𝑐 is the

unknown exchange-correlation interaction that considers all non-classical many-body effects

between electrons, and 𝐸𝑒𝑥𝑡 is the energy from the external field due to the positively charged

nuclei. Furthermore, one can rewrite the KS equation in a wave function form as follows:

[−1

2∇2 + 𝑉𝑒𝑓𝑓(𝑟 )]𝜓𝑖(𝑟 ) = 𝐸𝑖𝜓𝑖(𝑟 ) (2.7)

Here, the effective single-particle potential, 𝑉𝑒𝑓𝑓(𝑟 ) represents the addition of classical

Coulomb potential 𝑉𝐻(𝑟 ), exchange-correlation potential 𝑉𝑥𝑐(𝑟 ) =𝛿𝐸𝑥𝑐[𝜌(𝑟 )]

𝛿𝜌(𝑟 ), and the external

potential 𝑉𝑒𝑥𝑡(𝑟 ). To sum-up, the KS formalism allows to calculate the ground-state energy

and density exactly (see flow-chart in Figure 2.1). However, the exact form of the exchange-

correlation energy, 𝐸𝑥𝑐, is unknown yet, so further approximations are needed to enable

calculations practically in the framework of KS-DFT.

2.2.3 Exchange-Correlation Functionals

The exchange-correlation (XC) energy functional, 𝐸𝑥𝑐[𝜌(𝑟 )], defined in the KS-DFT can be

expressed as a sum of two terms. One is exchange 𝐸𝑥[𝜌(𝑟 )] and another is correlation part

𝐸𝑐[𝜌(𝑟 )] as follows in equation 2.8.

𝐸𝑥𝑐[𝜌(𝑟 )] = 𝐸𝑥[𝜌(𝑟 )] + 𝐸𝑐[𝜌(𝑟 )] (2.8)

There are different approximations to the XC energy functional available and in the following

subsection, we will describe them in an order of increasing complexity.

Page 28: Theoretical Investigations of the Photophysical Properties

22

2.2.3.1 Local Density Approximation

The simplest XC functional depends only on the electron density and occupies the first rung of

Jacob’s Ladder (as shown in Figure 2.2) known as local density approximation (LDA). In this

approximation, a homogeneous electron gas (HEG) is considered, and can be expressed in

equation 2.9 as follows:

𝐸𝑥𝑐[𝜌(𝑟 )] ≈ 𝐸𝑥𝑐𝐿𝐷𝐴[𝜌(𝑟 )] = ∫𝑑3𝑟 𝜌(𝑟 ) Ɛ𝑥𝑐

𝐻𝐸𝐺[𝜌(𝑟 )] (2.9)

where, Ɛ𝑥𝑐𝐻𝐸𝐺[𝜌(𝑟 )] is the XC energy density for a homogeneous electron gas that can, then, be

solved individually as follows:

Ɛ𝑥𝑐𝐻𝐸𝐺 = Ɛ𝑥

𝐻𝐸𝐺 + Ɛ𝑐𝐻𝐸𝐺 (2.10)

LDA mostly works well for materials, where electron density varies slowly, but also displays

shortcomings and large inaccuracies in cases of energetics details, since most real systems have

inhomogeneous density distributions. Therefore, an attempt to correct shortcoming of the LDA,

is addressed by the so-called Generalized Gradient Approximation (GGA).

Figure 2.2 Schematic representation of five different rungs of Jacob’s Ladder (1-5)

Page 29: Theoretical Investigations of the Photophysical Properties

23

2.2.3.2 Generalized Gradient Approximations

GGA employs an ingredient that can account for inhomogeneities in the density distributions:

the density gradient, 𝛻ρ in real systems. The GGA functionals occupy the second rung of

Jacob’s Ladder (see Figure 2.2) and the general form is given in equation 2.11 as follow:

𝐸𝑥𝑐𝐺𝐺𝐴[𝜌(𝑟 )] = ∫ 𝑑3𝑟 𝜌(𝑟 ) Ɛ𝑥𝑐

𝐺𝐺𝐴(𝜌(𝑟 ), 𝛻𝜌(𝑟 )) (2.11)

where, Ɛ𝑥𝑐𝐺𝐺𝐴(𝜌(𝑟 ), 𝛻𝜌(𝑟 )) is the XC energy per electron gradient that can, then, be solved

individually as follows:

Ɛ𝑥𝑐𝐺𝐺𝐴 = Ɛ𝑥

𝐺𝐺𝐴 + Ɛ𝑐𝐺𝐺𝐴 (2.12)

The GGA functionals are improved significantly over the LDA and work well for the real

systems, where electron density varies rapidly. However, to further improve their accuracy,

two additional local ingredients such as either the Laplacian of the electron density, ∇2𝜌(𝑟 ), or

the kinetic energy density, 𝜏(𝑟 ) = ∑1

2|∇ 𝜓𝑖

(𝑟 )|2𝑖 can be utilized. From a chemical point of

view, the kinetic energy density is by far more popular and useful ingredient for predicting

charge delocalisation.81 To sum up, XC functionals that rely on either of above-mentioned local

ingredients are known as meta-generalised gradient approximations (meta-GGA or mGGA)

and occupy the third rung of Jacob’s Ladder (see Figure 2.2).

2.2.3.3 Global Hybrids

Despite the systematic improvement offered by the inclusion of physical ingredients (𝛻𝜌, ∇2𝜌,

and 𝜏), gradient corrected functionals are still prone to originate systematic errors due to self-

interaction error (SIE) limitation. 82 The simplest way to explain self-interaction (i.e. electrons

interact with themselves) is to consider the HF description of the one-electron system (e.g. H-

atom), where the electron-electron interaction energy should be zero. That is, the classical

Page 30: Theoretical Investigations of the Photophysical Properties

24

“Coulomb” and the non-classical “exchange” energy terms cancel out exactly, making HF one-

electron SIE-free. In case of KS-DFT, since the exact exchange term is replaced by the XC

functional, most functionals are not SIE-free (i.e. 𝐸𝐻[𝜌(𝑟 )] + 𝐸𝑥𝑐[𝜌(𝑟 )] ≠ 0 for the H-atom)

To address the SIE problem, an alternative scheme is to replace the local exchange functional

with the exact exchange functional, while taking a local correlation functional that yields

correlation energy exactly zero for one-electron systems. However, early trials to combine

exact exchange with local correlation of DFT were unsuccessful. In the early 1990s, Becke

introduced an imperfect yet highly effective scheme83 of mixing only a global fraction of exact

exchange with the XC functional from considerations of the adiabatic connection formula.84-87

This successful functional scheme defines the Rung 4 of Jacob’s Ladder (Figure 2.2) and

known as the global hybrid (GH) which take the form given in equation 2.13,

Ɛ𝑥𝑐𝐺𝐻 = 𝑐𝑥Ɛ𝑥

𝐻𝐹 + (1 − 𝑐𝑥) Ɛ𝑥𝐷𝐹𝑇 + Ɛ𝑐

𝐷𝐹𝑇 (2.13)

While given the ongoing rapid pace of functionals development, there is a plethora of XC

functionals available in most of the quantum chemistry codes, so many, in fact, that it would

be impossible to cover all of them in this thesis. The choice of functional mostly depends on

the system of interest and the property to be evaluated. Here, for an overview and extensive

assessment of different rungs of XC functionals (including range-separation or double hybrids),

we would like to refer this excellent review article by N. Mardirossian and M. Head-Gordon.88

2.3 Time-Dependent Density Functional Theory

As mentioned in the previous section, time-independent KS-DFT provides quite an accurate

ground-state electronic structure. So, basic questions from a theoretical chemist and physicist

perspective would be: Is there any time-dependent analogue of the most fundamental HK

theorems and KS equation? Also, what about the characteristic properties which are evolving

Page 31: Theoretical Investigations of the Photophysical Properties

25

with time, to name a few, such as UV-Vis spectroscopy, non-linear optics, and photochemistry?

To address these questions, time-dependent density functional theory (TD-DFT) extends the

ideas based on ground-state DFT for the treatment of excited-state and the time-dependent

phenomena.89 In the following, two possible strategies to obtain the excited-state energies and

spectra are discussed. The first one is to propagate the time-dependent KS formalism, referred

to as real-time TD-DFT,90-91 and the other one is the time-dependent linear-response

formalism,92 which rapidly took off the application of TD-DFT.

2.3.1 The Runge-Gross Theorem and Time-Dependent Kohn Sham Formalism

The first theoretical foundation of TD-DFT was established in 1984 by the Runge-Gross (RG)

theorem,93 which is a time-dependent analogue of the HK theorem. It states that for a system

initially in its stationary ground-state exposed to time-dependent perturbation, the time-

dependent charge density, 𝜌(𝑟 , 𝑡), determines the time-dependent external potential, 𝑉𝑒𝑥𝑡(𝑟 , 𝑡)

up to a spatially constant function of time, 𝐶(𝑡), and thus a time-dependent wave function can

be expressed as follows:

𝛹(𝑟 , 𝑡) = 𝛹[𝜌(𝑡)](𝑡)𝑒−𝑖𝛼(𝑡) with (𝑑

𝑑𝑡)𝛼(𝑡) = 𝐶(𝑡) (2.14)

The RG theorem implies that there is one-to-one correspondence of the time-dependent

external potential and time-dependent density, which constitutes the foundation for the time-

dependent KS equations, (analogy to the time-independent KS equations 2.7) as follows:

𝑖𝜕

𝜕𝑡𝜓𝑖(𝑟 , 𝑡) = [−

1

2∇2 + 𝑉𝑒𝑓𝑓

𝑇𝐷 (𝑟 , 𝑡)] 𝜓𝑖(𝑟 , 𝑡) (2.15)

Here 𝑉𝑒𝑓𝑓𝑇𝐷 is the effective single-particle potential (extended from time-independent KS-DFT)

evolving with time is the sum of classical Coulomb potential 𝑉𝐻(𝑟 , 𝑡), exchange-correlation

potential 𝑉𝑥𝑐(𝑟 , 𝑡), and the external potential 𝑉𝑒𝑥𝑡(𝑟 , 𝑡). It can be expressed as follows:

Page 32: Theoretical Investigations of the Photophysical Properties

26

𝑉𝑒𝑓𝑓𝑇𝐷 (𝑟 , 𝑡) = 𝑉𝐻(𝑟 , 𝑡) + 𝑉𝑥𝑐(𝑟 , 𝑡) + 𝑉𝑒𝑥𝑡(𝑟 , 𝑡) (2.16)

The time-dependent density can thus be calculated from the time-dependent KS equations (see

flow-chart in Figure 2.3) by solving equation 2.15, is given by 𝜌(𝑟 , 𝑡) = ∑ |𝜓𝑖(𝑟 , 𝑡)|2

𝑖 . Like

in DFT, the exact functional form of the time dependent XC potential, 𝑉𝑥𝑐(𝑟 , 𝑡)

𝑉𝑥𝑐(𝑟 , 𝑡) =𝛿𝐸𝑥𝑐[𝜌(𝑟 ,𝑡)]

𝛿𝜌(𝑟 ,𝑡) (2.17)

is not known, therefore approximations are required. To address this problem, the simplest

possible approximation of the time-dependent XC potential is the adiabatic local-density

approximation (ALDA). It employs the functional form of the static LDA with a time-

Figure 2.3 Flow chart of a typical time-dependent KS-DFT calculation

Initial Guess

Calculate Effective Potential

Solve Kohn Sham Equation

Calculate Electron Density

Convergence ?No

Yes

Output

Time-Dependent Properties

Page 33: Theoretical Investigations of the Photophysical Properties

27

dependent density, where the value of 𝑉𝑥𝑐(𝑟 , 𝑡) is equal to the time-independent HEG potential.

This approximation allows to use the standard ground-state functionals for the TD-DFT

calculations, i.e. it can also be extended to adiabatic GGA, meta-GGA and hybrids.94

Despite some recent successful applications of the time-dependent KS equations in the

literature,95-96 it is still having the status of an expert’s approach. In the following sub-section,

we introduce the linear response TD-DFT, which is more convenient and easily implemented

in the most standard quantum chemistry codes.

2.3.2 The Linear Response Formalism

2.3.2.1 The Linear Response TD-DFT and TDA Approaches

The RG theorem and the time-dependent KS formalism represents the most essential

ingredients to construct the linear response (LR) formalism for TD-DFT, which was first

developed by M.E. Casida in the mid-1990s.92 The LR-TD-DFT is also referred as Casida’s

equation and can be defined by the linear time-dependent response of the system to a time-

dependent external electric field, which describes the change in electron density in response to

a time-varying external potential. Here, we would like to refer to the excellent review articles

by A. Dreuw et al.97 and M. E. Casida et al.89 for extended details. The TD-DFT response

problem is expressed by the following non-Hermitian eigenvalue equation:

( 𝐴 𝐵𝐵∗ 𝐴∗) (

𝑋𝑌) = 𝜔 (

1 00 −1

) (𝑋𝑌) (2.18)

Here, A and B are excitation and de-excitation matrices, respectively, X and Y are eigenvectors

and ω is the eigenvalue. For hybrids, the elements of matrices A and B are defined as:

Aia,jb = δij δab (ϵa - ϵi) + 2(ia | jb) – ax (ij | ab) + (1 - ax) (ia | fxc | jb) (2.19)

Page 34: Theoretical Investigations of the Photophysical Properties

28

Bia,jb = 2(ia | bj) – ax (ib | aj) + (1 - ax) (ia | fxc | bj) (2.20)

The first term in matrix A is the difference of the orbital energies, where δ is the Kronecker

delta, i,j and a,b correspond to occupied and unoccupied orbitals respectively, and ax represents

the amount of non-local Fock exchange (ax = 0 for pure DFT functionals). The last two terms

in matrix A and the elements of matrix B are the two-electron integrals, stem from the linear

response of the Coulomb and XC ( 𝑓xc ) operators. Here 𝑓xc corresponds to second functional

derivative of the XC energy (or XC kernels) and can be written using ALDA:

𝑓xc =𝛿2𝐸𝑥𝑐

𝛿𝜌(r, t) 𝛿𝜌(r´, t´) (2.21)

A popular approximation to the linear response is the Tamm-Dancoff approximation (TDA).98

This approximation corresponds to neglecting the matrix B, thus, instead of solving two

eigenvalue problems as in TD-DFT, only one eigenvalue problem needs to be solved which is

given in equation 2.22:

𝐴𝑋 = 𝜔𝑋 (2.22)

Here, the definition of the matrix elements of A is still the same as in equation 2.19. In the

TDA, the calculated excitation energies are only slightly larger than the full linear response

TD-DFT, but come at lower computational cost (an order of 2).98 Furthermore, the TDA can

be utilized to circumvent the so-called triplet instabilities.99 However, a well-known drawback

of the TDA is that the sum rules are no longer fulfilled, which results in poor description of

calculated oscillator strengths.100 To sum up, the TDA suggests that excitation energies can be

obtained in a good approximation within simplifying scheme of full TD-DFT.

Despite successful applications, TD-DFT on the basis of popular XC functionals and the so-

called adiabatic approximation still suffers accuracy limitations from the electronic SIE which

Page 35: Theoretical Investigations of the Photophysical Properties

29

is most evident in the wrong description of charge-transfer (CT), extended π-systems and

Rydberg states.101-105 Thus, like in DFT, it is necessary to include hybrids which employ a large

fraction of HF exact exchange to alleviate the SIE as well as the CT and related problems.

On the other hand, while TD-DFT can efficiently deal with molecular systems beyond the level

of traditional wave function-based methods, the computation of an entire UV-Vis or circular

dichroism (CD) spectrum of larger molecular or extended biological systems remains a

challenge. To address this challenge, the Grimme group has recently reported two advanced

simplifications to TD-DFT response, the simplified Tamm–Dancoff-Approximation106 (sTDA)

and the simplified Time-Dependent Density Functional Theory approach107 (sTD-DFT). Both

approaches allow extremely fast computations of UV-Vis (and CD) spectra for systems with

up to 1000 atoms. The computational bottleneck of the extremely fast sTDA (or sTD-DFT) is

the determination of the ground state KS orbitals and eigenvalues. In the following sub-section,

the sTDA and sTD-DFT approaches will be described in some more detail.

2.3.2.2 The sTDA and sTD-DFT Approaches

The sTDA approach begins by employing step-by-step three unique simplifications into the

TD-DFT/TDA equation 2.22: (1) The response of the XC functional (last term in equation

2.19) is neglected, to avoid the expensive numerical integration. (2) The two electron integrals

in matrix A are approximated by short-range damped Coulomb interactions of transition density

monopoles. (3) The reduction of the single excitation space. Here, we would like to refer to the

original articles by S. Grimme 106 and C. Bannwarth et al.107 for extended details. To sum up,

the elements of the approximated matrix A (denoted as A′) are then given by in equation 2.23:

A′ ia,jb = δij δab (ϵa - ϵi) +∑ (2𝑞𝑖𝑎𝐴𝑁 𝑎𝑡𝑜𝑚𝑠

𝐴,𝐵 𝛤𝐴𝐵𝐾 𝑞𝑗𝑏

𝐵 − 𝑞𝑖𝑗𝐴𝛤𝐴𝐵

𝐽 𝑞𝑎𝑏𝐵 ) (2.23)

Page 36: Theoretical Investigations of the Photophysical Properties

30

Here, the transition charge density 𝑞 is obtained from a Löwdin population analysis.108 A, B are

defined as atoms. 𝛤𝐴𝐵 with the Mataga-Nishimoto-Ohno-Klopman109-111 damped Coulomb

operator. Note that the pre-factor 𝛼𝑥 is dropped, but its effect is accounted for in the 𝛤𝐴𝐵𝐽

.

The modifications explained above are consistently applied also in the sTD-DFT approach.

The matrix A is just replaced by the approximate matrix A′ from sTDA and the matrix B is set

up in a consistent, simplified manner. As mentioned above, the term originating from the non-

local Fock exchange is of Coulomb type in matrix A, but of exchange-type in matrix B. Since

fitting the another set of parameters avoided, the exchange-type Mataga-Nishimoto-Ohno-

Klopman damped Coulomb interaction has been exploited, while keeping the scaling factor

𝛼𝑥. The elements of approximated matrix B (denoted as B′) can be expressed as:

B′ ia,jb = ∑ (2𝑞𝑖𝑎𝐴𝑁 𝑎𝑡𝑜𝑚𝑠

𝐴,𝐵 𝛤𝐴𝐵𝐾 𝑞𝑏𝑗

𝐵 − 𝛼𝑥𝑞𝑖𝑏𝐴 𝛤𝐴𝐵

𝐾 𝑞𝑎𝑗𝐵 ) (2.24)

In summary, these three simplifications along with the simple eigenvalue problem allow

extremely fast computations for a broad energy range spectrum of very large molecular and

extended biological systems.

2.4 Density Functional Based Tight Binding

While DFT methods have been successfully applied to systems of increasing size and

complexity, methods that can incorporate approximations to reduce further the computational

demand, without compromise much with accuracy, are still required. In this case, the density

functional tight binding method (DFTB) is quite reliable. In this section, we will briefly

introduce the DFTB method and mainly focus on one of its flavors, the self-consistent charge

DFTB (SCC-DFTB) which is thoroughly utilized in this thesis (see Chapters 4-6).

Page 37: Theoretical Investigations of the Photophysical Properties

31

DFTB112-114 is an approximate method based on KS-DFT (see section 2.2) within the Linear

Combination of Atomic Orbitals (LCAO) ansatz. Besides, DFTB is more align towards tight-

binding (TB) scheme and in fact, can be visualized as a non-orthogonal tight-binding scheme

based on DFT parameterization. Since DFTB was first introduced in the 1980s as a non-self-

consistent method112, three major extensions of DFTB have been systematically developed

over the years. These include (1) the self-consistent charge extension,115 known as SCC-DFTB,

(2) the formulation for the spin-dependent calculations116 and (3) the time-dependent treatment

of excited states.117 Generally, TB methods are utilized to calculate the electronic band

structures for solids or periodic systems. However, SCC-DFTB is not restricted to this and can

be utilized for calculating the full electronic structures i.e. total energies. SCC-DFTB

(approximation to KS-DFT) is computationally much faster than DFT and does not demand

multiple empirical parameters. In fact, the parameters are consistently gathered from DFT

treatments of few molecular systems. Since SCC-DFTB (with right parameter set) is

chemically more precise than the force-field methods, it is well utilized for larger systems118-

119 and even at longer time scales than DFT.

The theoretical approximation which is utilized for DFT as a basis for TB method is given by

Foulkes and Haydock,113 where the electronic density 𝜌 is expressed as a reference density 𝜌0,

plus a small fluctuation 𝛿𝜌, and finally expressed in equation 2.25,

𝜌 = 𝜌0 + 𝛿𝜌 (2.25)

This electronic density 𝜌 = [𝜌0 + 𝛿𝜌] is then employed in the KS-DFT total energy (see

equation 2.6) and can be rewritten as:

𝐸𝐾𝑆−𝐷𝐹𝑇[𝜌0 + 𝛿𝜌] = 𝑇𝐾𝑆[𝜌0 + 𝛿𝜌] + 𝐸𝐻[𝜌0 + 𝛿𝜌] + 𝐸𝑥𝑐[𝜌0 + 𝛿𝜌] + 𝐸𝑒𝑥𝑡[𝜌0 + 𝛿𝜌] (2.26)

Page 38: Theoretical Investigations of the Photophysical Properties

32

Afterwards, 𝐸𝑥𝑐[𝜌0 + 𝛿𝜌] is expanded in a Taylor series up to the second-order term, and it is

given in equation 2.27 as follows:

𝐸𝑥𝑐[𝜌0 + 𝛿𝜌] = 𝐸𝑥𝑐[𝜌0] + ∫𝛿𝐸𝑥𝑐

𝛿𝜌 𝛿𝜌 𝑑𝑟 +

1

2 ∬

𝛿2𝐸𝑥𝑐

𝛿𝜌𝛿𝜌′ 𝑉𝑥𝑐 𝛿𝜌𝛿𝜌′𝑑𝑟 𝑑𝑟 ′ (2.27)

followed by substitution of equation 2.27 into 2.26 by using the definition, 𝑉𝑥𝑐 =𝛿𝐸𝑥𝑐

𝛿𝜌, with this

total energy can be written as follows:

𝐸𝑡𝑜𝑡 ≈ 1

2 ∬

𝛿2

𝛿𝜌𝛿𝜌′ 𝑉𝑥𝑐 𝛿𝜌𝛿𝜌′𝑑𝑟 𝑑𝑟 ′ + 1

2∬

𝛿𝜌𝛿𝜌′

|𝑟 −𝑟 ′|𝑑𝑟 𝑑𝑟 ′

−1

2∬

𝜌0𝜌0′

|𝑟 −𝑟 ′|𝑑𝑟 𝑑𝑟 ′ + 𝐸𝑥𝑐[𝜌0] − ∫𝑉𝑥𝑐[𝜌0] 𝜌0 𝑑𝑟 + 𝑉𝑛𝑛 (2.28a)

In SCC-DFTB,115 the total energy expression, 𝐸𝑡𝑜𝑡𝑆𝐶𝐶−𝐷𝐹𝑇𝐵 is derived from the second order-

expansion of KS-DFT energy, and can be expressed in equation 2.28b:

𝐸𝑡𝑜𝑡𝑆𝐶𝐶−𝐷𝐹𝑇𝐵 = 𝐸𝑒𝑙𝑒𝑐 + 𝐸𝑟𝑒𝑝 = 𝐸𝐵𝑆 + 𝐸𝑆𝐶𝐶 + 𝐸𝑟𝑒𝑝 (2.28b)

where, 𝐸𝑒𝑙𝑒𝑐 denotes the electronic and 𝐸𝑟𝑒𝑝 the repulsive energy term. The 𝐸𝑒𝑙𝑒𝑐 consists of

the zero order (𝐸𝐵𝑆) and the second order (𝐸𝑆𝐶𝐶) terms. The zero-order term, also known as

the band-structure (BS) term, is the summation of energy over all the occupied eigenstates

represented by LCAO and is given in equation 2.29 as follows:

𝐸𝐵𝑆 = ∑ [𝑛𝑖 휀𝑖]𝑣𝑎𝑙𝑒𝑛𝑐𝑒 𝑒𝑖 = ∑ 𝑛𝑖⟨𝜓𝑖|𝐻0|𝜓𝑖⟩

𝑣𝑎𝑙𝑒𝑛𝑐𝑒 𝑒𝑖 (2.29)

In equation 2.29, 𝑛𝑖 denotes occupation, and, 휀𝑖 is the orbital energy for each of the occupied

electronic eigenstates signified by 𝜓 and the basis-set used for these calculations is minimal.

The diatomic distance related Hamiltonian matrix elements are calculated by DFT and stored

in SCC-DFTB parameter files, called a Slater-Koster tables. Besides 𝐸𝐵𝑆, the second term, i.e.

=𝐸𝑒𝑙𝑒𝑐

=𝐸𝑟𝑒𝑝

Page 39: Theoretical Investigations of the Photophysical Properties

33

𝐸𝑆𝐶𝐶 which is derived from the charge density fluctuations is included in total energy

expression. The 𝐸𝑆𝐶𝐶 term is calculated via expansion of the energy corresponding to Coulomb

interaction (taking charge fluctuation into account) is given by,

𝐸𝑆𝐶𝐶 = 1

2∑ [𝛾𝐴𝐵∆𝑞𝐴∆𝑞𝐵]𝑎𝑡𝑜𝑚𝑠

𝐴𝐵 (2.30)

where, 𝛾𝐴𝐵 is the distance-dependent, mainly a Coulomb charge-density interaction functional.

The variables ∆𝑞𝐴 and ∆𝑞𝐵 are induced Coulomb charges on atom A and B respectively,

derived from Mulliken population analysis.120 Here we would like to refer the original article

on SCC-DFTB by Elstner et al.115 for extended details.

To sum up, by utilizing approximations and proper parameterization, SCC-DFTB is well-

qualified method and expected to keep the accuracy of full DFT while computationally faster

and efficient. In this thesis, we have thoroughly used SCC-DFTB for the electronic structure

calculations of extended system of interest. (more details can be found in Chapter 4, 5, and 6)

2.5 Periodic Treatment of Crystalline Materials

To study the electronic structure properties of crystalline materials based on the periodic

arrangement of the atoms, we first must introduce several definitions related to crystal

structure. The fundamental property of a crystal or crystalline material is an ordered repetitive

arrangement of its constituents (atoms, molecules, or ions) forming a crystal lattice that extends

in all the directions. The terms “periodic” and “order” are not equivalent, and most recent

definition by the “International Union of Crystallography” thus states: A material is a crystal

if it has essentially a sharp diffraction pattern.121 This means periodic crystal forms subset and

usually mapped onto itself by a certain translation. Therefore, to describe a system which shows

a well-defined structure such as crystals or surface-mounted MOFs (SURMOFs) periodic

boundary conditions (PBCs) can be studied.

Page 40: Theoretical Investigations of the Photophysical Properties

34

In PBCs, periodicity is treated by a set of boundary conditions which are often selected for

approximating a large (infinite) system by using a small repeating unit that reflects the

symmetry of the system, called as unit cell. The periodicity can be 1D (e.g. polymer), 2D (e.g.

a surface) or 3D (e.g. a crystal), with the latter being most common. Every periodic system has

two lattices associated with it. The first one, is known as the real space lattice in which any

periodic distribution of a set of objects can be characterized by a certain translation that repeats

the set of objects periodically. This set of translations generate real space lattice and is defined

by a vector �� ,

�� = 𝑛1𝑎1 + 𝑛2𝑎2 + 𝑛3𝑎3 (2.31)

Here, the set of points defined by �� is called Bravais lattice (as shown in Figure 2.4a) where

𝑛1, 𝑛2, and 𝑛3 are integers and the vectors 𝑎1 , 𝑎2 , and 𝑎3 are the basis vectors in 3D space.

These three vectors are not necessarily orthogonal, and they restrict to angles 𝛼, 𝛽, and 𝛾. The

length of unit cell can be given by 𝑎, 𝑏, and 𝑐 along the axes of three vectors. All the physical

dimension of unit cells such as 𝑎1 , 𝑎2 , 𝑎3 , 𝑎, 𝑏, 𝑐, 𝛼, 𝛽, and 𝛾 in a crystal lattice are collectively

referred as lattice parameters or even unit cell parameters (as shown in Figure 2.4b).

Figure 2.4 (a) A 3D Bravais lattice. The choice of the primitive vectors a1, a2, and a3 is not

unique. (b) represents the lattice parameters includes lattice vectors, lengths, and angles

Page 41: Theoretical Investigations of the Photophysical Properties

35

The other lattice is the reciprocal space lattice (also known as k-space),122 which determines

how the periodic structure interacts with waves and plays a fundamental role in most analytic

studies of periodic structures, particularly mapping the diffraction pattern. The reciprocal space

lattice vector, 𝐺 is defined as a linear sum of new set of basis vectors 𝑏1 , 𝑏2

, and 𝑏3 derived

from the 𝑎1 , 𝑎2 , and 𝑎3 vectors of the real cell, and obeying the orthonormality condition

𝑎𝑖𝑏𝑖 = 2𝜋 𝛿𝑖𝑗. The reciprocal lattice vector, 𝐺 then can be expressed as follows:

𝐺 = 𝑚1𝑏1 + 𝑚2𝑏2

+ 𝑚3𝑏3 (2.32)

where 𝑚1, 𝑚2, and 𝑚3 are integers, and �� is real space lattice vector. The equivalent of a unit

cell in reciprocal space is called the (first) Brillouin zone (Figure 2.5a). The point in a reciprocal

space is described by a vector k. Since k has units of inverse length, it is called a wave vector.

The periodicity of the particle in the system means that the square of the wave function must

exhibit the similar periodicity. This is inherent by the Bloch theorem,123 which states that for a

periodic potential with translational symmetry, 𝑉(𝑟 ) = 𝑉(𝑟 + �� ), the corresponding time-

independent SE for a particle yields a set of eigen-functions,

Figure 2.5 (a) Brillouin zone for the face-centered cubic. (b) An illustration of the electronic band

structure diagram for a semiconducting material, Gallium phosphide (GaP)

Page 42: Theoretical Investigations of the Photophysical Properties

36

𝛹𝑛�� (𝑟 ) = 𝑢𝑛�� (𝑟 ) 𝑒𝑖�� .�� (2.33)

where 𝑢𝑛�� (𝑟 ) is a periodic function with same periodicity as the potential, 𝑢𝑛��

(𝑟 ) = 𝑢𝑛�� (𝑟 +

�� ), the solution of the SE are characterized by an integer number n (called band index) and the

wave vector �� . The eigenvalues 휀𝑛�� are dependent on the wave vector �� , this dependency is

the basis for the band structure model and forms a continuous spectrum, so called band structure

diagram (see Figure 2.5b). Like in non-periodic system (or molecular system), these bands are

filled with electrons according to energy. The “top” highest occupied filled band is called

valence band (VB) and the “bottom” lowest unoccupied one which is called as conduction band

(CB). The energy difference between the top of the VB and the bottom of CB is called as band

gap and is equivalent to HOMO-LUMO gap in non-periodic system. The band gap values are

generally zero (metallic systems) and finite (insulators or semi-conductors) depending on

whether the band gap is large or small compared with the thermal energy kBT.

The application of Bloch's theorem to the Kohn-Sham wavefunctions greatly facilitates the

treatment of periodic crystals by exploiting their lattice periodicity. Due to this tremendous

computational simplification, DFT has become a working horse for the calculations of

structures and properties of crystalline framework materials. Unfortunately, there is no DFT

software available that concomitantly offers all the required features for above-mentioned class

of materials. In this thesis, we have employed DFT for electronic structure and band structure

calculations for PP-SURMOFs. For this 3D bulk system of interest, we have discovered that

the DFT (GGA-level itself) has not only suffered computationally but also encountered

troubles in their convergence. Furthermore, the well-known shortcomings for electronic

structure calculations using DFT, most notably the too-narrow band gaps were observed. To

overcome these shortcomings, we have employed the DFTB (an approximate-DFT) approach

which keeps the accuracy of the full-DFT while computationally faster and efficient. However,

Page 43: Theoretical Investigations of the Photophysical Properties

37

DFT has been proven to efficiently calculate the electronic band structure properties for

SURMOFs at most prominently the GGA-level of approximations, while the band gaps were

improved at hybrid-level.

2.6 Applied Computational Chemistry Packages

In this section, we briefly introduce the computational chemistry packages used to investigate

the electronic structure properties of the molecular and extended systems of interest.

AMS is the Amsterdam Modeling Suite that contains various set of modules for

computationally tackling research problems of various types in diverse areas of chemistry and

materials science. Below we list the AMS modules that have been used in our studies:

ADF, the Amsterdam Density Functional,124 has been used for molecular systems,

particularly in predicting electronic structure and spectroscopy.

AuToGraFS, the Automatic Topological Generator for Framework Structures,125 has

been used for stitching the periodic framework structures from topological information,

and specification of individual linkers and connectors. This coupling framework

generation to a force field is quite essential to produce starting structures that should be

followed by higher-level methods. (see Chapters 4, 6)

UFF, the universal force field,126 is an all atom potential containing parameters for

every atom. It has been used to provide parameters for each atomic type within a

framework to dictate the strength of local chemical bonds / bends. (see Chapters 4, 6)

DFTB, the Density Functional Tight Binding method.127 It is available in three variants,

the DFTB0, DFTB2, and DFTB3. The self-consistent charge (SCC) approximation,

also known as DFTB2, improves the transferability, in particular to polar systems and

Page 44: Theoretical Investigations of the Photophysical Properties

38

has been thoroughly used for determining the (3D) bulk structure, stacking, electronic,

including topological, properties in a first screening approach. (see Chapters 4,5 and 6)

TURBOMOLE is an ab initio computational chemistry program.128 It has been used for

ground-state electronic structure calculations for molecular systems of interest (see Chapter 3).

ORCA is a general-purpose quantum chemistry program package.129 It is a molecular code

lacking periodic boundary conditions. It features all modern electronic structure methods with

specific emphasis on spectroscopic properties. It has been thoroughly used for the calculations

of ground-state as well as excited-state electronic structure properties for molecular system of

interest (see Chapters 3, 4, 5 and 6).

CRYSTAL is a general-purpose program for the study of crystalline solids.130 It computes the

electronic structure properties of 0D, 1D, 2D and 3D systems within wave function and density-

based approximations. It has been used for determining the (3D) bulk electronic band gap and

band structure properties within DFT-level (see Chapters 4, 5 and 6).

Page 45: Theoretical Investigations of the Photophysical Properties

39

Chapter 3

Benchmarking the Performance of Time-Dependent

Density Functional Theory for Predicting the UV-Vis

Spectral Properties of Porphyrinoids

"Because the theory of quantum mechanics could explain all of chemistry

and the various properties of substances, it was a tremendous success.

But still there was the problem of the interaction of light and matter." – Richard P. Feynman

The studies summarized in this Chapter have been published as:

Benchmark of Simplified Time‐Dependent Density Functional Theory for UV–Vis Spectral

Properties of Porphyrinoids

Kamal Batra, Stefan Zahn, and Thomas Heine

Adv. Theory Simul. 2020, 3, 1900192 © Wiley-VCH Verlag GmbH & Co. KGaA

Page 46: Theoretical Investigations of the Photophysical Properties

40

This Chapter investigates the capability of various variants of time-dependent density-

functional theory (TD-DFT) for predicting the UV-Vis spectra (including Soret- and Q-bands)

of porphyrinoids with the aim to identify a computationally feasible approach for large-scale

applications such as molecular framework materials. In the following, we mainly focus on the

characteristic Q-bands because they absorb in the visible light range.

In Section 3.1, after a brief introduction of porphyrinoids followed by the shortcomings of the

ab-initio approaches, we introduce the canonical TD-DFT and the popular Tamm-Dancoff

Approximation (TDA) for predicting the Q-bands of porphyrinoids. Then we summarize the

TD-DFT shortcomings and the state-of-the-art for porphyrinoids. Finally, we present two

highly efficient approaches developed by Grimme et al., namely, the simplified TDA (sTDA)

and the simplified TD-DFT (sTD-DFT).

Section 3.2 lays out a summary of computational methodologies which include various TD-

DFT approaches, basis-sets, density functional flavors (gradient corrected, global hybrids,

range-separated, and double hybrids), and some ab-initio approaches for the comparison.

Besides, we briefly present the benchmark-set of porphyrinoids having a diverse extent of π-

conjugation, ring functionalization, as well as inclusion/modification of a central metal atom.

In Section 3.3, we discuss the performance of the canonical and the simplified version of TD-

DFT approaches by exploiting various basis-sets and density functional flavors with respect to

the experimental references. Furthermore, we discuss some traditional ab-initio e.g. CIS and

CIS(D) approaches with comparable computational cost to density functional flavors.

Finally, in Section 3.4, we discuss the range-separated functional CAM-B3LYP in combination

with the sTDA approach and a basis-set of double-ζ quality, def2-SVP, as an excellent choice

for these calculations. This approach outperforms more expensive approaches, even double

hybrids, and incrementally improves the results by systematic excitation energy scaling.

Page 47: Theoretical Investigations of the Photophysical Properties

41

3.1 Introduction

Porphyrins (PPs) and their derivatives can be found in many natural biological systems and

offer potential solutions to a wide range of applications. In plants, PPs are an essential part of

the chlorophyll pigment that converts solar energy into chemical energy.131 Porphyrinoids also

have been proven to be efficient sensitizers132 and catalysts133 in several chemical processes,

including medical applications such as photodynamic therapy.134 They have been incorporated

both as linkers and connectors in metal-organic frameworks (MOFs)72, 135 and covalent-organic

frameworks (COFs).136 A good light-harvesting material efficiently absorbs photons from the

highly abundant visible solar spectrum. This property can be probed by UV-Vis spectroscopy.

The characteristic absorption bands of porphyrinoids are displayed for an example of

tetraphenyl PP, see supporting information (SI) Figure S3.1 in the appendix B.1. The intense

Soret-band, also called B-band, commonly arises in the UV from 350 to 450 nm. In metal free

porphyrins, four transitions with much lower intensity are found in the spectral range from 450

to 800 nm, which are called Q-bands. All transitions between the frontier orbitals are allowed

based on symmetry rules. However, both highest occupied molecular orbitals (HOMO-1 and

HOMO) as well as both lowest unoccupied molecular orbitals (LUMO and LUMO+1) are close

in energy and thus, nearly degenerated as in a simplified model of a 18π cyclic polyene, as

employed by Gouterman13. In this model, two transitions are allowed between the degenerated

frontier orbitals while two are forbidden. Indeed, the frontier molecular orbitals of porphyrin

and the model system show strong similarities. Furthermore, a strong mixing of the transitions

was observed for Soret and Q-bands by quantum chemical calculations.14-15 Opposing

transition dipoles reduce the intensity of the Q-bands while a parallel orientation of both

transition dipoles contributes to the Soret band and thus, a more intense absorption band is

observed for the latter. Therefore, tuning the energy levels of the frontier molecular orbitals

strongly affects the absorption intensity of the characteristic Q-bands. The higher the energy

Page 48: Theoretical Investigations of the Photophysical Properties

42

gap between HOMO-1 and HOMO as well as LUMO and LUMO+1, the stronger will be the

absorption intensity of the Q-bands. Nonetheless, predicting the final spectra is challenging.

Obviously, correlated ab-initio approaches such as coupled cluster theory137-139 (CC2, CCSD,

and CC3), the algebraic diagrammatic construction through second order140 (ADC2), and

complete active space second order perturbation theory139, 141-142 (CASPT2) deliver reliable

absorption energies in accordance with experimental results. However, these approaches are

computationally quite expensive for tackling a conjugated molecular system beyond the basic

PP. Other prominent approaches, such as symmetry adapted cluster-configuration

interaction143 (SAC-CI) and similarity transformed equation-of-motion coupled-cluster144

(STEOM-CC) are more accurate than CASPT2, but limited to molecular system up to 50 atoms

due to their high computational cost. To overcome the limits of CASPT2, second order N-

electron valence state perturbation theory145-146 (NEVPT2) has been introduced, which is more

efficient than CASPT2, size consistent and intruder-state-free, but like all multi-reference

approaches, the computational cost of NEVPT2 is still high for larger systems. Overall,

previously mentioned approaches yield reliable absorption energies for PPs, but suffer from a

high computational cost for increased molecular systems, making them practically unsuitable

for large systems.

To find an alternative for the prediction of excited state properties of large systems at a

moderate cost, time-dependent density functional theory (TD-DFT) appears to be a promising

candidate. TD-DFT is an extension of Kohn-Sham DFT and based on almost 35-years old

Runge-Gross theorem93 which has been thoroughly reviewed in the literature89, 92-93, 147-151.

Almost 25 years ago, Casida developed a constructive linear-response formalism for TD-DFT,

known as Casida equations92 but which we will refer to as the random-phase approximations

Page 49: Theoretical Investigations of the Photophysical Properties

43

(RPA), allowing to efficiently determine the solution of the TD-DFT equations, which are

formulated in matrix equation involving the excitation and de-excitation matrices.

A popular approximation to the Casida equations is the Tamm-Dancoff approximation98

(TDA), which simplifies the algebra and associated algorithms to obtain the electronic

excitations, yet it typically yields electronic excitations close to those obtained by TD-DFT.152-

153 Unfortunately, TD-DFT on the basis of popular exchange correlation functionals suffers

accuracy limitations which are most evident in the failure to correctly describe Rydberg and

charge transfer (CT) states.103, 154 These drawbacks can usually be overcome by range-

separated hybrid (RSH) functionals, which employ a large amount of HF exchange at large

electron-electron distances and, therefore, reflect the correct asymptotic exchange potential.

TD-DFT in numerous variants has been applied to PPs, and in the following we are

summarizing the state of the art as found in the current literature:

In 1996, Bauernschmitt et al.155 employed TD-DFT to compute the first four electronic

excitations of PP to validate exchange correlation functionals, including the local density

approximation (LDA: S-VWN), the generalized gradient approximation (GGA: BP86) and

hybrid functional (B3LYP). Their results showed that TD-DFT excitation using the BP86

functional are in better accordance with experiment than CIS and TD-HF. Also, CASPT2

possesses an error of more than 0.3 eV for the Q-bands compared to experimental results.

However, the employed basis set was overall small for post-HF approaches.

In 2010, Tian et al.156 examined the performance of global hybrids (PBE0, B3LYP, M06, M06-

2X, M06HF) and long-range corrected (LC) hybrid functionals (ωB97X-D, ωB97X, ωB97,

LC-ωPBE and CAM-B3LYP) in TD-DFT calculations to predict the spectral properties of PP

analogues. Among the many functionals tested, the LC functional ωB97X-D results in an error

of 0.05 eV for Qy band. Moreover, they concluded that the results are robust with respect to

Page 50: Theoretical Investigations of the Photophysical Properties

44

subtle geometry changes resulting from the functional choice for geometry optimization and

showed that diffuse functions have only a minor effect on calculated absorption spectra.

However, the quite general study included only two porphyrinoids.

Eriksson et al.157 (2011) investigated the ability of LC hybrid functionals ωB97, ωB97X and

ωB97X-D within the TD-DFT framework. They found that ωB97X reproduces the experiment

best with an error of up to 0.09 eV. Additionally, it was confirmed that the applied functional

for geometry optimization has only a small influence on the calculated spectra.

Lee et al.158 (2012) benchmarked five DFT functionals (B3LYP, LC-ωPBE, LC-BLYP, CAM-

B3LYP and ωB97X-D) using TD-DFT for PP derivative. It was found that ωB97X-D yields

the best agreement to the reference for the LC functionals (Qave bands: 0.055 eV). Overall,

better results were obtained for B3LYP for the Soret and Q-bands. However, it was not

recommended due to the susceptibility for charge transfer excitations.

A benchmark set of 66 medium-sized and large aromatic organic molecules, including five

porphyrinoids, has been studied by Winter et al.159 in 2013. B3LYP was outperformed by the

investigated post-HF approaches (ADC (2), CC2, SOS-CC2, SCS-CC2).

Fang et al.160 (2014) compiled a subset of 96 excitations of 79 different organic and inorganic

molecules, including basic PP.161 They have assessed diverse DFT functionals (BP86, B3LYP,

PBE0, M06-2X, M06-HF, CAM-B3LYP and ωB97XD) and two wave-function based

approaches (CIS and CC2). Overall, the lowest error was produced by CC2 with MAE of 0.19

eV. However, it was found that CC2 approach did not perform well for inorganic systems

(MAE: 0.31 eV) while the MAE of B3LYP is only 0.22 eV.

Theisen et al.162 (2015) validated the performance of diverse DFT functionals (B3LYP, PBE0,

CAM-B3LYP, M062X, M06, M11) for Zn-phthalocyanine (ZnPC). Interestingly, the extra

Page 51: Theoretical Investigations of the Photophysical Properties

45

diffuse function in 6-31+g(d) caused convergence problems in the TD-DFT calculations.

Among the investigated functionals, M11 showed the best accordance with experiment with an

error of 0.13 eV for the Qx (0-0) band.

Despite the many successful applications of TD-DFT on a wide range of molecular systems, it

is often challenging in TD-DFT to calculate a sufficient number of excited states for a complete

spectrum or spectra of extended biological systems. To overcome this challenge, the Grimme

group presented two highly efficient approaches, the simplified Tamm–Dancoff-

Approximation106 (sTDA) and simplified Time-Dependent Density Functional Theory

approach107 (sTD-DFT). In both approaches, the computational resources needed to tackle a

targeted system is solely determined by the ground state DFT calculation. This is achieved by

approximating Coulomb and exchange interactions of the electrons by monopole interactions.

Additionally, the CI space is truncated with a screening based on second-order perturbation

theory. Note, the central concepts of the sTDA approach to increase the computational

efficiency can be also employed in tight-binding approaches and, thus, allows fast access to

excited state properties of systems with an amazing size.163 Computational studies validating

sTDA or sTD-DFT for PPs are missing in the literature so far.

With the goal to identify a computational feasible approach to investigate the absorption

properties of PP-containing materials and extended biological systems, we compare the semi-

empirical sTDA, sTD-DFT and canonical TD-DFT (RPA and TDA) for UV-Vis spectra

calculations of porphyrinoids. After a short summary of computational details, we assess

diverse DFT functionals regarding their performance for the calculations of absorption energies

of the Q- bands. This includes the (for sTDA unsuccessful and unnecessary) attempt to improve

the results by an empirical scaling of excitation energies can improve results significantly.

3.2 Computational Methodologies

Page 52: Theoretical Investigations of the Photophysical Properties

46

All geometries have been fully optimized using the Turbomole-suite128, employing the BLYP

functional with Grimme’s D3 correction for London dispersion (BLYP-D3)164-166 in

combination with the resolution of identity (RI) approximation,167-169 and the TZVP170 split-

valence basis set of triple-ζ quality with polarization functions. The convergence criterion for

the self-consistent field approach was increased to 10−8 Hartree. This approach is very fast for

the molecules studied here, but also easily affordable in periodic calculations that suffer severe

performance loss for hybrid functionals. As hybrid functionals are known to produce more

accurate structures, we assessed the influence of the geometry on the excited state properties.

For that purpose, we reoptimized all geometries using the B3LYP-D3 hybrid functional164-166,

171 , again employing the TZVP basis set and the RI.

For excited state properties we applied the ORCA code129 with a wide range of functionals,

basis sets and TD-DFT approaches. A summary of the calculation types is given in Table S3.1

in the SI. In detail, we calculated UV-Vis spectra using the following density-functionals:

GGA and mGGA functionals: BLYP,164-165 BP86,164, 172 PBE,173-174 TPSS,175 M06-L176

Global hybrid functionals: B3LYP,164-165, 171 PBE0,173-174, 177 B3P86,171-172 BHLYP,178

TPSS0,179 M06,180 M06-2X,180

Range separated functionals: ωB97,181 ωB97X,181 LC-BLYP,182 CAM-B3LYP183

Double hybrid functionals: B2PLYP,184 B2GP-PLYP,185 mPW2PLYP186

The motivation of this work is to quantitatively reproduce the Q-bands of porphyrinoids with

the possibly lowest computational cost which plays a vital role for the simulations in bio- and

material related chemistry. Hence, each method is validated with the relatively small basis set

def2-SVP187 which is of double-ζ quality. For basis set validation, we repeated the calculations

with the def2-TZVP basis set187 for the sTDA and sTD-DFT approaches, and we also

investigated the impact of diffusive functions for sTDA (def2-SVPD188 and def2-TZVPD).188

Page 53: Theoretical Investigations of the Photophysical Properties

47

To speed up the calculations, we employed the RI approximation throughout, including its

variant for double hybrid functionals,189-190 and the RIJCOSX approximation191 was employed

for the global hybrid and range-separated hybrid functionals. For comparison, we have also

included the post-Hartree-Fock methods CIS192 and CIS(D).193-194

The performance of each approach was assessed by calculation of the mean error (ME), the

mean absolute error (MAE), and the absolute maximum error (MAXE) to the experimental

reference values.

Benchmark Set: We have included diverse variants of porphyrinoids starting from basic PP

to the extension of conjugated π-system of the central core followed by ring functionalization

and modification of metal atoms. The molecules included in our benchmark set are given in

Figure 3.1 while Table 3.1 list the experimental references.

Table 3.1 Experimental references of benchmark-set of investigated porphyrinoids

Porphyrinoids Benchmark-Set Abbrev. Ref.

[1] Porphyrin H2PP [195-197]a

[2] Octaethylporphyrin H2OEP [196-198]b

[3] Magnesium Octaethylporphyrin MgOEP [196-197, 199]a

[4] Zinc Octaethylporphyrin ZnOEP [196-197, 200]a

[5] Tetraphenylporphyrin H2TPP [196-197, 201]a

[6] Magnesium Tetraphenylporphyrin MgTPP [196-197, 202]a

[7] Zinc Tetraphenylporphyrin ZnTPP [196-197, 201]a

[8] Tetrakis(o-aminophenyl) porphyrin H2TAPP [196-197, 203]a

[9] Zinc tetrakis(4-carboxyphenyl) porphyrin ZnTCPP [204]c

[10] Zinc [5,15-dipyridyl-10,20-bis(pentafluorophenyl) porphyrin F-ZnP [205]d

[11] Zinc [5,15-di(4-pyridylacetyl)-10,20-diphenyl] porphyrin DA-ZnP [205]d

[12] Octabromotetraphenyl porphyrin H2OBP [206]e

Toluenea ; Benzeneb ; THFc ; DMFd ; CH2Cl2e

Page 54: Theoretical Investigations of the Photophysical Properties

48

3.3 Results and Discussion

In this section, the performance of diverse functionals is presented. We will start the validation

with the computationally least costly DFT approach, the GGA functionals, and will finish with

the most expensive one, the double hybrid functionals. We would like to add here that the

present TD-DFT studies involve only transitions in the frozen ground state of the molecules.

Therefore, only the 0 → 0 transitions of the Q-bands can be obtained from the calculations. We

Figure 3.1 Molecular structures of porphyrinoids included in our benchmark set.

Page 55: Theoretical Investigations of the Photophysical Properties

49

will focus on the Q-bands in the following because they absorb in the visible light range.

Furthermore, these transitions can be clearly distinguished from other transitions in the excited

state calculations.

3.3.1 GGA and meta-GGA Functionals

GGA and meta-GGA functionals do not require four-center integrals as the Coulomb

interaction can be calculated directly via the electron density, which is particularly beneficial

for periodic calculations and for codes employing different basis functions than Gaussian-type

orbitals. On the other hand, pure Kohn-Sham DFT is very prone to errors originating from the

self-interaction error. Thus, an overall poor general performance for excited state calculations

can be expected due to weakly bound electrons.

As can be seen in Table 3.2 and Figure 3.2a, the performance of TD-DFT for the GGA

functionals PBE, BP86 and BLYP is very similar. Compound 8 shows a large error resulting

in an outlier of about 0.5 eV. This can be attributed to a significant contribution of charge

transfer excitations to the Q-bands (see SI, Table S3.2 and Figure S3.2a, S3.2b for a detailed

description). Employing the meta-GGA functionals TPSS and M06-L reduces significantly the

MAXE. However, the MAE (denoted by ‘+’ sign in Figure 3.2a) is not improved and still

exceeds 0.12 eV. Furthermore, meta-GGAs tend stronger to over-estimate the absorption

energies compared to GGAs. Employing the RPA-approach does not result in significant

improvements in comparison to the TDA-approach (see Table 3.2 and Figure 3.2b). For

instance, the calculated MAE from both approaches in combination with the GGA-functionals

is nearly similar. Only for M06-L, the MAE is reduced by about 0.03 eV. Additionally, the

RPA-approach tends to lower absorption energies than the TDA-approach.

Table 3.2 Calculated original error values in eV for the GGA and meta-GGA functionals

Page 56: Theoretical Investigations of the Photophysical Properties

50

GGA and meta-

GGA

Functionals

TDA

(def2-

SVP)

RPA

(def2-

SVP)

sTDA

(def2-

SVP)

sTD-DFT

(def2-

SVP)

sTDA

(def2-

TZVP)

sTD-DFT

(def2-

TZVP)

PBE

ME 0.02 -0.04 -0.08 -0.12 -0.18 -0.20

MAE 0.11 0.11 0.09 0.13 0.18 0.21

MAXE 0.48 0.48 0.52 0.53 0.49 0.50

BP86

ME 0.02 -0.04 -0.08 -0.12 -0.18 -0.20

MAE 0.11 0.11 0.09 0.13 0.18 0.21

MAXE 0.47 0.47 0.52 0.52 0.48 0.49

BLYP

ME 0.01 -0.05 -0.08 -0.13 -0.18 -0.21

MAE 0.10 0.10 0.09 0.14 0.19 0.21

MAXE 0.46 0.46 0.49 0.50 0.47 0.48

TPSS

ME 0.06 0.00 -0.02 -0.06 -0.12 -0.15

MAE 0.12 0.10 0.08 0.09 0.13 0.16

MAXE 0.38 0.38 0.41 0.42 0.39 0.41

M06-L

ME 0.11 0.05 0.05 0.01 -0.05 -0.08

MAE 0.14 0.11 0.10 0.08 0.09 0.11

MAXE 0.25 0.26 0.28 0.30 0.29 0.31

The computationally cheaper approaches, sTDA and sTD-DFT, show a comparable MAXE for

GGAs (see Table 3.2 and Figure 3.3a, 3.3b). However, the MAE of the GGA-functionals is

strongly affected by the selected approach and basis set. For example, the ME and MAE of

sTDA with the functional BLYP increases by about 0.1 eV when the larger def2-TZVP basis

set is employed (see Table 3.2 and Figure 3.4a). Changing from sTDA to sTD-DFT results also

in large MAE values. The ME and MAE of a given approach-basis-set-combination is overall

close to each other in all cases and thus, highlighting systematic deviations. Please note, that

the sTDA approach includes empirical parameters optimized for functionals with a Hartree-

Fock exchange contribution between 20% and 60% and not for pure DFT functionals.106 TPSS

Page 57: Theoretical Investigations of the Photophysical Properties

51

shows an improvement to the GGAs but is still significantly affected by the selected basis-set.

The MAE of M06-L for sTDA and sTD-DFT is only slightly affected by the choice of basis

set which is in contrast to the GGAs. To sum up, the MAE of GGAs and meta-GGAs exceeds

0.08 eV while the MAXE is reduced to 0.30 eV only for the M06-L functional. Finally, a

systematic under-estimation of absorption energies, especially for the sTDA and sTD-DFT in

combination with a large basis set, is observed suggesting a global scaling of the obtained

energies to match better the experimental reference. This approach is rather semi empirical but

allows to access excited state calculations of extended system sizes due to the low cost of pure

Kohn-Sham density functional theory.

After scaling of energies, significant improvements of the MAE and MAXE are only obtained

for the sTDA and sTD-DFT approaches in combination with the large basis-set, (see Figure

3.4b for e.g. BLYP). All functionals with scaled error value are listed in Table S3.3 and a

graphical illustration can be seen in Figure S3.5. Nonetheless, the MAE is still above 0.05 eV

and high MAXE is obtained as well. Thus, GGAs and meta-GGAs cannot be recommended for

calculations of UV-Vis-spectra of porphyrinoids.

3.3.2 Hybrid Functionals

A hybrid functional is defined as an approximate KS density functional where a part or all the

semi-local DFT exchange expression 𝐸𝑋𝐷𝐹𝑇 is replaced by exact Hartree-Fock (HF) exchange

𝐸𝑋𝐻𝐹. The amount of HF exchange for typical hybrid functionals lies in the range of 10-25%

but can be as high as 50-55% like in the BHLYP and the M06-2X functional. Increasing the

amount of HF exchange also increases the likelihood of encountering triplet instabilities (i.e.,

imaginary triplet excitation energies which indicate that the ground state is unstable with

respect to symmetry breaking. Moreover, Incorporation of exact HF exchange reduce the self-

interaction error (SIE) which is a significant troublemaker in TD-DFT. However, it also

Page 58: Theoretical Investigations of the Photophysical Properties

52

introduces a four-index integral into the Hamiltonian which leads to higher computational cost

in comparison to the GGA and meta-GGA functionals.

As shown in Table 3.2 and 3.3, similar trends are observed for global hybrids as for pure DFT

functionals, e.g. RPA tends to lower absorption energies than TDA. In contrast, the large error

which stem from the outliers are significantly reduced for the global hybrids with respect to

GGAs (see Table 3.3 and Figure 3.2a, 3.2b). TDA in combination with global hybrids tends

strongly to over-estimate absorption energies of Q-bands which can be reduced by employing

the RPA approach, especially in combination with the BHLYP and M06-2X functional.

Nonetheless, MAE is still above 0.10 eV and thus, this approach cannot be recommended.

Table 3.3 Calculated original error values in eV for the global hybrid functionals

Global Hybrid

Functionals

TDA

(def2-

SVP)

RPA

(def2-

SVP)

sTDA

(def2-

SVP)

sTD-DFT

(def2-

SVP)

sTDA

(def2-

TZVP)

sTD-DFT

(def2-

TZVP)

PBE0

ME 0.22 0.15 -0.09 -0.12 -0.22 -0.24

MAE 0.22 0.15 0.10 0.12 0.22 0.24

MAXE 0.31 0.25 0.23 0.27 0.40 0.42

B3P86

ME 0.19 0.13 -0.02 -0.06 -0.15 -0.17

MAE 0.19 0.13 0.06 0.08 0.16 0.18

MAXE 0.28 0.24 0.18 0.22 0.34 0.35

B3LYP

ME 0.18 0.12 -0.03 -0.07 -0.16 -0.18

MAE 0.18 0.12 0.06 0.08 0.17 0.19

MAXE 0.27 0.23 0.18 0.22 0.35 0.36

TPSS0

ME 0.24 0.16 -0.04 -0.08 -0.17 -0.20

MAE 0.24 0.16 0.07 0.09 0.18 0.20

MAXE 0.32 0.27 0.18 0.24 0.35 0.37

M06 ME 0.15 0.07 -0.16 -0.20 -0.31 -0.33

Page 59: Theoretical Investigations of the Photophysical Properties

53

MAE 0.15 0.10 0.16 0.20 0.31 0.33

MAXE 0.24 0.17 0.30 0.34 0.48 0.51

BHLYP ME 0.28 0.14 -0.30 -0.36 -0.44 -0.53

MAE 0.28 0.14 0.30 0.36 0.44 0.53

MAXE 0.36 0.22 0.42 0.47 0.61 0.69

M06-2X ME 0.28 0.15 -0.40 -0.49 -0.54 -0.66

MAE 0.28 0.15 0.40 0.49 0.54 0.66

MAXE 0.35 0.23 0.54 0.61 0.73 0.83

The approximate sTDA and sTD-DFT approaches fail significantly for the global hybrids with

large amount of HF exchange like BHLYP and M06-2X. However, global hybrids with HF-

exchange contributions in the range of 20-25% show significant improvement compared to the

pure DFT functionals (see Table 3.3 and Figure 3.3a, 3.3b). For example, sTDA in combination

with the B3LYP functional and def2-SVP basis set possesses a MAE of 0.06 eV while the

MAXE is 0.18 eV. Increasing the basis set increases the deviation, similarly as for the GGA

functionals (see Figure 3.4a for e.g. B3LYP and Figure S3.5 for all the tested functionals).

The MAE and MAXE can be reduced by energy scaling due to systematic deviations. A global

hybrid with large HF-exchange contribution works best: sTDA/sTD-DFT in combination with

the functionals BHLYP or M06-2X and the def2-SVP basis set possess MAE’s of only 0.06

eV (see SI, Table S3.4 and Figure S3.5). Moreover, MAXE is reduced to 0.12 eV for M06-2X

functional in combination with the RPA-approach and def2-SVP basis set. Therefore,

employing global hybrids with large HF-exchange contribution and energy scaling might be a

suitable, albeit somewhat empirical approach to estimate the absorption energies of PPs.

3.3.3 Range Separated Hybrid Functionals

Page 60: Theoretical Investigations of the Photophysical Properties

54

Range separated hybrid functionals (RSH) possess a different contribution of HF exchange in

short and long interelectronic distances. Short range corrected functionals, like HSE06, possess

a medium amount of HF exchange in the short range while it drops commonly to zero at long

interelectronic distances. This allows a faster calculation of solid-state properties compared to

global hybrid functionals but an improvement for excited states cannot be expected for this

type of functional. In contrast, long-range corrected (LC) RSH possess a large amount of HF

exchange at long interelectronic distances. This significantly reduces errors originating from

Rydberg states and charge transfer excitations in TD-DFT calculations.183, 207-209 Therefore, we

will focus solely on LC-RSH.

The performance of TDA for the investigated RSH functionals is overall comparable to that of

global hybrids, as visible in Table 3.3, 3.4 and Figure 3.2a. Additionally, TDA tends stronger

to over-estimate the absorption energies. The RPA approach for CAM-B3LYP results in

overestimated absorption energies, while other LC-RSH functionals tend to underestimate

these energies. Overall, RPA has a strong dependency on the type of functional and produce

lower absorption energies than TDA (see Table 3.4 and Figure 3.2b).

Table 3.4 Calculated original error values in eV for investigated RSH functionals

Range Separated

Hybrid Functionals

TDA

(def2-

SVP)

RPA

(def2-

SVP)

sTDA

(def2-

SVP)

sTD-DFT

(def2-

SVP)

sTDA

(def2-

TZVP)

sTD-DFT

(def2-

TZVP)

ωB97

ME 0.21 -0.16 -0.16 -0.45 -0.18 -0.45

MAE 0.21 0.16 0.16 0.45 0.18 0.45

MAXE 0.31 0.23 0.39 0.69 0.40 0.68

ωB97X

ME 0.22 -0.08 -0.18 -0.43 -0.20 -0.42

MAE 0.22 0.09 0.18 0.43 0.20 0.42

MAXE 0.30 0.17 0.41 0.67 0.42 0.65

Page 61: Theoretical Investigations of the Photophysical Properties

55

LC-BLYP

ME 0.20 -0.09 -0.13 -0.35 -0.16 -0.35

MAE 0.20 0.10 0.15 0.35 0.16 0.35

MAXE 0.28 0.18 0.37 0.59 0.38 0.58

CAM-

B3LYP

ME 0.24 0.07 0.03 -0.02 -0.13 -0.17

MAE 0.24 0.08 0.05 0.05 0.13 0.17

MAXE 0.31 0.16 0.10 0.14 0.27 0.29

On the other hand, as can be seen in Table 3.4, Figure 3.3a and 3.3b, the performance of sTDA

and sTD-DFT is improved over global hybrids only for CAM-B3LYP which amounts up to

46% HF-exchange. In contrast to CAM-B3LYP, LC-RSH functionals incorporating up to 85-

100% HF-exchange result in remarkably large errors. This is in agreement with the original

work where sTDA was developed for long-range corrected hybrids where best results of

excitation energies were obtained with CAM-B3LYP for charge transfer free systems.210

Employing large basis sets does not improve the results based on simplified approaches. (see

Figure 3.2 Boxplot displaying original error values in eV for the variant of density-functionals in

combination with TD-DFT types (a) TDA, (b) RPA, and def2-SVP basis set. Here, MAE is

denoted by (+) while outliers (•) termed as extremes error values which are outside the range

given in bars. (scaled error values can be seen in SI, Figure S3.3)

Page 62: Theoretical Investigations of the Photophysical Properties

56

Figure 3.4a for e.g. CAM-B3LYP case and Figure S3.5 for all the tested functionals). Thus,

results obtained with the def2-SVP in combination with simplified approaches are most reliable

and reasonable.

Figure 3.3 Boxplot displaying original error values in eV for the variant of density-functionals in

combination with TD-DFT types (a) sTDA, (b) sTD-DFT, and def2-SVP basis set. (scaled error

values can be seen in SI, in Figure S3.4).

Figure 3.4 Boxplot displaying error values in eV, where (a) original and (b) scaled for the selected

DFT functional in combination with different basis-set qualities (def2-SVP and def2-TZVP) and

simplified time dependent approaches (sTDA and sTD-DFT).

Page 63: Theoretical Investigations of the Photophysical Properties

57

To sum up and as shown in Figure 3.5a, GGA functionals produce large errors in the form of

outliers e.g., BLYP in combination with def2-SVP basis set. Global hybrid functionals, e.g.

B3LYP, produces MAE of 0.06 eV in combination with sTDA and the def2-SVP basis set and

appears to produce overall reliable results. However, best results, and indeed excellent ones,

are obtained with the RSH, CAM-B3LYP in combination with sTDA and an overall small

def2-SVP basis set which yields a MAE of about 0.05 eV. This can be barely improved by

energy scaling (see Figure 3.5b and scaled error values are listed in Table S3.5).

3.3.4 Double Hybrid Functionals and post-Hartree Fock approaches

In addition to the exact HF exchange, double hybrid functionals include a second order

perturbation theory correction term (MP2) for the correlation part of the functional. This

improves mainly the consideration of dispersion forces. However, the computational time is

comparable to MP2. Therefore, we have also included some traditional post-HF approaches

with comparable computational cost such CIS and CIS(D), in our study.

Figure 3.5 Boxplot displaying error values in eV, where (a) original and (b) scaled for the selected

functionals from DFT-group in combination with TD-DFT types and def2-SVP basis set. A detailed

box-plot representation (original and scaled error values) of all the investigated density functional-

approaches-basis set combinations can be seen in SI, Figure S3.5 (in eV) and Figure S3.6 (in nm).

Page 64: Theoretical Investigations of the Photophysical Properties

58

Table 5. Calculated original error values in eV for double hybrids and post HF methods

Approach def2-SVP

B2PLYP

ME 0.24

MAE 0.24

MAXE 0.31

B2GP-PLYP

ME 0.29

MAE 0.29

MAXE 0.37

mPW2PLYP

ME 0.25

MAE 0.25

MAXE 0.33

CIS

ME 0.26

MAE 0.26

MAXE 0.39

CIS (D) ME 0.51

MAE 0.51

MAXE 0.60

Figure 3.6 Boxplot displaying error distribution in eV, where (a) original value and (b) scaled values

for the double hybrid functional and post HF methods in combination with def2-SVP basis-set

Page 65: Theoretical Investigations of the Photophysical Properties

59

As can be seen in Table 3.5 as well as in Figure 3.6a, double hybrid functionals produce large

errors comparable to CIS. Including perturbative double corrections results in even larger errors

of the CI approach. However, we would like to highlight that the employed basis set is only of

double-ζ quality due to the system size. The strong systematic overestimation of absorption

energies for the double hybrid functionals and post HF methods suggests a scaling of the

obtained absorption energies. Indeed, the results are significantly improved and an accuracy

comparable to CAM-B3LYP can be reached (see Figure 3.6b and Table S3.6). Thus, CIS(D)

with scaled absorption energies might be a suitable approach to verify results obtained with

sTDA, def2-SVP and CAM-B3LYP.

3.3.5 Influence of diffuse basis set functions and ground state structure

Commonly, diffuse basis sets are recommended for weakly bound electrons found in anions or

in excited states. Therefore, we have selected three functionals, BLYP, B3LYP and CAM-

B3LYP, and extended the employed Ahlrichs basis set by diffuse functions. Independent of the

employed functional type, including diffuse basis set functions provides poorer results

compared to the def2-SVP double-ζ basis set, (see SI, Figure S3.7a and Table S3.7). The worse

performance cannot explained by the ϵ (HOMO) criterion211 (see SI, Table S3.8). Also scaling

of energy does not improve results since including diffuse basis sets increases the scattering of

the calculated absorption energies in most cases (see SI, Figure S3.8a). Thus, unintuitively, the

smallest basis set provides the most accurate results.

Finally, we investigated the influence of the electronic structure method during structure

optimization. Instead of the GGA BLYP, the more expensive hybrid functional B3LYP was

selected for structure optimization. The influence is overall negligible for absorption energies

obtained by BLYP and B3LYP, compare Table S3.7 and S3.9 in SI. In case of the RSH CAM-

Page 66: Theoretical Investigations of the Photophysical Properties

60

B3LYP, the errors without energy scaling are even increased pointing to some error

compensation for the most reliable approach (see SI, Figure S3.7b). Nonetheless, global scaling

of energy provides nearly identical results. Thus, as long as the correct combination of scaling

factor, structure optimization setup and absorption energies calculation approach are selected,

results can be barely improved.

3.4 Conclusions

We have presented a detailed validation of the simplified time dependent density functional

theory method developed by Grimme et al. for the calculation of UV-Vis-spectra of

porphyrinoids including free base and metal containing PPs. The original RPA-approach tends

to smaller absorption energies than TDA, which is also visible for the simplified versions.

Local GGA functionals produce large errors and therefore cannot be recommended. In contrast

to local DFT functionals, global hybrids yield significantly improved results only after energy

scaling, especially BHLYP and M06-2X. We can recommend as global hybrid B3LYP which

produces MAE of 0.06 eV in combination with sTDA and the def2-SVP basis set which can

be barely improved by energy scaling. Best results without energy scaling are obtained with

the RSH CAM-B3LYP in combination with sTDA and the def2-SVP basis set yielding a MAE

of about 0.05 eV. Significantly more expensive perturbative corrected double hybrid

functionals tend to yield results comparable to CAM-B3LYP solely when energies are scaled.

Apart from that, employing a hybrid instead of a GGA functional for geometry optimization

has less-significant effect on the calculated absorption bands whereas increasing the basis set

does not improve the calculated absorption bands. Most notable, including diffuse basis

functions even leads to worse results. Thus, employing a cheap GGA like BLYP for structure

optimization, selecting an overall small basis set of double-ζ quality in combination with a

CAM-B3LYP and the sTDA approach provides a cost-efficient approach to estimate the

Page 67: Theoretical Investigations of the Photophysical Properties

61

absorption spectra of porphyrinoids which can be barely improved by more expensive

approaches. Unfortunately, none of the local functionals has sufficient predictive power, which

is an obstacle in particular for periodic calculations.

Page 68: Theoretical Investigations of the Photophysical Properties

62

Page 69: Theoretical Investigations of the Photophysical Properties

63

Chapter 4

Computational Screening of Surface Mounted Metal-

Organic Frameworks Assembled from Porphyrins

"No amount of experimentation can ever prove me right;

a single experiment can prove me wrong." – Albert Einstein

The studies summarized in this Chapter have been published as:

Bridging the Green Gap: Metal-Organic Framework Heteromultilayers Assembled from

Porphyrinic Linkers Identified Using Computational Screening

Ritesh Haldar+, Kamal Batra+, Stefan Michael Marschner, Agnieszka B. Kuc, Stefan Zahn,

Roland A. Fischer, Stefan Bräse, Thomas Heine, and Christof Wöll

Chem. Eur. J. 2019, 25, 7847-7851 © Wiley-VCH Verlag GmbH & Co. KgaA

Page 70: Theoretical Investigations of the Photophysical Properties

64

This Chapter investigates the potential of computational screening methods for tailoring the

photophysical properties of porphyrin-based surface-mounted metal-organic framework (PP-

SURMOF) thin films. This study involves a close collaboration between experiment and theory.

First, we briefly mention their contributions followed by the Chapter’s outline:

Experimental contributions: (i) Synthesis and characterization of multi-functionalized PP-

linkers are led by S. Marschner and S. Bräse; (ii) Thin film depositions and photophysical

characterization of PP-based SURMOFs are led by R. Haldar, C. Wöll, and R. Fischer.

Theoretical contributions: (i) Electronic structure and properties calculations of multi-

functionalized PP-linkers are led by me, S. Zahn, and T. Heine (ii) Computational screening

and band structure calculations of PP-based SURMOFs are led by me, A.B. Kuc, and T. Heine.

In Section 4.1, after a brief outline of organic photovoltaics followed by their shortcomings,

we introduce the benefits of crystalline systems and an approach based on the regular assembly

of PP-linkers into thin-films of MOFs. Then we summarize the shortcomings of other PP-

based thin-film approaches and the scope of PP-linkers to assemble in SURMOFs. Finally, we

investigate the tuning of characteristic Q-bands of the multi-functionalized PPs by using a

validated computational protocol (see Chapter 3) for improving their absorption efficiency.

Section 4.2 lays out a summary of the computational methodologies utilized in this study.

Besides, we highlight the three interesting strategies for tuning the structure and absorption

efficiency of PPs on behalf of computational screening. Finally, we select three promising PPs

exhibiting important differences in their characteristic Q-bands and utilize them for

assembling into SURMOF structures.

In Section 4.3, first, we compare the predicted absorption spectra with experimental results for

the promising PPs. Next, to rationalize the experimental absorption spectra of the PP-

Page 71: Theoretical Investigations of the Photophysical Properties

65

SURMOFs, we investigate the impact of the arrangement, also known as stacking, of the PP-

linkers in the framework structures by calculating the respective electronic band structures.

Furthermore, we estimate an indirect band gap having band dispersion for one of the

promising structures, that is substantially larger (≈200meV) than the previously reported

(≈5meV) value for a different PP‐SURMOF.

Finally, in Section 4.4, we present an attractive route to prepare chromophoric assemblies by

the SURMOF-based approach and demonstrate the power of computational screening methods

to identify the most promising PP linkers for the construction of layered PP-SURMOFs with

the desired photophysical properties.

Page 72: Theoretical Investigations of the Photophysical Properties

66

4.1 Introduction

While presently the highest conversion efficiencies are achieved with inorganic (Si, Ge)212-213

and hybrid (Grätzel-cell, perovskites)214-217 semi-conductor based photovoltaic (PV) devices,

organic materials are an interesting alternative with advantages in more specific

applications.218-219 Main obstacles in this area are efficiency and stability issues. Progress in

this field is only gradual, and the search for organic molecules suited as an active PV material

is to a large extent still dominated by empirical approaches.218-221 Use of computer-based

screening methods is hampered by the fact that the characteristics of most organic PV (OPV)

materials are strongly impacted by intermolecular interactions.222 and, for disordered systems,

many configurations have to be samples to make reliable predictions.223

Major progress has thus to be expected when employing crystalline systems with exactly

known structures.224 In such cases, periodic boundary conditions (PBCs) can be applied, which

allows for a thorough theoretical analysis utilizing the solid state analysis toolbox.223 The OPV

device characteristics can then be predicted on the basis of first-principles electronic structure

calculations carried out both for single molecules as well as for the condensed phases of these

organic semiconductors. In the latter case, the precise arrangement of the functional molecules

in the crystalline state is taken fully in account. In addition to a reliable theoretical analysis

with high predictive power, for crystalline systems additional effects may be employed, for

example the emergence of band dispersion favoring large charge carrier mobilities and the

formation of indirect band gaps.225-227 Indirect band gaps are beneficial in photovoltaics since

they favor a fast and highly efficient charge separation and thus charge carrier recombination

is strongly suppressed.

Herein, we demonstrate an approach based on assembly of multi-functionalized PP linkers into

thin films of MOFs.228-229 PPs are a particular promising class of organic compounds for

Page 73: Theoretical Investigations of the Photophysical Properties

67

investigating the beneficial effect of a regular arrangement of photoactive molecules with

regard to light harvesting.71, 230-233 PPs form a very rich class (>100,000 compounds known)131,

234-240 which are also common in nature. In plants, PPs like chlorophyll transform solar energy

into chemical energy. Since PPs are among the best-performing organic compounds regarding

photon absorption, charge separation, and stability,131, 235-240 numerous previous works have

been carried out with the aim to prepare well-defined, thin films of PP aggregates deposited on

conducting and transparent substrates.241-243 The approaches employed previously include

vapor-phase deposition of PP molecules, a rather sophisticated method, or self-assembly from

appropriate solutions. The latter strategy has been rather successful, but the resulting systems

do not exhibit a high degree of ordering, thus largely excluding the formation of band-structure

effects. In addition, the different types of molecular packing in the amorphous films make more

precise predictions of absorption properties difficult.

The typical absorption spectrum of a PP chromophore consists of a sharp, high-intensity Soret

band, which is typically located in the UV region, and four Q bands located in the visible range

(Figure 4.1a). In principle all transitions of the frontier orbitals are symmetry-allowed.

However, as the frontier orbitals are nearly degenerate, the electronic structure can be

approximated by a 18π cyclic polyene model, as suggested by Gouterman,13 in which two

Figure 4.1. a) Typical spectrum of a porphyrin. Blue lines in the Lewis structure highlight the π-

electrons on which a 18π cyclic polyene model is reasonable; b) Comparison of the frontier orbitals

obtained from DFT calculations of porphyrin and the 18π cyclic polyene model.

Page 74: Theoretical Investigations of the Photophysical Properties

68

transitions are allowed between the degenerated frontier orbitals (Δk=1), whereas two are

forbidden (Δk=9) (Figure 4.1b). Detailed quantum chemical studies confirm Gouterman’s

conclusions and explain the typical absorption spectra of PPs in detail.14-15 Tuning the energy

levels of HOMO-1, HOMO, LUMO and LUMO+1 strongly affects the absorption intensity of

the characteristic absorption bands of porphyrins. The higher the energy gap between HOMO-

1 and HOMO as well as LUMO and LUMO+1, the stronger will be the absorption intensity of

the Q-bands. While this allows the rational design of PP molecules, the prognostication of

absorption spectra for PP-based SURMOFs remains difficult. This is mainly due to the fact

that the bulk structure is not known a priori, and changes of absorption intensity and band

positions resulting from intermolecular interactions are thus difficult to predict.

Herein, we circumvent the problem of structure prediction for the condensed phase by focusing

on porphyrinic dicarboxylic acids that can be used as ditopic linkers to assemble PP-MOFs. In

order to identify a small set of chromophoric MOF linkers to cover the green gap we first

carried out a computational screening of a library containing 14 PP structures, see Supporting

Information (SI), Figure S4.1 in appendix B.2). This library was generated by functionalization

of two of the phenyl rings at the meso-positions with carboxylate groups to produce ditopic

MOF linkers. In the first group of PP derivatives, the H atoms at the β-positions are substituted

by halides and methyl groups (1-X group, X = F, Cl, Br, CH3), and in the second group the two

phenyl rings that do not serve as linking groups are substituted. The library contains

substitution groups which can modulate PP ring planarity (e.g. substitution at β-position by

bulky Br atoms), and electron density (electron donating /pulling).

To obtain theoretical predictions with sufficient reliability, simplified time-dependent

functional approach was applied (see section 3.2, computational methodologies). As expected,

functionalization is found to be a suitable means for both Q-band intensity enhancement and

Page 75: Theoretical Investigations of the Photophysical Properties

69

tuning of absorption band positions. For example, β-substitution by Br as well as meso-

positioned phenyl group substitution creates a large red shifts of Q bands by >50 nm in

comparison to 1-H (see Table S4.2). Additionally, the absorption of the Q bands is significantly

enhanced (Figure S4.3).

3.2 Computational Methodologies

For molecular systems: The CAM-B3LYP functional183 in combination with the TZVP basis

set170 was employed in all calculations. The energy convergence criterion of the self-consistent

field cycle was set to 10−8 Hartree and dispersion forces were considered by the 3rd version of

Grimme’s empirical dispersion correction D3 in combination with the improved Becke–

Johnson damping.166, 244 To speed up the density functional theory calculations, the resolution

of identity approximation167-168, 245 (RI) was employed for the Coulomb integrals while

exchange contributions were accelerated by the “chain of spheres” approximation191 (COSX).

After structural optimizations, excited states were calculated by the simplified time-dependent

functional theory.107 We have employed the CAM-B3LYP functional, since this long-range

corrected hybrid functional keeps the accuracy of B3LYP for excited states, but nearly

eliminates errors from charge transfer excitations.183 All calculations were carried out

employing the ORCA 4.1.1 program package.129

For solid-state systems: The structure models have been created using AuToGraFS125. The

frameworks structure and lattice parameters of porphyrin based SURMOFs were pre-optimized

using Universal Force Field (UFF)126 employing UFF4MOF parameters.246 The bond order

specified between paddlewheel Zn atoms was set to 0.25, paddlewheel Zn-O bonds were set to

0.50, all other bond orders were specified according to the standard chemical notation.

Frameworks geometry optimizations have been carried out using the self-consistent charge,

density functional based tight-binding (SCC-DFTB) method including UFF-dispersion along

Page 76: Theoretical Investigations of the Photophysical Properties

70

with DFTB.org/3ob-3-1 parameters. The optimized lattice constants (a, b, and c) calculated for

the periodic frameworks are given in Table S4.5. Single-point calculations for band gaps and

band structures were performed using CRYSTAL17130 program along with DFT PBE173-174

functional, pob-TZVP247 basis type and 100 k-points in the Monkhorst-Pack mesh. Calculated

band-gaps for 1-Br ', 5 ' and 10 ' are given in the SI, Table S4.6. Also, stacking information of

PP-SURMOFs are given in the SI, Table S4.7.

3.3 Results and Discussion

On the basis of this computational screening, three PPs (Figure 4.2a) have been selected for

synthesis (see experimental section in the SI) and validation of the theoretical predictions.

Selection criteria were the enhancement of Q-band intensity and the tuning of band position.

These selected three PPs exhibit important differences: i) the presence of electron withdrawing

functional groups (fluorinated phenyl), ii) π-conjugation of the substituent (phenyl group or

acetylene group), and iii) planarity of the porphyrin core (twisted octabromo porphyrin). The

Figure 4.2. a) Three selected PP linkers; (b) TD-DFT predicted (left) and experimental (right) UV-

Vis spectra of the linkers. Experimental spectra were recorded for 20 M ethanolic solution of the

linkers at RT.

Page 77: Theoretical Investigations of the Photophysical Properties

71

experimental methods developed for organic synthesis of the multi-substituted porphyrins were

successful. The required A2B2-type porphyrins were assembled in a modular fashion starting

from dipyrromethene building blocks and functional aldehydes. Subsequent cross-coupling

reactions led to the designed porphyrins (see Experimental Section in appendix B.2). Figure

4.2b, shows the predicted (left) and experimental (right) UV-Vis spectra of 1-Br, 5 and 10.

Generally, a good agreement is seen between the TD-DFT gas-phase predictions and the

experimental results (solvated). Note, that the agreement is not expected to be quantitative since

solvatochromic shift affect the experimental absorption spectra.248-249 In these synthesized PPs,

apart from the Q-bands in near-IR region, also the Soret band shifts to visible range, in

comparison to the 1-H.

For applications, the assembly of PP dyes into thin films of high optical quality is crucial.

Therefore, in a final step, the selected PP linkers were assembled into SURMOF-2 thin films

containing optically silent Zn2 paddle wheel units, following the spin-coated variant of the

layer-by-layer (LbL) liquid phase epitaxy method.68, 250-253 The SURMOF-2 structure consists

of Zn2-paddle-wheel type secondary building units (SBUs) tethered with ditopic linkers to yield

two-dimensional planes stacked along the (010) crystallographic direction (Figure 4.3).60 Note

that the framework of SURMOF-2 has P4 symmetry, and the lattice constant only depends on

Figure 4.3. a) Simulated XRD pattern of 1-Br', and experimental out-of-plane XRD patterns of 1-

Br', 5' and 10'; calculated atomistic PP-based Zn-SURMOF structures of b) 1-Br', c) 5', and d) 10',

view along [010] direction

Page 78: Theoretical Investigations of the Photophysical Properties

72

the length of the PP linkers (i.e. distance between the two carboxylic acid groups), while the

addition of side groups to the PPs are not expected to create changes of the SURMOF unit cell.

In all three cases, SURMOF-2 structures, labelled as 1-Br ', 5 ' and 10 ', exhibit well-defined

out-of-plane XRD patterns (Figure 4.3a). The position of the diffraction peaks is almost

identical, yielding unit cell dimensions with a = b = 2.3 nm, in excellent agreement with the

prediction. These observations confirm the expectation that irrespective of the type of group

attached at meso or β-position to the PPs, the unit cell parameters of the resulting MOF

structures remain unchanged. (Figure 4.3, for details, see SI). Figures 3b-d show the DFTB

optimized atomistic structures of 1-Br ', 5 ' and 10 '. In 10, fluorination of the linkers creates

significant electrostatic repulsion, so the inter-linker distance is maximized when the PPs are

untwisted. This maintains the overall P4 symmetry in 10 with a distance between the PP units

of 6.3 Å. In contrast, the longer linkers in 5′ interact strongly, causing a tilting of 32° and

reducing the distance between the PP planes to 3.3 Å.

Figure 4.4 compares the UV-Vis spectra of 1-Br ', 5 ' and 10 ' with those of the solvated linker

molecules 1-Br, 5 and 10. In 1-Br ', the distorted basal planes of the linker (1-Br) prevent

periodicity along the PP stacks (along (010)). Consequently, the bands are not shifted compared

to those of the solvated 1-Br, but significantly broadened (Figure 4.4a). However, the

Figure 4.4 Experimental UV-Vis comparison spectra of the synthesized PP linkers (solid line) and

their corresponding fabricated SURMOF-2 structures (broken line); a) 1-Br, b) 5 and c) 10. Note

that the molecular spectra are recorded in ethanol at RT.

Page 79: Theoretical Investigations of the Photophysical Properties

73

experimental data reveal that crystallization of 5 in the SURMOF-2 structure causes a

significant redshift of all bands in 5 ' (Figure 4.4b). Notably, this redshift affects both the Soret

and the Q-bands. In addition, all bands are substantially broadened. The spectra of 10 and 10 '

are very similar, indicating negligible interaction between the PP linkers (also in line with the

predicted structure) (Figure 4.4c).

To rationalize the experimental UV-Vis spectra recorded for the PP-SURMOFs, we have

calculated the band structures of 1-Br ', 5 ' and 10 ' (for details, see SI), which are shown in

Figure 4.5(a-c). The band structure of 1-Br ' reflects the idealized structure and therefore is not

directly suitable to interpret the UV-Vis spectrum (Figure 4.5a). For 5 ', we observe significant

band dispersion along the Z direction of the Brillouin zone (see Figure 4.5b, 4.5d), which is

responsible, both for the redshift of the absorption bands, as well as for the charge carrier

mobility along the PP stacks (along (010)). We observe an indirect band gap of ~200 meV,

Figure 4.5. (a-c) Band structure of 1-Br, 5′ and 10′; (d) brillouin zone; (e) a schematic illustration

of trilayer SURMOF (left) and corresponding UV-Vis spectrum (right). (The gray filled box = SBU,

filled diamond shape = PP).

Page 80: Theoretical Investigations of the Photophysical Properties

74

which is substantially larger than that (5 meV) reported in previous work for a different PP-

SURMOF-2 structure.72 For 10 ', no band dispersion is observed, confirming the hypothesis

that the linker molecules 10 do not interact in the SURMOF-2 structure (Figure 4.5c). As a

result, the UV-Vis data for the corresponding SURMOF-2 film mainly show the molecular

features. Due to the lack of band dispersion, no ballistic transport is possible for such structure.

While for the 1-Br′ and 10′ no substantial aggregation-induced shifts of the absorption band

positions were observed, a broadening of the absorption bands is evident. This is a beneficial

effect, and together the three PP-linkers allow covering the full visible spectrum ranging from

violet to near-IR. Inspired by such broad absorption, we have fabricated a multilayer

SURMOF-2 structure by employing heteroepitaxy.59, 254 Layers of 1-Br ', 5 ' and 10 ' were

sequentially deposited on top of each other to make a crystalline thin film with a broad

absorption ranging UV-to-NIR, as shown in Figure 4.5e. Such straightforward fabrication

method, combining all the potential photon absorbing dyes as a thin film is a promising strategy

towards further improvement of OPV materials.

4.4 Conclusions

In conclusion, we present here an attractive route to create OPV absorber layers in the visible

regime, covering the entire range from the violet to the near infrared, including the green gap.

Starting point is a set of candidate structures, in this case, tetraphenyl porphyrin derivatives

with different substitution patterns. Instead of scheduling a vast amount of structures for

synthesis, we used a computational approach to first select promising chromophoric MOF

linkers from an in-silico library. Indeed, after synthesis, the three candidates, showed excellent

performance. 1-Br exhibits distorted basal plane that imposes a significant red shift of both

Soret and Q-bands in the free molecule. The second one (5) has a large aromatic ligand

(extended conjugation), which enhances the Q-band intensity, but does not affect the position

Page 81: Theoretical Investigations of the Photophysical Properties

75

of the Soret band. The third one (10) is functionalized with strong electrostatically active

groups. Arranging these dyes into a SURMOF-2 structure yields three different pictures: In 1-

Br′, distorted basal planes of the linkers avoid regular stacking, and while the positions of the

absorption bands of the free linkers are maintained, they are significantly broadened in the

SURMOF structure. In 5′, the strong intermolecular interactions lead to a red shift and

broadening of all bands. Finally, in 10′, electrostatic repulsion reduces intermolecular

interactions and thus the absorption spectrum of the corresponding SURMOF is dominated by

the molecular properties of 10.

Altogether our results reveal that the strategy to prepare chromophoric assemblies via a

SURMOF-based approach carries a huge potential. Depending on the stacking within the MOF,

one can realize systems with small inter-chromophore coupling, which are essentially

dominated by features of the individual molecules. By adjusting the MOF topology together

with chromophore substitution patterns, intermolecular couplings can be introduced, which

allow to invoke band structure effects, thus increasing charge carrier mobility and allowing of

indirect band gap formation. Finally, the SURMOF approach then provides the prospect to

realize energy funneling via fabrication of hetero-multilayers or gradient structures.

Page 82: Theoretical Investigations of the Photophysical Properties

76

Page 83: Theoretical Investigations of the Photophysical Properties

77

Chapter 5

The Proximity Effect in Porphyrin-based Surface-

Mounted Metal-Organic Frameworks

"The important thing in science is not so much to obtain new facts

as to discover new ways of thinking about them." – Sir William Bragg

The studies summarized in this Chapter have been published as a part of the

Progress Report:

Proximity Effect in Crystalline Framework Materials: Stacking-Induced Functionality in

MOFs and COFs

Agnieszka B. Kuc, Maximilian A. Springer, Kamal Batra, Rosalba Juarez-Mosqueda,

Christof Wöll, and Thomas Heine

Adv. Funct. Mater. 2020, 30, 1908004 © Wiley-VCH Verlag GmbH & Co. KGaA

Page 84: Theoretical Investigations of the Photophysical Properties

78

This Chapter discusses the tuning of stack interactions in layered porphyrin-based surface-

mounted metal-organic frameworks (PP-SURMOFs) with a future goal for identifying the most

promising molecular framework structures by proper selection of functional groups in the PP-

linkers. The exact stacking in PP-SURMOF layers is of paramount importance for their

photophysical properties.

In Section 5.1, after an introduction of molecular framework materials followed by discussing

the proximity effect, which is caused by van der Waals interactions between stacked aromatic

molecules, we present the impact of proximity effect that results in the first reported indirect

band gap formation in a PP-SURMOF. Then we briefly summarize the shortcomings of the

reported PP-SURMOF and the scope of functionalizing the PP-linkers. Finally, we present

three most promising PP-linkers and incorporate them into SURMOF to examine their effects.

Section 5.2 lays out a short summary of the computational methodologies utilized in this study

to predict and analyze the proximity effect.

In Section 5.3, we discuss the impact of the proximity effect by analyzing the stack interactions

within three most promising PP-SURMOFs. For one of these PP-SURMOFs, we investigate

the band dispersion as a function of the rotation of the functionalized PP-linker in a bulk

framework. Furthermore, we present that different degrees of structural rotations lead to

different structures in the electronic bands. i.e. dispersion in the valence and conduction bands.

Finally, in Section 5.4, we observe substantial dispersion of both band edges for a structure

with out-of-plane rotated (by 110) substituents and this shows that the proximity effect can

render strong electronic effects if the aromatic molecules are arranged in exact or well-

controlled stacks.

Page 85: Theoretical Investigations of the Photophysical Properties

79

5.1 Introduction

Molecular framework materials, including metal-organic frameworks (MOFs),26, 255-256

coordination polymers,257 and covalent organic frameworks (COFs),258-259 provide an

intriguing bridge between chemistry and solid-state physics. They are composed of molecular

units that may carry the whole range of functional groups known to chemistry. These molecular

building blocks are stitched together by strong bonds. This setup provides much higher

chemical, thermal, and mechanical stability as compared to molecular crystals and, thus, allows

the formation of large pores which can reach up to 10 nm in diameter,60, 260-262 as well as

extremely large internal surface areas.261 These unique properties have motivated intense MOF

research during the past years: they allow high gas uptake capacities and, thus, application in

methane and hydrogen storage.263-264 Coupling these properties, intrinsic to porous materials,

with molecular functionalities integrated into nodes and linkers can yield to multifunctionality:

selective uptake, CO2 capture265-266, hydrogen isotope separation,267-268 as well as switching

permeance and selectivity via optical269 or electrical270 switching. Even though the IUPAC

recommendations for nomenclature explicitly specifies that these materials are not necessarily

crystalline,271 high crystal order is observed for many of them.

Crystallinity allows high-quality structural analysis by experimental methods, e.g., via X-ray

diffraction (XRD). Consequently, theoretical work can be carried out in a straightforward

fashion and can then be compared directly to experimental results, thus, allowing for a direct

validation of computational approaches. The well-defined structure also makes a proper

physicochemical characterization of these materials possible and provides, thus, the basis for

the high level of understanding of their structure and electronic structure. At the same time,

crystal order is the reason for solid-state effects that are caused by the translational symmetry,

such as indirect band gaps and ballistic charge transport.272 As in other crystalline solids,

Page 86: Theoretical Investigations of the Photophysical Properties

80

defects play a critical role for certain properties, including electronic and optical properties,

thus, a proper characterization of defect types and their density is crucial.58, 273 This point is

particularly important, as the properties discussed below are a direct consequence of the

crystallinity of the framework materials.

Ballistic transport is typically hindered either by the large effective masses of the typically

dispersion less bands or by the chemical composition of the frameworks, where non-carbon

centers (for example oxygen, boron, or nitrogen) can effectively block electron conjugation.

Only recently, ballistic transport, facilitated by electron conjugation, has been demonstrated in

layers of two-dimensional (2D) crystalline frameworks.274-277 There is, however, a more subtle

way to implement strong band dispersion in crystalline molecular framework materials:

Controlled stacking of aromatic molecules, incorporated into the materials as linkers or as

pillars with suitable intermolecular distance, is subject to the proximity effect, more precisely,

the molecules interact via π-stacking, which causes strong alterations of the electronic structure

of the framework materials. While this type of stacking is rather obvious for the class of layered

crystalline frameworks278 with atomically thin layers and with aromatic connectors or linkers,

it can also be achieved in MOF thin films with suitable crystal structure.72, 135 A further option

of MOFs is to grow superstructures by applying layer-by-layer (LbL) procedures (MOF-on-

MOF) and to fabricate structurally well-defined organic/organic heterointerfaces.251, 279

Several groups have introduced LbL methods to deposit MOF thin films on substrate, first, in

2007, Wöll and Fischer reported on the LbL route to MOF synthesis,50, 68, 280 and later H.

Kitagawa and coworkers.69 Such surface-mounted MOFs are referred to as SURMOFs. They

exhibit high crystallinity and can be investigated using virtually all surface science techniques,

see also Section on SURMOFs synthesis in the SI. SURMOF-2, being an iso-reticular series

based on MOF-2,281-282 is one of the simplest MOF architectures suited for LbL growth.60 They

Page 87: Theoretical Investigations of the Photophysical Properties

81

are derived from MOF-2, a bulk framework material based on paddle-wheel units with four

dicarboxylate groups and typically Cu2+ or Zn2+-dimers connected to ditopic organic linkers of

different length, the shortest one being 1,4-benzene dicarboxylate. The length of the linkers

determines the lattice constant and, thus, the pore size of the resulting SURMOFs, where up to

4 nm in diagonal have been reached up thus far.60 Layers of such SURMOFs form square

lattices and, theoretically, could be stacked together in three different arrangements i.e.

eclipsed, slipped, and inclined. The most symmetric P4 variant is the eclipsed stacking with

linkers and connectors in one layer directly on top of linkers and connectors in another layer.

This stacking is found in all SURMOF-2 derivatives discussed in this Chapter and leads to a

less stable system due to unfavorable non-covalent interactions. Computational investigations

showed60 that two other stackings are energetically more favorable. These are slipped and

inclined stackings, with P2 and C2 symmetries, respectively, which emerge in bulk

synthesis.281, 283-284 However, in the SURMOF approach, the metastable P4 symmetry with

eclipsed stacking is enforced by the anchoring of the first MOF layers to the nucleating surface.

The first report on indirect band gap formation in MOFs was published in 2015.72 Thin films

of epitaxial MOFs have been studied, and photoinduced charge-carrier generation was

observed. The investigated Zn-paddle wheel SURMOF-2 derivative utilized Pd-porphyrinoid

linkers (Pd-PP-Zn-SURMOF, see Figure 5.1a). As parent SURMOF-2, this structure is also a

square lattice in-plane and layers are stacked in the AAAA fashion. Such a system results in

fairly dispersion less bands of the electronic structure (see Figure 5.1b), however, zoom-in to

the conduction and valence bands reveals a small but distinct dispersion in an order of 5meV.

This value corresponds to a mobility of about 0.003 cm2 V-1 s-1, which at that time was larger

than for any other MOF. This MOF exhibits an indirect band gap, which should result in

suppressed electron-hole recombination and improved photovoltaic properties in such organic-

semiconductor based devices.

Page 88: Theoretical Investigations of the Photophysical Properties

82

The photovoltaic efficiency of the reported PP-SURMOF amounts to only 0.2%,72 and is thus

far too low for realizing a competitive device. Another issue is the absorbance of the PP itself:

the strongly absorbing Soret band is in the ultraviolet, while the Q-bands, which are located in

the visible spectrum, are only weakly absorbing. PP functionalization can strongly enhance the

absorbance of the Q-bands. Among the large number of possibilities, three particularly

interesting ones have been identified by combining rational design with computational

screening.135 (i) 1-Br, distorting the planarity alters the selection rules and, thus, enhances the

intensity of the Q bands. This can be achieved by bromination of the PP core (twisted

octabromo porphyrin). (ii) 5, the π-conjugation of the PP can be extended by adding a phenyl-

acetylene (PhA) group. (iii) 10, the π-system can be affected by the presence of electron-

withdrawing fluorinated phenyl substituents.135 All three strategies show an effect on the

Figure 5.1 (a) Building block (Pd-porphyrinoid, Pd-PP) together with the top and side views of Pd-

PP-based Zn-SURMOF, and (b) the corresponding band structure with zoom-in to the top of valence

and bottom of conduction bands adapted from Ref. [72]. Bands are fairly flat, however, small out-

of-plane dispersion occurs in the direction perpendicular to the layers. The dispersion is in the limit

of a couple of meV. Pictures of structures made with VESTA.

Page 89: Theoretical Investigations of the Photophysical Properties

83

isolated PP linker molecules, in particular, the Q-band intensity increases. Incorporated to a

SURMOF, these three linkers, however, show very different absorbance (see Chapter 4). This

can, again, be attributed to the proximity effect and will be discussed in section 5.3 after a short

summary of the computational methodologies used in this Chapter to further exploit this effect

by probing different degree of structural rotation for one of the promising PP-based SURMOFs.

5.2 Computational Methodologies

This section gives an overview of methods that are useful for computationally tackling layered

SURMOFs, from constructing the atomistic structures to obtain high-level electronic structure

data. It includes the methods that have been used (as in Chapter 4) to obtain the results.

The structure models have been created using AuToGraFS125. The frameworks structure and

lattice parameters of PP-based SURMOFs were pre-optimized using Universal Force Field

(UFF)126 employing UFF4MOF parameters. 246 The bond order specified between paddlewheel

Zn atoms was set to 0.25, paddlewheel Zn-O bonds were set to 0.50, all other bond orders were

specified according to the standard chemical notation. Frameworks geometry optimizations

have been carried out using the self-consistent charge, density functional based tight-binding

(SCC-DFTB) method including UFF-dispersion along with DFTB.org/3ob-3-1 parameters.

Single-point calculations for band gaps and band structures were performed using

CRYSTAL17130 program along with DFT PBE173-174 functional, pob-TZVP247 basis type and

100 k-points in the Monkhorst-Pack mesh. The optimized lattice constants (a, b, and c) and the

calculated band-gaps for 1-Br ', 5 ' and 10 ' are given below in Table S4.1 and S4.6 respectively

Also, the stacking information of above mentioned PP-SURMOFs are given in Table S4.7. (see

SI of Chapter 4). Besides, the calculated band gap values for rotations angles, αi and βi, of the

functional groups with respect to the PP plane of SURMOF (5’) are given in Table S5.1.

Page 90: Theoretical Investigations of the Photophysical Properties

84

5.3 Results and Discussion

For the brominated linker, the geometrically distorted building blocks fail to arrange

themselves into well-ordered stacks. As a result, loosely packed linker stacks with distances of

≈6.1 Å, too far to cause a significant proximity effect, are obtained. Hence, the incorporation

of the brominated PP (1-Br) into the MOF lattice does not affect the absorber properties

significantly. For the fluorinated species (10), also the formation of a well-ordered lattice is

Figure 5.2 (Top panel) Building blocks together with the top and side views of three Zn-SURMOFs

assembled from multi-functionalized PP linkers: (a) 1-Br ', (b) 5 ' and (c) 10 ' (Bottom panel) The

corresponding band structures. Adapted from the Chapter 4 or Ref. [135]. Strong band dispersion

observed for the PP-Zn-SURMOF (5 ') in the stacking direction, due to enhanced London dispersion

interactions between the PP linkers. Pictures of structures made with VESTA.

Page 91: Theoretical Investigations of the Photophysical Properties

85

reported. However, Coulomb repulsion keeps the linkers at the widest possible distance (≈6.3

Å) and the resulting absorption spectrum is very similar to that of the individual molecules. If

large and aromatic linkers are used (as in case of PA, (5)), the attractive London dispersion

interaction fosters linker rotation to the extent that the PP molecules form well-ordered stacks

Figure 5.3 (a) Top and side view of the 5 '. The interlayer distance (d = 3.3 Å) corresponds to the

distance between PP units in adjacent layers; (b) Cluster structure of the linker with possible

rotations angles, i and i, of the functional groups with respect to the PP plane; (c) Band structures

corresponding to different values of i and i. The strongest band dispersion was obtained for both

angles of about ±110°. The band structure from the Chapter 4 or Ref. 135 corresponds to the case

of i = 0° and i = 110°. Pictures of structures made with VESTA.

Page 92: Theoretical Investigations of the Photophysical Properties

86

with intermolecular distance of ≈3.3 Å, very close to that in graphite. Consequently, the band

structure shows strong dispersion, which results in a red shift of the Soret band and enhances

and broadens the Q-bands. These results are summarized in Figure 5.3 and reported in Ref.135.

To what degree is the spatial extension of the aromatic system of the individual porphyrinic

linkers relevant to the properties, which are mainly caused by the proximity effect in the linker

stack? To answer this question, we analyzed the band dispersion as function of the rotation

angle between the PP moiety and the PhA substituent (see Figure 5.3). This is achieved by

rotating parts of the linker (PhA and the benzene rings connected to the paddle wheel). Indeed,

such a rotation results in a strong manipulation of the electronic bands, from almost flat

conduction band in a hypothetical structure with all likers at 0 with respect to the PP to very

strong dispersion in both band edges for rotations of both parts by 110. We believe that such

rotations can be achieved by proper selection of functional groups (steric control units, SCUs,

see below) in the PP linkers.

5.4 Conclusions

This Chapter shows that, in addition to molecular functional groups, undercoordinated metal

sites, porosity, and large surface areas, a further possibility of property control can be

incorporated into crystalline framework materials, e.g. SURMOFs: If aromatic molecules are

placed in well-controlled stacks, the proximity effect gives raise to strong electronic effects. If

the intermolecular distance between the basal planes of the aromatic molecules is in the range

of the interlayer distance of graphene (≈3-3.5 Å), disperse electronic bands emerge, resulting

in a ballistic charge carrier transport with appreciable mobilities. Thus, while the electronic

properties of most framework materials are merely the superposition of the electronic

properties of the constituting molecular building blocks, a suitable stacking of aromatic

Page 93: Theoretical Investigations of the Photophysical Properties

87

building blocks can turn them into semiconducting materials with particular electronic and

optoelectronic properties.

Without steric control and sufficient flexibility, van-der-Waals interactions result in self-

assembly of PP stacks with strong proximity effect. Mutual shift and twist between the basal

planes of the aromatic PP linkers and intermolecular distance have a strong impact on the

resulting electronic structure change, due to the proximity effect. It is possible to control the

stacking by strong interlayer interactions, functional groups (or SCUs), and the introduction of

SCUs at the proper positions in the PP linkers. These control mechanisms are still beyond the

state-of-the-art and subject of the author’s ongoing research efforts.

Page 94: Theoretical Investigations of the Photophysical Properties

88

Page 95: Theoretical Investigations of the Photophysical Properties

89

Chapter 6

Computational Screening of Phthalocyanine-based

Surface Mounted Metal-Organic Framework

"Somewhere, something incredible is waiting to be known."

– Carl Sagan

This Chapter contains work that has not been published yet

Page 96: Theoretical Investigations of the Photophysical Properties

90

This Chapter investigates the photophysical properties of phthalocyanine-based surface-

mounted metal organic framework (PC-SURMOF) thin films. We are interested in exploring

PCs as alternative SURMOF building blocks, because they are not only more stable, but also

possess enhanced absorption in the visible and the near IR spectral regions in comparison to

PPs. Hence, the exploitation of PCs could enrich the library of building blocks for crystalline

materials with controlled optical properties.

In Section 6.1, after a brief introduction to PCs followed by their pros and cons over PPs, we

introduce the benefits of crystalline systems and an approach based on the regular assembly

of PC-linkers into thin-films of MOFs. Then we briefly summarize the state-of-the-art

concerning PCs integration into MOFs and the scope of layered PC-SURMOF. Finally, we

estimate the absorption spectra for a series of multi-functionalized PCs by using the efficient

computational protocol as identified in Chapter 3.

Section 6.2 lays out a summary of the computational methodologies utilized in this study.

Besides, we briefly present the series of PC-derivatives having a diverse extent of π-

conjugation, ring functionalization, as well as inclusion of a central metal atom.

In Section 6.3, first, we discuss the predicted absorption spectra for a series of PCs. Next, we

select exemplary PCs exhibiting improvement in the characteristic Q-bands and thus utilize

them for assembling in a SURMOF structure. Finally, we investigate the impact of the

arrangement of PC-linkers in the theoretically modelled PC-SURMOF by calculating its

electronic band structure.

Finally, in Section 6.4, we summarize our preliminary findings while demonstrating the power

of computational screening protocols to identify the promising PC-linkers for the construction

of layered PC-SURMOFs with the desired photophysical properties.

Page 97: Theoretical Investigations of the Photophysical Properties

91

6.1 Introduction

Phthalocyanines (PCs) and their derivatives are a group of heterocyclic macrocycles, best-

known as synthetic porphyrin (PP) analogues,285-286 consisting of four iso-indole units linked

together through nitrogen atoms. PCs possess an 18π-electron aromatic cloud delocalized over

an arrangement of alternating carbon and nitrogen atoms as shown in Figure 6.1. In a free base

PC, the two hydrogen atoms can be replaced by different central metal atoms and a variety of

substituents can be incorporated, both at the periphery of macrocycle and the axial positions,

thus allowing fine-tuning of the physical responses. PCs and their derivatives are chemically

and thermally stable, which makes them promising candidates to be integrated into devices

such as solar cells,287 sensors,288 transistors289 etc. Moreover, they have recently attracted an

increasing interest as a building blocks for the construction of new molecular materials that

give rise to promising photoelectric and photophysical properties.290-291

Like in porphyrinoids, the light harvesting properties of PCs and their derivatives constitute

one of their most fascinating attributes, which can be probed by UV-Vis spectroscopy. Also,

in this case, we can follow the 18-π cyclic polyene model, as suggested by Gouterman13 to

explain their UV-Vis spectrum. In PCs, the formal introduction of heteroatoms (nitrogen) at

Figure 6.1 The unsubstituted phthalocyanine macrocycle

Page 98: Theoretical Investigations of the Photophysical Properties

92

the four meso positions, combined with four fused benzo rings significantly breaks the

degeneracy of the frontier orbitals, causes a progressive red-shift of the Q-band to the region

of 670 nm, and a blue-shift of the Soret-band, appearing at around 300 nm. Furthermore, the

strong mixing between the Soret and Q-band transitions is reduced. As a result, the weakly

intense Q-bands acquire significant intensity. To sum up, PCs are not only owing structural

robustness and stability, but also possess red-shifted Q-bands with a significantly enhanced

absorption intensity in the visible and near IR spectral regions as compared to PPs (Figure 6.2).

In contrast to above-mentioned advantages of PCs over PPs, it is worth to highlight that the

synthesis of PCs is more exhausting and challenging as compared to PPs. This is reflected by

the rough numbers of successful PCs syntheses, which amounts to over 100,000 for PPs

(according to Sci-Finder), but only to 20,000 for PCs (both including metalation reactions).

Given all the facts and figures, PCs are among the best-performing organic molecules regarding

light harvesting properties and perfectly suitable for their integration in solar energy conversion

devices e.g. organic photovoltaics (OPVs). Thin films of PCs have been playing an important

role in incorporating them into devices. Generally, in the field of molecular electronics, ordered

(crystalline) PC materials are preferred over disordered (amorphous) PC materials, which not

Figure 6.2 Comparison of calculated UV-Vis spectrum of PP-1H and PC-1H molecules

Page 99: Theoretical Investigations of the Photophysical Properties

93

only facilitates their characterization and theoretical analysis, but also anticipate to

significantly enhance the device performance. Several previous investigations with different

approaches292-296 (e.g. sublimation, Langmuir-Blodgett [LB] etc.) have been performed with

the goal to develop well-defined, thin PC-layers on various substrates. The LB approach has

proved to be particularly useful for the fabrication of organized thin films of PCs.297 However,

the resulting systems do not exhibit a highest degree of ordering and orientation in the thin

films. In the recent work (see Chapter 4), we have demonstrated that PP-based SURMOFs

outperform the other type of PP-based thin films17-19 (made, e.g., using vapor phase deposition,

self-assembly etc.) as regards structural order and optical quality. The SURMOF approach is

well suited to grow crystalline, highly oriented, monolithic thin films which greatly simplify

the integration of chromophoric MOF materials into devices. Therefore, in this Chapter, we

propose to extend the successful fabrication of PP-SURMOFs to PC-SURMOFs. This will

enable the extension of MOF application areas to photo-electrochemistry, where the conditions

are much harsher.

As per state of the art, only very few articles reporting the integration of PCs into MOFs have

been published to date.54-56 Basically, all of these studies are linked to tetra-topic PCs

functionalized with -NH2/-OH groups. Reducing the connectivity to ditopic PCs can give more

structural flexibility, and variation. However, studies where ditopic PC-linkers were employed

for assembly of MOFs are not known as per our knowledge. The same is applicable for thin

films of PC-MOF, no prior literature was found on this area. The fact that works on PC-related

MOFs is limited results from the rather extensive synthesis effort need to synthesize PCs, and

in particular functionalized PCs. To circumvent the huge synthesis efforts and resources, we

estimate the absorption spectra for a series of multi-functionalized PCs by using the efficient

computational protocol as identified in Chapter 3. Only the most promising PC candidates will

then be synthesized and used to fabricate SURMOFs. Depending on the PC substitution pattern

Page 100: Theoretical Investigations of the Photophysical Properties

94

and/or the choice of PC metal reaction, different intermolecular interactions can be supported,

reduced, or even blocked, thus affecting their electrical and optical properties.

6.2 Computational Methodologies

For molecular systems: All the geometries have been fully optimized by using BLYP

functional164 in combination with the TZVP basis set.170 The energy convergence criterion of

the self-consistent field cycle was set to 10−8 Hartree and dispersion forces were considered by

the 3rd version of Grimme’s empirical dispersion correction D3 in combination with the

improved Becke–Johnson damping.166, 244 To speed up the DFT calculations, the resolution of

identity approximation167-168, 245(RI) was employed for the Coulomb integrals, while exchange

contributions were accelerated by the “chain of spheres” approximation191 (COSX). After

structural optimizations, excited states were calculated by the simplified Tamm-Dancoff

approximation (sTDA).106 in combination with CAM-B3LYP functional183 and the def2-SVP

basis set.188 All calculations were carried out employing the ORCA 4.1.1 program package.129

For solid-state system: The structure models have been created using AuToGraFS.125 The

frameworks structure and lattice parameters of phthalocyanine based SURMOFs were pre-

optimized using Universal Force Field (UFF)126 employing UFF4MOF parameters.246 The

bond order specified between paddlewheel Zn atoms was set to 0.25, paddlewheel Zn-O bonds

were set to 0.50, all other bond orders were specified according to the standard chemical

notation. Frameworks geometry optimizations have been carried out using the self-consistent

charge, density functional based tight-binding (SCC-DFTB) method including UFF-dispersion

along with DFTB.org/3ob-3-1 parameters. Single-point calculations for band gaps and band

structures were performed using CRYSTAL17130 program along with DFT PBE173-174

functional, pob-TZVP247 basis type and 100 k-points in the Monkhorst-Pack mesh.

Page 101: Theoretical Investigations of the Photophysical Properties

95

PC-derivatives set: We have included a series of PC-derivatives having a variety of

substituents, both at the periphery and the axial positions of the PC macrocycle: (i) ring

functionalization with electron donating and electron withdrawing groups; (ii) extension of π-

conjugation by including phenyl and phenylacetylene groups (iii) inclusion of a central metal

atom. The molecules include in our PC-derivative set are given in Figure 6.3

6.3 Results and Discussions

In this section, the UV-Vis spectrum of multi-functionalized PC-derivatives 1-10 is presented.

We will mainly discuss the positions and intensities of the characteristic Q-bands of PC-

derivatives in the following because they absorb light in the visible to near IR region of solar

spectrum. We would like to add here that the simplified time-dependent studies involve only

transitions in the frozen ground state of molecules. Furthermore, these transitions can be clearly

Figure 6.3 Molecular structures of diverse phthalocyanine derivatives [1-10]

Page 102: Theoretical Investigations of the Photophysical Properties

96

distinguished from other transitions in the excited state calculations. Next, we investigate the

impact of the arrangement of PC-linkers in the theoretically optimized PC-SURMOF structure

by calculating and analyzing its electronic band structure diagram.

The precursors for the synthesis of the PC can be modified through different substitution

pattern at the periphery of macrocycle and the axial positions in silico. Numerous possible

modifications of the molecular structure permit the fine-tuning of physical, electronic, and

optical responses. The peripheral substitution of PC-derivatives plays a significant role not only

in the fine-tuning of their characteristic absorption bands, but also improves their solubility and

permits the depositions onto substrates. The attached peripheral substituents can be divided

into electron donating or electronic withdrawing groups. The incorporation of metal at the

central cavity affects the electronic structure in a way that as thermodynamically stable

delocalized dianion with higher symmetry is obtained. The extension of π-conjugated system

Figure 6.4 Comparison of calculated UV-Vis-spectrum for phthalocyanine-derivatives [1-10]

Page 103: Theoretical Investigations of the Photophysical Properties

97

(e.g. phenyl or phenylacetylene unit) offers another possibility to influence the absorption

properties which is present in the all the structures included in our set as shown in Figure 6.3.

Table 6.1 Calculated absorption wavelengths of PC-derivatives with employed abbreviations.

To examine the influence of attached substituents with respect to light absorbing properties of

PCs, we first investigate the simple free base and metal-containing PC-derivatives abbreviated

as 1 and 2, respectively. It can be deduced from Table 6.1 and Figure 6.4, the calculated Q-

bands values appears in the visible and near IR region. Moreover, an incremental red-shift in

the Qy-band through Zn-metal implementation is observed in the visible region (2 in Figure

6.4). Next, we examine the effect of four large bromine groups (replace the hydrogen-atoms in

1) at the β-positions of free base PC-derivatives abbreviated as 3. It can be deduced from Table

6.1, structure 3 exhibits a Qx-band ~30 nm at longer wavelength with respect to the structure

1 in the near IR region while there is no improvement of Qy-band in the visible region.

Likewise, 1 and 3, the similar trend is observed for Q-bands in the Zn-metal containing

Structures Abbrev.

Qx Band (nm)

Qy Band (nm)

1 722 665

2 703 671

3 750 667

4 732 668

5 723 668

6 751 686

7 728 694

8 722 670

9 750 685

10 727 695

Page 104: Theoretical Investigations of the Photophysical Properties

98

structures 2 and 4, while a somewhat pronounced variation in their band width (see Table 6.1

and Figure 6.4).

Apart from ring functionalization effect at β-positions, we examine the impact of introduced

functionality (propyl groups) at the extent of π-conjugated structure in the form of cis and trans

conformers, abbreviated as 5 and 6 respectively. It can be deduced from Table 6.1, the

estimated Qx-band in near IR region for structures 5 and 6 are quite similar to structures 1 and

3 respectively, while the structure 6 outperforms 1 and 3 exhibiting a Qy-band ~20 nm at longer

wavelength in the visible region (see Table 6.1). To compare 5 and 6 conformers, 6 displays

progressive red shift in the Q-bands. Moreover, in the structure 6, an increased absorption and

incremental red-shift in Qy-band through Zn-metal implementation is observed in the visible

region (7 in Figure 6.4). Likewise 5-7, the similar trend is observed for the structures 8, 9 and

10 as well by extending the π- conjugated system with acetylene group forming cis/trans and

employing the Zn-metal to the central cavity of trans conformer (see Table 6.1 and Figure 6.4).

On the basis of computational screening of series of PC-derivatives, we have initially started

with an exemplary and simplest PC-derivative abbreviated as 1, exhibiting improvement in the

characteristic Q-bands as compared to PP/PC-derivatives and thus utilized for assembling in a

SURMOF structure. Based on our previous experience of PP-based SURMOF, we have

successfully optimized the geometry and lattice constants for the theoretically modeled PC-

SURMOF (see Figure 6.5a). The optimized lattice constants (a, b, and c) calculated for the

periodic framework are given in Table 6.2. Next, we have investigated the impact of the

arrangement of PC-linkers by calculating its band structure and observed significant band

dispersion along the Z direction of the Brillouin zone (see Figure 6.5b-c) which is responsible,

both for the bathochromic shift (red-shift) of the absorption band, as well as for the charge

carrier mobility along the PC-stacks. We have observed a formation of direct band gap with

Page 105: Theoretical Investigations of the Photophysical Properties

99

significant band dispersion ≈280 meV (in the conduction and the valence band), which is larger

than that for (≈200 meV) reported in previous study (see Chapter 4) for PP-based SURMOFs

Table 6.2 Calculated lattice parameters for the exemplary PC-based SURMOF structure

Structures a (in Å) b (in Å) c (in Å)

PC-SURMOF 34.26 34.26 5.14

6.4 Conclusion

In conclusion, we present here an attractive route to model chromophoric assemblies by

SURMOF-based approach and demonstrate the power of computational screening methods to

identify the promising PC-linker for the construction of layered PC-SURMOFs with the desired

photophysical properties. The starting point is to set a candidate structure, in the case ditopic

PC-derivative with variant substitution pattern are optimized in silico. Based on the predictive

absorption properties, we select an exemplary PC-derivative exhibiting enhanced Q-bands

Figure 6.5 a) Top and side view of calculated atomistic PC-SURMOF structure; d) Brillouin Zone;

c) Electronic band structure diagram of PC-SURMOF structure.

Page 106: Theoretical Investigations of the Photophysical Properties

100

absorption in the visible/near-IR region and thus utilized for modeling the SURMOF structure.

Based on successfully optimized geometry, we have calculated its electronic band structure

and observed a direct band gap with significant dispersion ≈280 meV, which larger than that

(≈200 meV) reported in previous work (see Chapter 4) for PP-SURMOF / (5´) structure.

Altogether, our preliminary findings demonstrate that photophysical properties of PC-based

SURMOF offers a huge potential for opto-electronic devices with covering the entire solar

spectrum, ranging from ultraviolet to infrared.

Page 107: Theoretical Investigations of the Photophysical Properties

101

Chapter 7

Summary

"Science cannot solve the ultimate mystery of nature. And that is because, in the last analysis, we

ourselves are a part of the mystery that we are trying to solve." – Max Planck

For inorganic semiconductors such as

silicon, crystalline order leads to bands

in the electronic structure which give

rise to drastic differences with respect to

disordered materials. Distinct band

features lead to photo-effect, and the

band structure can be tuned to optimize

the performance of the photovoltaic

(PV) device. An example is the presence of an indirect band gap. For organic semiconductors,

such effects are typically precluded, since most organic materials employed are disordered,

which hampers their characterization and theoretical analysis. The inspiration for this thesis

came from the very first evidence of an indirect band gap exhibited by highly ordered and

crystalline porphyrin-based surface-mounted metal-organic framework (Pd-PP-based

SURMOF) material.72 The presence of an indirect band gap should in principle result in

suppressed charge recombination and efficient charge separations which would significantly

enhance the PV device performance. However, the energy gain from the electronic band

dispersion in the reported Pd-PP-based SURMOF is far too low (≈5 meV) and results in a very

Page 108: Theoretical Investigations of the Photophysical Properties

102

low photocurrent generation (efficiency 0.2%), which is certainly not sufficient for the

application. Another noticeable shortcoming is the weakly absorbing Q-bands of the employed

PP chromophore (Pd-metal containing porphyrinoid, Pd-PP) in the visible region of the solar

spectrum. Nevertheless, this novel research has highlighted the potential to improve the

photophysical properties of PP-SURMOFs by (i) introducing various functional groups or

metal ions to the PP-core and (ii) controlling the PP-stacking behavior in layered materials.

To overcome the posed shortcomings of the PP-MOF prototype PV material and to exploit the

potential of PP-SURMOFs, we have employed the following approach to increase the light

absorption and the electronic band dispersion.

Firstly, we proposed a computationally feasible protocol to investigate the light

absorption properties of PP derivatives or related PP-containing materials.

Secondly, we predicted the light absorption properties of multi-functionalized PPs (i.e.

tuning the weakly absorbing Q-bands) by using a validated computational protocol,

thus allowing us to identify different PP linkers with different light absorption

properties, allowing to bridge the so-called green gap.

Finally, we incorporated the most promising PP linkers for the construction of layered

SURMOF materials and optimized the PP-stacking behavior to achieve the desired

photophysical properties.

Besides PPs, we have extended our investigations to phthalocyanines (PCs) as alternative

individual SURMOF building blocks, because they do not only exhibit structural robustness

and stability but also possess enhanced absorption in the visible and the near IR spectral regions

in comparison to PPs. Hence, the exploitation of PCs could enrich the library of SURMOF

materials with the desired optical quality.

Page 109: Theoretical Investigations of the Photophysical Properties

103

In line with these considerations, the objective of this thesis is two-fold. We first develop a

computationally feasible protocol to investigate the absorption properties of chromophores or

extended biological systems. Our second aim is to assemble the most promising chromophores

for the construction of layered SURMOF materials with the desired photophysical properties

such as increased electronic band dispersion or photocurrent.

Chapter 1 introduced the research topic and systems of interest. In Chapter 2, we discussed

theory and methods, as well as the computational protocols that have been applied to

investigate the posed problems and achieve the respective goals are briefly summarized.

In Chapter 3, we presented a detailed

validation of various variants of time-

dependent density-functional theory

(TD-DFT) for predicting the UV-Vis

spectra of PP derivatives with diverse extent of π-conjugation, ring functionalization, as well

as inclusion or modification of a central metal atom. With the aim to provide an approach that

is computationally feasible for large-scale applications such as molecular framework materials,

we have assessed the performance of the simplified Tamm-Dancoff approximation (sTDA),

sTD-DFT, and canonical TD-DFT (including TDA). We have compared the results given by

various computational protocols by exploiting various basis sets and density-functionals

(gradient corrected local functionals, hybrids, range separated, and double hybrids) with

respect to the experimental references. An excellent choice for these calculations is the range-

separated functional CAM-B3LYP in combination with sTDA and a basis set of double-ζ

quality (Mean Absolute Error [MAE] ≈0.05 eV). This is not surpassed by more expensive

approaches, not even by double hybrid functionals, and solely systematic excitation energy

scaling slightly improves the results (MAE ≈0.04 eV). Unfortunately, none of the gradient

Page 110: Theoretical Investigations of the Photophysical Properties

104

corrected local functionals have sufficient predictive power (for a detailed discussion see

Chapter 3), which is an obstacle especially for periodic calculations.

In Chapter 4, by using the simplified

time-dependent approach identified

in Chapter 3 as the most cost-efficient

yet reliable method for accurate

prediction of the absorption spectra

of PPs, we have examined how the molecular and electronic structure of variously substituted

PPs can be tuned for increasing their absorption efficiency. Based on this computational

screening, three interesting strategies for tuning the structure and light-absorbing properties of

PPs have been identified:

(i) Distortion of the planarity of the PP core,

(ii) Extension of π-conjugation by adding a phenyl-acetylene substituent, and

(iii) Introduction of electron‐withdrawing functional groups.

These structural modifications cause pronounced changes in the positions and intensities of the

Q-bands of the respective PP molecules. For applications, the assembly of PP linkers into thin

films of high optical quality is crucial. In a final step, the identified PP linkers were assembled

into SURMOFs. To rationalize the experimental UV-Vis spectra of the synthesized PP‐based

SURMOFs, we have estimated the impact of the arrangement, also known as stacking, of the

PP linkers in the framework structures by calculating the respective electronic band structures

(for a detailed discussion see Chapter 4). We have calculated an indirect band gap dispersion

of ≈200 meV that is 40 times larger than the previously reported value of ≈5 meV for a Pd-PP‐

SURMOF.72 In conclusion, our study has demonstrated the power of computational screening

Page 111: Theoretical Investigations of the Photophysical Properties

105

methods to identify the most promising PP derivatives for the construction of layered PP-

SURMOF materials with the desired photophysical properties.

In Chapter 5, we have provided in-

depth analysis of the proximity effect

and stacking control in the PP-

containing SURMOFs modeled in

Chapter 4. For one of the most

promising PP-based SURMOFs i.e. 5´, we have studied the band dispersion as a function of

the rotation of the introduced functional groups, such as phenyl and phenyl-acetylene, with

respect to the PP-core in a bulk framework (for a detailed discussion see Chapter 5). The

different degree of structural rotation causes different structure of the electronic bands - from

an almost flat conduction band in a hypothetical structure with all functional groups in-plane

with the PP core, to a pronounced dispersion of both band edges for a structure with out-of-

plane rotated (by 110) substituents. We believe that the desired molecular and electronic

structure of the PP-based SURMOFs can be achieved by a careful selection of the functional

groups and their introduction at the proper positions in the PP-linkers.

In Chapter 6, we presented theoretical

findings based on the screening of a

series of PC linker molecules and an

exemplary PC-based SURMOF. By

utilizing the simplified time-dependent

approach identified in Chapter 3, we

predicted the UV-Vis spectra of various

PC derivatives having diverse extent of π-conjugation, ring functionalization, inclusion of a

Page 112: Theoretical Investigations of the Photophysical Properties

106

central metal atom, as well as conformer variants. The structural modifications cause

pronounced changes in the positions and intensities of the absorption bands of PC molecules,

covering the entire range from violet to the near infrared, including the green gap (for a detailed

discussion see Chapter 6). Based on the estimated light absorption properties of PC derivatives,

an exemplary PC derivative exhibiting enhanced Q-band absorption is thus utilized for

modeling the SURMOF structure. Next based on fully optimized geometry and lattice

parameters, we have calculated the electronic band structure and observed a direct band gap

with substantial electronic band dispersion of ≈280 meV. Altogether, our results demonstrate

that the photophysical properties of PC-based SURMOFs present a huge prospective for the

construction of optical devices.

By leveraging the light-harvesting properties of the chromophores, one can rationally design

novel chromophore-based SURMOFs. On one end, we have demonstrated from the electronic

structure and properties calculations (validated with experiments), that tuned and well-

controlled arrays of PP chromophores in layered SURMOFs display enhanced photophysical

properties compared to the previously reported Pd-PP-SURMOF. The substantial indirect band

dispersion of the PP-SURMOF i.e. (5´) reported in our work is ≈200 meV, which is 40 times

higher than the previously reported72 value of ≈5 meV (see illustrating Figure 7.1). On the other

end, our screening of an exemplary PC-based SURMOF results in a direct bandgap displaying

substantial band dispersion of ≈280 meV. Altogether, our results demonstrate that the solid-

state properties of chromophoric MOFs offer a huge potential for optoelectronic devices.

Page 113: Theoretical Investigations of the Photophysical Properties

107

Figure 7.1. Left (a) Building block (Pd-porphyrinoid, Pd-PP) together with the top and side views

of Pd-PP-based Zn-SURMOF, and the corresponding band structure adapted from Ref. [72]. Bands

are fairly flat, and the dispersion is in the limit of ≈5 meV; Center (b) Building blocks together with

the top and side views of three Zn-SURMOFs assembled from multi-functionalized PP linkers: (i)

1-Br ', (ii) 5 ' and (iii) 10 '; Right (c) Strong band dispersion (≈200 meV) observed for the 5 ' in the

stacking direction, due to enhanced London dispersion interactions between the PP linkers.

Page 114: Theoretical Investigations of the Photophysical Properties

108

Page 115: Theoretical Investigations of the Photophysical Properties

109

A. Acronyms

2D Two Dimensional

3D Three Dimensional

ADC2 Second Order Algebraic Diagrammatic Construction

ADF Amsterdam Density Functional

ALDA Adiabatic Local Density Approximation

AMS Amsterdam Modeling Suite

AuToGraFS Automatic Topological Generator for Framework Structure

BO Born-Oppenheimer

BS Band Structure

CASPT2 Complete Active Space Second Order Perturbation Theory

CB Conduction Band

CC Coupled Cluster

CD Circular Dichroism

CI Configuration Interaction

COF Covalent-Organic Framework

COSX Chain of Spheres Approximation

CT Charge Transfer

DA-ZnP Zinc [5,15-di(4-pyridylacetyl)-10,20-diphenyl] porphyrin

DFT Density Functional Theory

DFTB Density Functional Tight Binding

DPA Diphenylamine

F-ZnP Zinc [5,15-dipyridyl-10,20-bis(pentafluorophenyl) porphyrin

GGA Generalized Gradient Approximation

GH Global Hybrid

H2OBP Octabromotetraphenyl porphyrin

H2OEP Octaethylporphyrin

H2TAPP Tetrakis(o-aminophenyl) porphyrin

H2TPP Tetraphenylporphyrin

HEG Homogeneous Electron Gas

HF Hartree-Fock

HK Hohenberg-Kohn

HOMO Highest Occupied Molecular Orbital

KS Kohn-Sham

LB Langmuir Blodgett

LbL Layer-by-layer

LC Long-range Corrected

LCAO Linear Combination of Atomic Orbitals

LDA Local Density Approximation

LPE Liquid Phase Epitaxial

LR Linear Response

LUMO Lowest Unoccupied Molecular Orbital

Page 116: Theoretical Investigations of the Photophysical Properties

110

MAE Mean Absolute Error

MAXE Absolute Maximum Error

ME Mean Error

mGGA meta Generalized Gradient Approximation

MgOEP Magnesium Octaethylporphyrin

MgTPP Magnesium Tetraphenylporphyrin

MOF Metal-Organic Framework

MP2 Second Order Møller–Plesset Perturbation Theory

M-PP Metal-Porphyrin

NAFS-1 Nanofilm of MOFs on Surface-No.1

NAFS-2 Nanofilm of MOFs on Surface-No.2

NEVPT2 Second Order N-electron Valence State Perturbation Theory

OPV Organic Photovoltaic

PBC Periodic Boundary Condition

PC Phthalocyanine

PCP Porous Coordination Polymer

Ph Phenyl

PhA Phenyl Acetylene

PP Porphyrin

PV Photovoltaic

RG Runge-Gross

RI Resolution of Identity

RPA Random Phase Approximation

RSH Range Separated Hybrid

SAC-CI Symmetry Adapted Cluster Configuration Interaction

SBU Secondary Binding Unit

SCC-DFTB Self Consistent Charge Density Functional Tight Binding

SCU Steric Control Unit

SE Schrödinger Equation

SI Supporting Information

SIE Self Interaction Error

sTDA simplified Tamm-Dancoff Approximation

sTD-DFT simplified Time-Dependent Density Functional Theory

STEOM-CC Similarity Transformed Equation of Motion Coupled Cluster

SURMOF Surface-mounted Metal-Organic Framework

TDA Tamm-Dancoff Approximation

TD-DFT Time Dependent Density Functional Theory

UFF Universal Force Field

UV-Vis Ultraviolet-Visible

VB Valence Band

XC Exchange-Correlation

XRD X-Ray Diffraction

ZnOEP Zinc Octaethylporphyrin

ZnTCPP Zinc tetrakis(4-carboxyphenyl) porphyrin

ZnTPP Zinc Tetraphenylporphyrin

Page 117: Theoretical Investigations of the Photophysical Properties

111

B. Appendices

B1. Supporting Information (SI) of Chapter 3 (S3)

Benchmarking the Performance of Time-Dependent Density

Functional Theory for Predicting the UV-Vis Spectral Properties of

Porphyrinoids

Table of Contents

I. UV-Vis Representation of Tetra-Phenyl-Porphyrin (H2TPP)

II. Computational Summary

III. Charge Transfer Excitations in Tetrakis(o-aminophenyl) Porphyrin

IV. Scaled Error Dataset for Diverse DFT Functional Approaches

V. Boxplot – Absolute Mean Error Representations

VI. Influence of Diffuse Functions and Geometry

Page 118: Theoretical Investigations of the Photophysical Properties

112

I. UV-Vis Representation of Tetraphenyl Porphyrin (H2TPP)

II. Computational Summary

Table S3.1 Summarize the computational protocols (including various functionals, approaches

and basis-sets) applied for the PP-benchmark study.

DFT

Group

DFT

Functionals

Computational Approaches (Basis Set Choices)

TDA

(SVP)

RPA

(SVP)

sTDA

(SVP)

sTD-DFT

(SVP)

sTDA

(TZVP)

sTD-DFT

(TZVP)

sTDA

(SVPD)

sTDA

(TZVPD)

GGA and

mGGA

PBE x x x x x x

BP86 x x x x x x

BLYP x x x x x x x x

*BLYP x x x

TPSS x x x x x x

M06-L x x x x x x

PBE0 x x x x x x

Figure S3.1 UV-Vis-spectra of H2TPP and the transitions based on the model of Gouterman’s

Page 119: Theoretical Investigations of the Photophysical Properties

113

Global

Hybrids

B3P86 x x x x x x

B3LYP x x x x x x x x

*B3LYP x x x

TPSS0 x x x x x x

M06 x x x x x x

BHLYP x x x x x x

M06-2X x x x x x x

Range

Separate

Hybrids

ωB97 x x x x x x

ωB97X x x x x x x

LC-BLYP x x x x x x

CAM-B3LYP x x x x x x x x

*CAM-B3LYP x x x

Double

Hybrids

and post

HF

B2PLYP x

B2GP-PLYP x

mPW2PLYP x

CIS x

CIS(D) x

Note: All the ground state geometries are fully optimized using BLYP-D3/TZVP level

Here (*) indicates the re-optimization of all geometries using B3LYP-D3/TZVP level to examine the

influence of applied functional on calculated absorption bands.

III. Charge Transfer Excitations in Tetrakis(o-aminophenyl) Porphyrin

Table S3.2: Charge transfer (CT) properties shown by GGAs in TD-DFT type (TDA and RPA)

for 8 (H2TAPP) molecule. Charge coefficients c yield the weights of the individual excitation,

calculated as 100% 𝑐2.

Page 120: Theoretical Investigations of the Photophysical Properties

114

BLYP/TDA

State 1

E= 1.83 eV

PBE /TDA

State 1

E= 1.81 eV

BP86 /TDA

State 1

E= 1.82 eV

172 177 18 % (c= 0.43) 172 177 15% (c= 0.38) 172 177 15 % (c= 0.39)

175 177 32 % (c= 0.57) 175 177 35% (c= 0.59) 175 177 34 % (c= 0.58)

176 177 43% (c= 0.66) 176 177 46% (c= 0.68) 176 177 55 % (c= 0.67)

BLYP/TDA

State 2

E= 1.86 eV

PBE /TDA

State 2

E= 1.83 eV

BP86 /TDA

State 2

E= 1.84 eV

174 177 99 % (c= 0.99) 174 177 98% (c= 0.99) 174 177 98 % (c= 0.99)

BLYP/RPA

State 1

E= 1.82 eV

PBE /RPA

State 1

E= 1.80 eV

BP86 /RPA

State 1

E= 1.81 eV

172 177 18 % (c= 0.43) 172 177 15 % (c= 0.39) 172 177 16 % (c= 0.40)

175 177 22 % (c= 0.47) 175 177 25 % (c= 0.50) 175 177 24 % (c= 0.49)

176 177 54 % (c= 0.73) 176 177 55 % (c= 0.74) 176 177 55 % (c= 0.74)

BLYP/RPA

State 2

E= 1.86 eV

PBE /RPA

State 2

E= 1.83 eV

BP86 /RPA

State 2

E= 1.84 eV

175 178 99 % (c= 0.99) 174 177 99 % (c= 0.99) 174 177 99 % (c= 0.99)

Figure S3.2 (a) Molecular orbitals of 8 (H2TAPP)

Page 121: Theoretical Investigations of the Photophysical Properties

115

IV. Scaled Error Dataset for Diverse Density Functional Approaches

Table S3.3 Calculated errors in eV for the GGA functionals with scaled absorption energies

GGA and m-GGA

Functionals

TDA

(def2-

SVP)

RPA

(def2-

SVP)

sTDA

(def2-

SVP)

sTD-DFT

(def2-

SVP)

sTDA

(def2-

TZVP)

sTD-DFT

(def2-

TZVP)

PBE

ME 0.01 0.01 0.01 0.01 0.01 0.01

MAE 0.11 0.11 0.10 0.10 0.09 0.09

MAXE 0.49 0.44 0.45 0.41 0.31 0.30

Scaling Factor 0.99 1.02 1.04 1.07 1.09 1.11

BP86

ME 0.01 0.01 0.01 0.01 0.01 0.01

MAE 0.11 0.11 0.10 0.10 0.09 0.09

MAXE 0.48 0.43 0.44 0.40 0.30 0.29

Scaling Factor 0.99 1.02 1.04 1.06 1.09 1.11

BLYP

ME 0.01 0.01 0.01 0.01 0.01 0.01

MAE 0.10 0.11 0.09 0.10 0.09 0.09

MAXE 0.46 0.41 0.41 0.38 0.28 0.27

Figure S3.2 (b) Density difference comparison of H2TAPP with representative functionals

Page 122: Theoretical Investigations of the Photophysical Properties

116

Scaling Factor 1.00 1.03 1.04 1.07 1.10 1.11

TPSS

ME 0.01 0.01 0.01 0.01 0.00 0.00

MAE 0.10 0.10 0.09 0.09 0.09 0.09

MAXE 0.43 0.38 0.39 0.36 0.27 0.26

Scaling Factor 0.97 1.00 1.01 1.03 1.06 1.08

M06-L

ME 0.00 0.00 0.00 0.00 0.00 0.00

MAE 0.09 0.09 0.08 0.08 0.08 0.09

MAXE 0.35 0.30 0.33 0.30 0.24 0.23

Scaling Factor 0.95 0.98 0.98 1.00 1.03 1.04

Table S3.4 Calculated errors in eV for the hybrid functionals with scaled absorption energies.

Global Hybrid

Functionals

TDA

(def2-

SVP)

RPA

(def2-

SVP)

sTDA

(def2-

SVP)

sTD-DFT

(def2-

SVP)

sTDA

(def2-

TZVP)

sTD-DFT

(def2-

TZVP)

PBE0

ME 0.00 0.00 0.00 0.00 0.00 0.00

MAE 0.07 0.08 0.07 0.07 0.08 0.08

MAXE 0.14 0.15 0.17 0.18 0.22 0.22

Scaling Factor 0.90 0.93 1.04 1.06 1.12 1.13

B3P86

ME 0.00 0.00 0.00 0.00 0.00 0.00

MAE 0.07 0.08 0.06 0.07 0.08 0.08

MAXE 0.15 0.16 0.17 0.18 0.22 0.22

Scaling Factor 0.91 0.94 1.01 1.03 1.08 1.09

B3LYP

ME 0.00 0.00 0.00 0.00 0.00 0.00

MAE 0.07 0.08 0.06 0.07 0.08 0.08

MAXE 0.15 0.15 0.17 0.18 0.21 0.22

Scaling Factor 0.92 0.95 1.01 1.03 1.08 1.10

TPSS0

ME 0.00 0.00 0.00 0.00 0.00 0.00

MAE 0.07 0.08 0.06 0.07 0.08 0.08

Page 123: Theoretical Investigations of the Photophysical Properties

117

MAXE 0.14 0.15 0.16 0.17 0.21 0.22

Scaling Factor 0.90 0.93 1.02 1.04 1.09 1.11

M06

ME 0.00 0.00 0.00 0.00 0.00 0.00

MAE 0.07 0.08 0.06 0.07 0.08 0.08

MAXE 0.14 0.15 0.17 0.18 0.21 0.22

Scaling Factor 0.93 0.97 1.08 1.11 1.17 1.19

BHLYP ME 0.00 0.00 0.00 0.01 0.00 0.00

MAE 0.07 0.07 0.06 0.06 0.08 0.08

MAXE 0.14 0.14 0.14 0.14 0.18 0.16

Scaling Factor 0.88 0.94 1.17 1.21 1.27 1.34

M06-2X ME 0.00 0.00 0.00 0.01 0.00 0.00

MAE 0.07 0.06 0.06 0.06 0.08 0.08

MAXE 0.13 0.12 0.14 0.15 0.19 0.18

Scaling Factor 0.88 0.93 1.24 1.31 1.35 1.47

Table S3.5 Calculated errors in eV for the RSH functionals with scaled absorption energies.

Range Separated

Hybrid Functionals

TDA

(def2-

SVP)

RPA

(def2-

SVP)

sTDA

(def2-

SVP)

sTD-DFT

(def2-

SVP)

sTDA

(def2-

TZVP)

sTD-DFT

(def2-

TZVP)

ωB97

ME 0.00 0.00 0.00 0.00 0.00 0.00

MAE 0.06 0.05 0.12 0.12 0.11 0.11

MAXE 0.12 0.11 0.23 0.23 0.22 0.23

Scaling Factor 0.91 1.08 1.08 1.28 1.09 1.27

ωB97X

ME 0.00 0.00 0.00 0.00 0.00 0.00

MAE 0.06 0.05 0.12 0.12 0.11 0.11

MAXE 0.12 0.10 0.23 0.24 0.21 0.22

Scaling Factor 0.90 1.04 1.09 1.26 1.10 1.25

LC-BLYP ME 0.00 0.00 0.00 0.00 0.00 0.00

Page 124: Theoretical Investigations of the Photophysical Properties

118

MAE 0.06 0.05 0.12 0.12 0.11 0.11

MAXE 0.12 0.10 0.24 0.24 0.22 0.22

Scaling Factor 0.91 1.05 1.07 1.20 1.08 1.20

CAM-

B3LYP

ME 0.00 0.00 0.00 0.00 0.00 0.00

MAE 0.06 0.05 0.04 0.05 0.06 0.07

MAXE 0.13 0.12 0.08 0.12 0.13 0.13

Scaling Factor 0.90 0.96 0.99 1.01 1.07 1.09

Table S3.6 Calculated errors in eV for the double hybrids and post-Hartree Fock (HF) methods

with scaled absorption energies

Double Hybrids and

Post-HF Methods (def2-SVP)

B2PLYP

ME 0.00

MAE 0.04

MAXE 0.08

Scaling Factor 0.90

B2GP-PLYP

ME 0.00

MAE 0.04

MAXE 0.09

Scaling Factor 0.88

mPW2PLYP

ME 0.00

MAE 0.05

MAXE 0.09

Scaling Factor 0.89

CIS

ME 0.00

MAE 0.10

MAXE 0.17

Page 125: Theoretical Investigations of the Photophysical Properties

119

Scaling Factor 0.89

CIS (D) ME 0.00

MAE 0.04

MAXE 0.08

Scaling Factor 0.80

V. Box-Plots Analysis - Absolute Mean Error Representations

Box-plot of variant DFT functionals displaying the error distribution of dataset based on the

following four number summary: (+) mean absolute error; (box) range between lower and

upper quartile; (dashed lines) minimum and maximum, excluding outliers; (black dots) outliers

termed as extremes error values which are outside the range given in bar. Variant color box

represents functional-approach-basis-set combination such as yellow: TDA (def2-SVP); cyan:

RPA (def2-SVP); orange: sTDA (def2-SVP); sky blue: sTD-DFT (def2-SVP); red: sTDA

(def2-TZVP); blue: sTD-DFT (def2-TZVP); white: TDA+MP2 (def2-SVP); grey: CIS and

CISD, (def2-SVP); light brown: sTDA (def2-SVPD) and dark brown: sTDA (def2-TZVPD)

Page 126: Theoretical Investigations of the Photophysical Properties

120

Figure S3.3 Boxplot displaying scaled error values in eV for the variant of density-functionals in

combination with TD-DFT types (a) TDA, (b) RPA, and def2-SVP basis set.

Figure S3.4 Boxplot displaying scaled error values in eV for the variant of density-functionals in

combination with TD-DFT types (a) sTDA, (b) sTD-DFT, and def2-SVP basis set.

Page 127: Theoretical Investigations of the Photophysical Properties

121

Figure S3.5 Full box-plot error distribution in eV for all approaches and basis set combinations.

Original error values are on the left while scaled error values are shown on the right.

Page 128: Theoretical Investigations of the Photophysical Properties

122

Figure S3.6 Full box-plot error distribution in nm for all approaches and basis set combinations.

Original error values are on the left while scaled error values are shown on the right.

Page 129: Theoretical Investigations of the Photophysical Properties

123

VI. Influence of Diffuse Functions and Geometry

Table S3.7 Calculated original and scaled error values in eV for the selected density functionals

BLYP-D3/TZVP

Optimized Geometry

Original Error Values Scaled Error Values

sTDA-

def2-

SVP

sTDA-

def2-

SVPD

sTDA-

def2-

TZVPD

sTDA-

def2-

SVP

sTDA-

def2-

SVPD

sTDA-

def2-

TZVPD

BLYP

ME -0.08 -0.23 -0.26 0.01 0.00 0.00

MAE 0.09 0.23 0.26 0.09 0.09 0.09

MAXE 0.49 0.50 0.51 0.41 0.27 0.27

Scaling Factor - - - 1.04 1.12 1.14

ME -0.03 -0.22 -0.25 0.00 0.00 0.00

B3LYP MAE 0.06 0.22 0.25 0.06 0.09 0.09

MAXE 0.18 0.42 0.45 0.17 0.27 0.24

Scaling Factor - - - 1.01 1.12 1.14

ME 0.03 -0.20 -0.23 0.00 0.00 0.00

CAM-

B3LYP

MAE 0.05 0.20 0.23 0.04 0.08 0.09

MAXE 0.10 0.35 0.14 0.08 0.15 0.15

Scaling Factor - - - 0.99 1.10 1.12

Page 130: Theoretical Investigations of the Photophysical Properties

124

Figure S3.7 Boxplot displaying original error values in eV for the selected density functionals with

and without diffuse (‘D’) basis set functions where (a) is the BLYP-D3-TZVP based geometry

optimization while (b) is B3LYP-D3/-TZVP based geometry optimization for the benchmark-set

Figure S3.8 Box-plot displaying scaled error values in eV for the selected density functionals with

and without diffuse(‘D’) basis set functions, where (a) is the BLYP-D3-TZVP based geometry

optimization while (b) is B3LYP-D3-TZVP based geometry optimization for the benchmark-set.

Page 131: Theoretical Investigations of the Photophysical Properties

125

Table S3.8: ϵ (HOMO) criterion based on BLYP-D3-TZVP geometries

BLYP-sTDA B3LYP-sTDA CAM-B3LYP-sTDA

Molecules ϵ

(eV)

def2-

SVP

def2-

SVPD

def2-

TZVPD

def2-

SVP

def2-

SVPD

def2-

TZVPD

def2-

SVP

def2-

SVPD

def2-

TZVPD

H2PP ϵH-L 1.90 1.92 1.92 2.87 2.88 2.88 4.68 4.68 4.68

- ϵH 2.74 2.91 2.91 2.43 2.58 2.57 1.69 1.85 1.84

H2OEP ϵH-L 1.94 1.96 1.96 2.91 2.88 2.86 4.60 4.56 4.53

- ϵH 2.42 2.57 2.55 2.11 2.24 2.21 1.39 1.51 1.49

MgOEP ϵH-L 2.02 2.01 2.00 2.89 2.86 2.85 4.58 4.55 4.52

- ϵH 2.37 2.50 2.48 2.06 2.17 2.14 1.33 1.44 1.41

ZnOEP ϵH-L 2.04 2.02 2.01 2.91 2.88 2.86 4.61 4.56 4.54

- ϵH 2.34 2.51 2.48 2.04 2.17 2.14 1.30 1.44 1.41

H2TPP ϵH-L 1.75 1.75 1.76 2.69 2.69 2.69 4.46 4.46 4.46

- ϵH 2.69 2.87 2.85 2.39 2.54 2.52 1.67 1.82 1.80

MgTPP ϵH-L 1.82 1.80 1.80 2.74 2.72 2.72 4.51 4.48 4.49

- ϵH 2.65 2.80 2.79 2.34 2.48 2.46 1.62 1.76 1.74

ZnTPP ϵH-L 1.88 1.87 1.87 2.81 2.80 2.80 4.59 4.57 4.58

- ϵH 2.63 2.80 2.79 2.32 2.47 2.46 1.59 1.75 1.73

H2TAPP ϵH-L 1.72 1.74 1.75 2.70 2.70 2.71 4.50 4.50 4.51

- ϵH 2.76 2.96 2.94 2.46 2.63 2.61 1.73 1.91 1.88

ZnTCPP ϵH-L 1.87 1.86 1.87 2.80 2.79 2.79 4.58 4.56 4.57

- ϵH 3.00 3.26 3.23 2.72 2.93 2.91 1.98 2.20 2.17

F-ZnP ϵH-L 1.89 1.88 1.88 2.82 2.81 2.82 4.59 4.55 4.53

- ϵH 3.29 3.56 3.51 3.06 3.27 3.21 2.36 2.58 2.52

DA-ZnP ϵH-L 1.42 1.41 1.41 2.25 2.23 2.24 3.91 3.89 3.90

- ϵH 3.33 3.50 3.49 3.09 3.24 3.22 2.43 2.58 2.56

H2OBP ϵH-L 1.36 1.37 1.37 2.21 2.22 2.23 3.90 3.90 3.92

- ϵH 3.31 3.55 3.51 3.12 3.31 3.26 2.47 2.65 2.60

Page 132: Theoretical Investigations of the Photophysical Properties

126

Table S3.9 Calculated original and scaled error values in eV for the selected density functionals

B3LYP-D3/TZVP

Optimized Geometry

Original Error Values Scaled Error Values

sTDA-

def2-

SVP

sTDA-

def2-

SVPD

sTDA-

def2-

TZVPD

sTDA-

def2-

SVP

sTDA-

def2-

SVPD

sTDA-

def2-

TZVPD

BLYP

ME -0.04 -0.19 -0.22 0.00 0.00 0.00

MAE 0.08 0.19 0.22 0.09 0.09 0.10

MAXE 0.46 0.46 0.47 0.42 0.27 0.27

Scaling Factor - - - 1.02 1.10 1.12

ME 0.02 -0.17 -0.21 0.00 0.00 0.00

B3LYP MAE 0.07 0.18 0.21 0.07 0.09 0.10

MAXE 0.18 0.37 0.41 0.16 0.26 0.25

Scaling Factor - - - 0.99 1.09 1.11

ME 0.08 -0.14 -0.18 0.00 0.00 0.00

CAM-

B3LYP

MAE 0.08 0.14 0.18 0.04 0.08 0.09

MAXE 0.16 0.30 0.34 0.07 0.15 0.16

Scaling Factor 0.96 1.07 1.09

Table S3.10: Molar attenuation coefficient, ε (cm-1/M) with BLYP functional in combination

of variant applied approaches-basis sets for the given PP-Benchmark-set. The ε (cm-1/M) for

all the porphyrinoids (benchmark-set) are taken from the references [195-206].

BLYP Functional

Molecules Ref. ε

(cm-1/M)

TDA

(def2-SVP)

RPA

(def2-SVP)

sTDA

(def2-SVP)

sTD-DFT

(def2-SVP)

sTDA

(def2-TZVP)

sTD-DFT

(def2-TZVP)

H2PP (1300) 100 253 70 186 39 114

(3000) 2 87 41 32 1 61

H2OEP (9000) 56 12 223 118 376 253

Page 133: Theoretical Investigations of the Photophysical Properties

127

(16000) 632 287 1315 664 820 430

MgOEP (23000) 600 517 843 645 660 480

ZnOEP (38000) 603 707 810 594 618 458

H2TPP (4500) 2369 3289 2492 3091 2265 2686

(8000) 2838 4221 3174 4508 3320 4248

MgTPP (11000) 2009 2873 2599 3447 2757 3330

ZnTPP (5000) 1454 2185 1730 2429 1935 2382

H2TAPP (14) 410 662 997 1131 1808 1949

(8) 4 4 2168 2428 2938 3399

ZnTCPP (NA) 3687 4342 4236 4951 3970 4413

F-ZnP (8400) 1711 2371 2249 2971 2456 2916

DA-ZnP (51000) 61 77166 26 79667 60 74145

H2OBP (7500) 7405 8432 8871 9137 8735 8712

(13200) 9495 10955 11839 12652 11342 11533

Table S3.11: Molar attenuation coefficient, ε (cm-1/M) with B3LYP functional in

combination of variant applied approaches-basis sets for the given PP-Benchmark-set.

B3LYP Functional

Molecules Ref. ε

(cm-1/M)

TDA

(def2-SVP)

RPA

(def2-SVP)

sTDA

(def2-SVP)

sTD-DFT

(def2-SVP)

sTDA

(def2-TZVP)

sTD-DFT

(def2-TZVP)

H2PP (1300) 1 29 0 22 10 0

(3000) 71 0 104 1 33 2

H2OEP (9000) 442 566 680 560 873 728

(16000) 1284 1264 1631 1182 1127 793

MgOEP (23000) 1398 1472 1431 1327 1196 1018

ZnOEP (38000) 2002 2240 2264 2150 1767 1583

H2TPP (4500) 1730 2517 1833 2364 1433 1821

(8000) 2314 3656 2600 3900 2304 3233

MgTPP (11000) 1428 2172 1899 2658 1704 2248

ZnTPP (5000) 929 1516 1153 1710 1158 1582

Page 134: Theoretical Investigations of the Photophysical Properties

128

H2TAPP (14) 1538 2223 1675 2075 1133 1414

(8) 1883 2686 2266 3108 1604 2259

ZnTCPP (NA) 1863 2505 2300 2948 2034 2549

F-ZnP (8400) 1076 1617 1476 2061 1407 1865

DA-ZnP (51000) 86077 77451 87082 76493 77904 69693

H2OBP (7500) 9855 11039 10808 10904 9556 9591

(13200) 11352 13415 12793 14121 11105 11807

Table S3.12: Molar attenuation coefficient, ε (cm-1/M) with CAM-B3LYP functional in

combination of variant applied approaches-basis sets for the given PP-Benchmark-set.

CAM-B3LYP Functional

Molecules Ref. ε

(cm-1/M)

TDA

(def2-SVP)

RPA

(def2-SVP)

sTDA

(def2-SVP)

sTD-DFT

(def2-SVP)

sTDA

(def2-

TZVP)

sTD-DFT

(def2-

TZVP)

H2PP (1300) 101 90 67 18 204 135

(3000) 252 112 702 231 329 137

H2OEP (9000) 1230 1920 1739 1595 2288 2106

(16000) 2120 2292 4807 4171 3035 2673

MgOEP (23000) 2503 2632 4285 4194 3513 3327

ZnOEP (38000) 3499 3674 6445 6334 5028 4830

H2TPP (4500) 948 1188 1472 1884 991 1210

(8000) 1679 2579 2083 2805 1880 2741

MgTPP (11000) 778 1169 1122 1818 1042 1423

ZnTPP (5000) 406 690 397 816 493 739

H2TAPP (14) 738 933 1076 1341 652 818

(8) 860 1477 738 1590 730 1225

ZnTCPP (NA) 716 1049 658 1071 702 985

F-ZnP (8400) 466 744 541 963 596 868

DA-ZnP (51000) 63927 57422 76664 69434 71986 66438

H2OBP (7500) 10087 9775 15068 14156 13147 12483

(13200) 11002 12109 18425 20654 14395 15424

Page 135: Theoretical Investigations of the Photophysical Properties

129

B.2 Supporting Information (SI) of Chapter 4 (S4)

Computational Screening of Surface-mounted Metal-Organic

Frameworks Assembled from Porphyrins

COMPUTATIONAL SECTION

Figure S4.1 Library of all porphyrins (PPs) selected for computational investigations.

Page 136: Theoretical Investigations of the Photophysical Properties

130

S4.2. Calculated and experimental (selected) absorption wavelengths of the investigated

porphyrins with employed abbreviations (in nm).

Porphyrin

Structures

Calculated

Experimental

1-H 641, 550 646, 546

1-F 624, 538 --

1-Cl 704, 603 --

1-Br 692, 614 740, 615

1-Me 692, 595 --

2 632, 551 --

3 637, 548 --

4 661, 576 --

5 692, 613 688, 595

6 662, 631 --

7 701, 663 --

8 788, 542 --

9 664, 583 --

10 638, 546 598, 512

Page 137: Theoretical Investigations of the Photophysical Properties

131

Figure S4.3 Comparison of the calculated absorption spectra of all the porphyrin linkers (see

Figure S1) with 1-H.

Page 138: Theoretical Investigations of the Photophysical Properties

132

S4.4 List of all investigated PPs molecules with employed abbreviation and IUPAC name.

Abbreviation IUPAC Name

1-H 5,15-diphenyl-10,20-bis(4-carboxyphenyl) porphyrin

1-F Octafluoro-5,15-diphenyl-10,20-bis(4-carboxyphenyl) porphyrin

1-Cl Octachloro-5,15-diphenyl-10,20-bis(4-carboxyphenyl) porphyrin

1-Br Octabromo-5,15-diphenyl-10,20-bis(4-carboxyphenyl) porphyrin

1-Me Octamethyl-5,15-diphenyl-10,20-bis(4-carboxyphenyl) porphyrin

2 5,15-bis(pyridyl)-10,20-bis(4-carboxyphenyl) porphyrin

3 5,15-bis(bipyridyl)-10,20-bis(4-carboxyphenyl) porphyrin

4 5,15-bis(ethynyl)-10,20-bis(4-carboxyphenyl)porphyrin

5 5,15-bis(phenylethynyl)-10,20-bis(4-carboxyphenyl) porphyrin

6 5,15-diphenyl-10,20-bis(4-carboxyphenyl) dibenzo[b,l]-porphyrin

7 5,15-diphenyl-10,20-bis(4-carboxyphenyl) tetrabenzo-porphyrin

8 5,15-diphenyl-10,20-bis(4-carboxyphenyl) bacteriochlorin

9 5,15-bis(tert-butyl)-10,20-bis(4-carboxyphenyl) porphyrin

10 5,15-bis(3,4,5-trifluorophenyl)-10,20-bis(4 carboxphenyl)porphyrin

S4.5 Calculated lattice parameters for the selected PP-based SURMOF structures.

Structures a (in Å) b (in Å) c (in Å)

1-Br ' 23.75 23.75 6.12

5 ' 23.85 23.85 6.22

10 ' 23.85 23.85 6.38

S4.6 Calculated band gap for the selected PP-based SURMOF structures.

Structures Band gap (in eV)

1-Br ' 1.15

5 ' 1.09

10 ' 1.72

Page 139: Theoretical Investigations of the Photophysical Properties

133

S4.7 Structural analysis of the selected PP-based SURMOFs shown in Fig. S8

Structures Distance, d (in Å) Angle, α (in °) Shift, s (in Å)

1-Br' (x-direction) 6.023 79.8 1.087

1-Br' (y-direction) 6.069 82.6 0.785

5' (x-direction) 3.314 32.2 5.262

5' (y-direction) 3.305 32.1 5.268

10' (x-direction) 6.358 85.2 0.529

10' (y-direction) 6.335 83.2 0.755

Note: slight differences in stacking values are due to the numerical noise.

-(d) distance between the PPs

-(α) angle with respect to crystal plane

-(s) shift from eclipsed stacking

-(x), -(y) denotes the stacking directions

Page 140: Theoretical Investigations of the Photophysical Properties

134

Figure S4.8 Stacking analysis of the selected tuned-PPs (a) 1-Br' (b) 5' and (c) 10'

Figure S4.9 Band structure of the 5' with zoom-in of the valence band and conduction band

at PBE-level of theory

Page 141: Theoretical Investigations of the Photophysical Properties

135

EXPERIMENTAL SECTION

Materials:

Reactions which require dry solvents were prepared using standard Schlenk conditions. Liquid

reagents were added via plastic syringes with stainless steel cannulas. Non-dry solvents were

used in p.a. quality (pro analysi) purchased from Fisher Scientific and Sigma Aldrich without

further purification. Dry THF was freshly distilled over potassium prior to use, dry

triethylamine was bought from Sigma-Aldrich and used without further purification.

Chemicals were purchased from ABCR, Alfa Aesar, Carbolution, Chempur, Sigma-Aldrich

and TCI and used without further purification unless stated otherwise. Monitoring of reactions

was done by TLC using silica gel coated aluminium plates (TLC silica gel 60 F254) purchased

from Merck and a UV lamp emitting with λ = 254 nm. Column chromatographies were

performed using silica gel 60 (0.040–0.063 mm, 230–400 mesh ASTM) purchased from Merck

as stationary phase and solvents in p.a. quality.

Zinc acetate dihydrate was purchased from Merck Millipore. 16-mercaptohexadecanoic acid

(MHDA, 97%), was purchased from Sigma-Aldrich (Germany). Absolute ethanol was

purchased from VWR (Germany).

Figure S4.10 Band structure of the 5' with zoom-in of the valence band and conduction

band at PBE0-level of theory

Page 142: Theoretical Investigations of the Photophysical Properties

136

Substrates:

The silicon substrates with a [100] orientation are from Silicon Sense (US). The quartz glasses

are from Alfa Aesar. These substrates were treated with plasma (Diener Plasma) under O2 (50

sccm) for 30 min to remove the impurities and generated a surface with hydroxyl groups.

X-ray diffraction (XRD)

The XRD measurements for out-of-plane (co-planar orientation) were carried out using a

Bruker D8-Advance diffractometer equipped with a position sensitive detector Lynxeye in

geometry, variable divergence slit and 2.3° Soller-slit was used on the secondary side. The Cu-

anodes which utilize the Cu Kα1,2-radiation (ʎ = 0.154018 nm) was used as source.

Characterizations of Porphyrin linkers by 1H-NMR, 13C NMR and HR MS

All NMR spectra were recorded on a BRUKER Avanche 400 (1H NMR: 400 MHz, 13C NMR:

100 MHz, 19F NMR: 377 MHz) at room temperature using deuterated chloroform purchased

from Eurisotop. Chemical shift δ were given in ppm, with the residual solvent peak as reference

(7.26 ppm for 1H and 77.16 ppm for 13C NMR). Coupling constants J were given in Hertz (Hz)

as absolute values. For examination of spectra the following abbreviations are used: s = singlet,

bs = broad singlet, d = doublet, t = triplet, q = quartet, dd = doublet of doublets, td = triplet of

doublets, m = multiplet. The spectra were analyzed according to first order. For multiplicities

in 13C spectra the following abbreviations were used: + = primary or tertiary C, – = secondary

C, Cq = quaternary C. For the assignment of signals the following indices were used: meso =

meso position of porphyrin, Pyr = pyrrolic positions of porphyrin, Ph = phenyl.

(High resolution) Mass spectra were recorded on a Finnigan MAT 95 instrument using either

FAB (fast atom bombardement), with 3-nitrobenzyl alcohol used as matrix or EI (electron

impact) with 70 eV as ionization method. Mass spectra were interpreted by listing the

mass/charge ratios (m/z) of molecule fragments together with their intensities relative to the

base peak (100%).

Syntheses of Porphyrin linkers:

Di(1H-pyrrol-2-yl)methane (11)298

Page 143: Theoretical Investigations of the Photophysical Properties

137

To 340 mL of freshly distilled pyrrole (330 g, 4.92 mol, 92.7 equiv.) were added 1.59 g of

paraformaldehyde (53.1 mmol, 1.00 equiv.) and the resulting suspension was stirred for 10 min

at 55 °C. Then 1.16 g InCl3 (5.25 mmol, 0.10 equiv.) were added and the resulting mixture was

stirred for additional 3 h at 55 °C. After cooling down to rt, 7.02 g powdered NaOH (176 mmol,

3.31 equiv.) were added and the mixture was stirred for another hour, followed by filtration.

The filtrate was evaporated under reduced pressure and the remaining crude product was

purified by column chromatography (CH/EE 10:1 with gradient to 2:1) to yield 5.05 g of 11

(34.5 mmol, 65%) as a white solid.

Rf (CH/EE 2:1) = 0.57. – 1H-NMR (400 MHz, CDCl3): (ppm) = 3.94 (s, 2H, CH2), 6.03–

6.09 (m, 2H), 6.18 (q, J = 2.9 Hz, 2H), 6.62 (td, J = 2.7, 1.6 Hz, 2H), 7.67 (bs, 2H, NH). – 13C-

NMR (100 MHz, CDCl3): (ppm) = 26.4 (–, CH2), 106.6 (+), 108.4 (+), 117.5 (+), 129.2 (Cq).

– IR (ATR): = 3325 (m), 1561 (w), 1468 (w), 1439 (w), 1327 (w), 1244 (w), 1181 (w), 1119

(w), 1108 (w), 1095 (w), 1024 (m), 961 (w), 884 (w), 857 (vw), 797 (m), 720 (s), 667 (m), 600

(m), 586 (m) cm–1. – MS (EI, 70 eV): m/z (%) = 146.1 (100) [M]+, 147.1 (11), 145.1 (70), 80.1

(30) [C5H6N]+. – HRMS (C9H10N2): ber.: 146.0844, gef.: 146.0845.

Ethyl 4-formylbenzoate (12)299

To a solution of 4.87 g 4-formylbenzoic acid (32.4 mmol, 1.00 equiv.) in 125 mL of DMF were

added 8.68 g K2CO3 (62.8 mmol, 1.94 equiv.) and 6.60 mL iodoethane (12.8 g, 62.8 mmol,

2.54 equiv.). After stirring the reaction for 3 h at rt, water was added, the phases were separated

and the aqueous phase was extracted with diethyl ether two times. The combined organic

phases were washed with brine, dried over MgSO4 and filtered. After removal of the solvent

under reduced pressure, the crude product was purified by column chromatography (CH/EE

6:1) to yield 4.90 g of 12 (27.5 mmol, 85%) as a light yellowish liquid.

Rf (CH/EE 6:1) = 0.83. – 1H-NMR (400 MHz, CDCl3): (ppm) = 1.38 (t, J = 7.1 Hz, 3H,

CH3), 4.38 (q, J = 7.1 Hz, 2H, CH2), 7.91 (d, J = 8.5 Hz, 2H, HPh), 8.16 (d, J = 8.5 Hz, 2H,

HPh), 10.06 (s, 1H, OCH). – 13C-NMR (100 MHz, CDCl3): (ppm) = 14.3 (+, CH3), 61.6 (–,

CH2), 129.5 (+, CPhH), 130.2 (+, CPhH), 135.5 (Cq), 139.2 (Cq), 165.6 (Cq, COO), 191.7 (+,

OCH). – IR (ATR): = 2924 (w), 2854 (w), 1701 (s), 1577 (w), 1503 (vw), 1448 (w), 1367

(w), 1272 (m), 1200 (w), 1172 (w), 1103 (m), 1017 (m), 854 (w), 818 (w), 758 (m), 733 (w),

690 (w), 630 (vw), 461 (vw) cm–1. – MS (EI, 70 eV): m/z (%) = 178.1 (44) [M]+, 179.1 (5),

Page 144: Theoretical Investigations of the Photophysical Properties

138

149.1 (24) [M – CO]+, 133.1 (100) [M – C2H5O]+. – HRMS (C10H10O3): ber.: 178.0630, gef.:

178.0630.

5,15-Bis(4-ethoxycarbonylphenyl)porphyrin (13)300

Through a solution of 1.10 g di(1H-pyrrol-2-yl)methane (11) (7.53 mmol, 2.00 equiv.) and

1.36 g ethyl 4-formylbenzoate (12) (7.62 mmol, 2.02 equiv.) in 1.50 L CHCl3 was passed Ar

gas for 30 min, followed by the dropwise addition of 580 µL TFA (858 mg, 7.53 mmol,

2.00 equiv.). The reaction was stirred for 17 h in the dark, after which time 3.21 mL NEt3

(2.35 g, 23.2 mmol, 6.16 equiv.) and 5.52 g p-chloranil (22.4 mmol, 5.96 equiv.) were added

in this order. The mixture was then refluxed for 90 min and the solvent removed under reduced

pressure. After a filtration through silica gel (DCM) to remove most of the oligomeric side

products, the crude product was purified by column chromatography (DCM/EE 1:0 with

gradient to 50:1). The obtained solid was thoroughly washed with MeOH, leaving 1.07 g of 13

(1.76 mmol, 47%) as a purple solid.

Rf (DCM) = 0.42. – 1H-NMR (400 MHz, CDCl3): (ppm) = –3.13 (bs, 2H, NH), 1.58 (t,

J = 7.1 Hz, 6H, CH3), 4.61 (q, J = 7.1 Hz, 4H, CH2), 8.36 (d, J = 8.2 Hz, 4H, HPh), 8.51 (d,

J = 8.2 Hz, 4H, HPh), 9.04 (d, J = 4.6 Hz, 4H, HPyr), 9.42 (d, J = 4.6 Hz, 4H, HPyr), 10.34 (s,

2H, Hmeso). – 13C-NMR (100 MHz, CDCl3): (ppm) = 14.7 (+, CH3), 61.5 (–, CH2), 105.8 (+),

118.3 (Cq), 128.3 (+), 130.1 (Cq), 130.9 (+), 132.2 (+), 135.0 (+), 145.5 (Cq), 146.2 (Cq), 146.8

(Cq), 167.0 (Cq, COO). – UV/VIS (CHCl3): max (log ) = 405 (5.27), 504 (4.22), 539 (3.85),

576 (3.74), 631 (3.33) nm. – IR (ATR): = 3484 (vw), 3276 (vw), 2976 (vw), 1703 (m), 1602

(w), 1437 (w), 1398 (w), 1363 (w), 1305 (vw), 1268 (m), 1239 (w), 1194 (w), 1173 (w), 1096

(m), 1051 (w), 1017 (w), 986 (w), 971 (w), 952 (w), 900 (w), 868 (m), 843 (w), 812 (w), 792

(m), 752 (w), 736 (m), 723 (m), 691 (m), 520 (vw), 489 (w), 435 (w), 411 (vw) cm–1. – HRMS

(C38H31O4N4): ber.: 607.2340, gef.: 607.2340.

Cross coupling reactions were performed according to a known procedure established by Senge

et al.301

5,15-Dibromo-10,20-bis(4-ethoxycarbonylphenyl)porphyrin (14)302-303

Page 145: Theoretical Investigations of the Photophysical Properties

139

To a solution of 933 mg porphyrin 13 (1.54 mmol, 1.00 equiv.) in 385 mL of CHCl3 were

added 0.38 mL pyridine (377 mg, 4.77 mmol, 3.10 equiv.). The mixture was cooled to 0 °C

and 646 mg NBS were added, as well as 50 mg in intervals of 30 min respectively, until a

complete conversion was observed by TLC (∑ 746 mg, 4.19 mmol, 2.72 equiv.). After addition

of the last portion, the reaction was stirred for 30 min and the mixture was directly filtered

through silica gel eluting with DCM. The solvent of the filtrate was removed under reduced

pressure and the crude product was purified by column chromatography (DCM). The obtained

solid was recrystallized from DCM/MeOH yielding 1.00 g of porphyrin 14 (1.31 mmol, 85%)

as a purple solid.

Rf (DCM) = 0.60. – 1H-NMR (400 MHz, CDCl3): (ppm) = –2.77 (bs, 2H, NH), 1.57 (t,

J = 7.1 Hz, 6H, CH3), 4.60 (q, J = 7.1 Hz, 4H, CH2), 8.23 (d, J = 8.1 Hz, 4H, HPh), 8.47 (d,

J = 8.1 Hz, 4H, HPh), 8.78 (d, J = 4.8 Hz, 4H, HPyr), 9.62 (d, J = 4.9 Hz, 4H, HPyr). – 13C-

NMR (100 MHz, CDCl3): (ppm) = 14.7 (+, CH3), 61.6 (–, CH2), 104.3 (Cq), 120.4 (Cq),

128.2 (+), 130.5 (Cq), 134.6 (+), 146.1 (Cq), 166.8 (Cq, COO). – UV/VIS (CHCl3): max

(log ) = 423 (5.42), 522 (4.21), 557 (4.01), 601 (3.66), 659 (3.62) nm. – IR (ATR): =3314

(vw), 2921 (w), 1709 (m), 1605 (w), 1557 (w), 1465 (w), 1397 (w), 1366 (w), 1336 (w), 1302

(w), 1272 (m), 1193 (w), 1175 (w), 1122 (w), 1106 (m), 1018 (w), 997 (w), 979 (w), 961 (m),

869 (w), 848 (w), 795 (m), 785 (m), 755 (m), 729 (m), 706 (m), 630 (w), 555 (vw), 523 (w),

500 (w), 458 (w), 394 (vw) cm–1. – HRMS (C38H29O4N479Br81Br): ber.: 765.0530, gef.:

765.0528.

5,15-Bis(3,4,5-trifluorophenyl)-10,20-bis(4-ethoxycarbonylphenyl)porphyrin (15)

Page 146: Theoretical Investigations of the Photophysical Properties

140

Under Ar atmosphere 95 mg of porphyrin 14 (124 µmol, 1.00 equiv.), 265 mg (3,4,5-

trifluorophenyl)boronic acid (1.51 mmol, 12.2 equiv.), 650 mg K3PO4 (3.06 mmol,

24.6 equiv.) and 15 mg Pd(PPh3)4 (13 µmol, 0.10 equiv.) were dissolved in 75 ml of dry THF

and the mixture was heated to 80 °C overnight under protection from light. After cooling of

the solution to rt the solvent was removed under reduced pressure. The residue was dissolved

in DCM, washed with a saturated solution of NaHCO3 and water and dried over Na2SO4. After

evaporation of the solvent under reduced pressure, the product was isolated by column

chromatography (DCM/CH 5:1) and a subsequent thorough wash with methanol, yielding

98 mg of porphyrin 15 (113 µmol, 91%) as a purple solid.

Rf (CH2Cl2) = 0.74. – 1H-NMR (400 MHz, CDCl3): (ppm) = –2.91 (bs, 2H, NH), 1.57 (t,

J = 7.1 Hz, 6H, CH3), 4.60 (q, J = 7.1 Hz, 4H, CH2), 7.80–7.91 (m, 4H, CHCF), 8.30 (d,

J = 8.2 Hz, 4H, HPh), 8.48 (d, J = 8.3 Hz, 4H, HPh), 8.79–8.93 (m, 8H, HPyr). – 13C-NMR

(100 MHz, CDCl3): (ppm) = 14.7 (+, CH3), 61.6 (–, CH2), 117.1 (Cq), 118.7, 118.8, 118.9,

118.9, 120.0 (Cq), 128.2 (+), 130.5 (Cq), 134.6 (+), 137.8, 137.8, 139.1, 141.6, 146.3 (Cq),

148.4, 148.5, 148.5, 148.6 , 150.9, 151.0, 151.0, 151.1, 166.8 (Cq, COO). 19F-NMR (377 MHz,

CDCl3): (ppm) = –140.0 (d, J = 20.6 Hz), –165.4 (t, J = 20.6 Hz). – UV/VIS (CHCl3): max

(log ) = 417 (5.41), 514 (4.26), 549 (3.76), 590 (3.73), 646 (3.30) nm. – IR (ATR): = 3302

(vw), 3070 (vw), 1707 (m), 1607 (w), 1525 (w), 1475 (w), 1423 (w), 1401 (w), 1365 (w), 1307

(vw), 1271 (m), 1237 (w), 1176 (w), 1108 (w), 1040 (m), 1022 (w), 973 (w), 928 (w), 869 (w),

849 (vw), 805 (m), 762 (w), 732 (w), 717 (w), 636 (vw), 561 (vw), 541 (w), 408 (vw) cm–1. –

HRMS (C50H33O4N4F6): ber.: 867.2401, gef.: 867.2403.

5,15-Bis(phenylethynyl)-10,20-bis(4-ethoxycarbonylphenyl)porphyrin (16)

Page 147: Theoretical Investigations of the Photophysical Properties

141

Under Ar atmosphere 77 mg of porphyrin 14 (101 µmol, 1.00 equiv.), 4 mg CuI (20 µmol,

0.20 equiv.) and 7 mg Pd(PPh3)2Cl2 (10 µmol, 0.10 equiv.) were dissolved in 10 mL of dry

THF and 21 mL of dry NEt3. The solution was then purged with Ar gas for 15 min, followed

by the addition of 44 µL phenylacetylen (41 mg, 404 µmol, 4.00 equiv.). The reaction was

stirred overnight at rt and subsequently directly filtered through silica eluting with DCM. The

solvent of the filtrate was removed under reduced pressure and the crude product was purified

by column chromatography (CHCl3), yielding 62 mg of porphyrin 16 (76.8 µmol, 76%) as a

greenish-purple solid.

Rf (DCM) = 0.69. – 1H-NMR (400 MHz, CDCl3): (ppm) = –2.06 (bs, 2H, NH), 1.58 (t,

J = 7.2 Hz, 6H, CH3), 4.61 (q, J = 7.1 Hz, 4H, CH2), 7.47–7.63 (m, 6H, HPh), 7.98–8.07 (m,

4H, HPh), 8.27 (d, J = 8.2 Hz, 4H, HPh), 8.48 (d, J = 8.2 Hz, 4H, HPh), 8.77 (d, J = 4.7 Hz, 4H,

HPyr), 9.68 (d, J = 4.7 Hz, 4h, HPh). – 13C-NMR (100 MHz, CDCl3): (ppm) = 14.7 (+, CH3),

61.5 (–, CH2), 91.8 (Cq), 97.8 (Cq), 101.8 (Cq), 120.8 (Cq), 123.8 (Cq), 128.2 (+), 128.9 (+),

129.0 (+), 130.4 (Cq), 131.9 (+), 134.6 (+), 146.1 (Cq), 166.9 (Cq, COO). – UV/VIS (CHCl3):

max (log ) = 307 (4.45), 442 (5.36), 599 (4.70), 690 (4.32) nm. – IR (ATR): = 2972 (w),

1709 (m), 1604 (w), 1553 (w), 1487 (w), 1470 (w), 1399 (w), 1362 (w), 1264 (m), 1176 (m),

1159 (w), 1099 (m), 1065 (m), 1019 (m), 973 (m), 923 (w), 866 (w), 807 (m), 788 (m), 761

(w), 749 (m), 716 (m), 683 (m), 633 (w), 581 (w), 565 (w), 540 (w), 514 (w), 441 (w) cm–1. –

HRMS (C54H39O4N4): ber.: 807.2966, gef.: 807.2964.

5,15-Diphenyl-10,20-bis(4-ethoxycarbonylphenyl)porphyrin (17)

Page 148: Theoretical Investigations of the Photophysical Properties

142

Under Ar atmosphere 152 mg of porphyrin 14 (198 µmol, 1.00 equiv.), 287 mg phenylboronic

acid (2.36 mmol, 11.9 equiv.), 1.04 g K3PO4 (4.91 mmol, 24.7 equiv.) and 27 mg Pd(PPh3)4

(23 µmol, 0.12 equiv.) were dissolved in 45 ml of dry THF and the mixture was heated to 80 °C

overnight under protection from light. After cooling of the solution the solvent was removed

under reduced pressure. The residue was dissolved in DCM, washed with a saturated solution

of NaHCO3 and water and dried over Na2SO4. After evaporation of the solvent under reduced

pressure, the product was isolated by column chromatography (DCM/CH 1:1 with gradient to

1:0), yielding 134 mg of porphyrin 17 (176 µmol, 89%) as a purple solid.

Rf (DCM) = 0.50. – 1H-NMR (400 MHz, CDCl3): (ppm) = –2.78 (bs, 2H, NH), 1.57 (t, J =

7.1 Hz, 6H, CH3), 4.59 (q, J = 7.2 Hz, 4H, CH2), 7.72–7.85 (m, 6H, HPh), 8.23 (dd, J = 7.6,

1.7 Hz, 2H, HPh), 8.32 (d, J = 8.2 Hz, 4H, HPh), 8.47 (d, J = 8.2 Hz, 4H, HPh), 8.82 (d,

J = 4.9 Hz, 4H, HPyr), 8.89 (d, J = 4.9 Hz, 4H, HPyr). – 13C-NMR (100 MHz, CDCl3): (ppm)

= 14.7 (+, CH3), 61.5 (–, CH2), 119.1 (Cq), 120.7 (Cq), 126.9 (+), 128.0 (+), 130.1 (Cq), 142.0

(Cq), 147.0 (Cq), 167.0 (Cq, COO). – UV/VIS (CHCl3): max (log ) = 419 (5.39), 516 (4.28),

551 (3.92), 591 (3.76), 646 (3.57) nm. – IR (ATR): = 3297 (vw), 3048 (vw), 3103 (vw), 2906

(w), 2978 (vw), 1713 (w), 1604 (vw), 1473 (vw), 1438 (vw), 1401 (vw), 1364 (vw), 1309 (vw),

1269 (w), 1177 (w), 1099 (w), 1023 (w), 980 (w), 964 (w), 863 (vw), 799 (w), 754 (w), 730

(w), 704 (w), 656 (vw), 635 (vw), 564 (vw), 455 (vw) cm–1. – HRMS (C50H39N4O4): ber.:

759.2971, gef.: 759.2969.

2,3,7,8,12,13,17,18-Octabromo-5,15-Diphenyl-10,20-bis(4-ethoxycarbonylphenyl)porphyrin

(18)

Page 149: Theoretical Investigations of the Photophysical Properties

143

The reaction was performed similar to a literature known procedure.304

To a solution of 39 mg of porphyrin 17 (51.4 µmol, 1.00 equiv.) in 16 mL of CHCl3 were added

821 mg copper acetate monohydrate (411 µmol, 8.00 equiv.). The reaction mixture was stirred

at rt until complete conversion to copper porphyrin was detected by TLC (< 3 h). Then 430 µL

of molecular bromine (1.36 g, 8.48 mmol, 165 equiv.) were added to the reaction mixture

directly and the solution was stirred overnight at rt. An aqueous solution of sodium thiosulfate

was added to quench the reaction, followed by washing of the reaction mixture with H2O for

four times. After filtration 4.0 mL of perchloric acid (70% solution in water, 6.6 g, 46 mmol,

900 equiv.) were added to the solution. The reaction mixture was vigorously stirred at rt

overnight again for demetalization. After addition of 10 mL of water, the organic layer was

separated and washed with water, a solution of sodium bicarbonate and water again, and dried

over Na2SO4. After removing the solvent under reduced pressure, the crude product was

isolated by column chromatography (DCM/EE 1:0 with gradient to 50:1) yielding 66 mg of the

octabromo porphyrin 18 (47.7 µmol, 93%) as a greenish-purple solid.

Rf (DCM) = 0.80. – 1H-NMR (400 MHz, CDCl3): (ppm) = –1.56 (bs, 2H, NH), 1.55 (t,

J = 7.1 Hz, 6H, CH3), 4.55 (q, J = 7.1 Hz, 4H, CH2), 7.72–7.90 (m, 6H, HPh), 8.17–8.25 (m,

4H, HPh), 8.31 (d, J = 8.3 Hz, 4H, HPh), 8.45 (d, J = 8.3 Hz, 4H, HPh). – 13C-NMR (100 MHz,

CDCl3): (ppm) = 14.6 (+, CH3), 61.6 (–, CH2), 119.6 (Cq), 121.4 (Cq), 128.6 (+), 129.5 (+),

130.1 (+), 131.5 (Cq), 136.9 (+), 137.1 (Cq), 141.2 (Cq), 166.9 (Cq, COO). – UV/VIS (CHCl3):

max (log ) = 372 (4.45), 472 (5.30), 572 (3.99), 627 (4.13), 740 (3.90) nm. – IR (ATR): =

2973 (vw), 2924 (vw), 1716 (w), 1605 (vw), 1464 (vw), 1403 (vw), 1365 (vw), 1269 (w), 1175

(vw), 1101 (vw), 1019 (vw), 1004 (w), 914 (vw), 818 (vw), 749 (vw), 727 (vw), 693 (vw), 607

(vw), 401 (vw) cm–1. – MS (FAB, 3-NBA): m/z (%) = 1390.3 (100) [M(79Br481Br4) + H]+.

(Fitting isotope pattern for eight bromines)

General procedure for saponification of porphyrins305

Page 150: Theoretical Investigations of the Photophysical Properties

144

To a solution of the corresponding ester porphyrin 15, 16 or 18 (1.00 equiv.) in THF/MeOH

(4:1, 9.95 M) was added a 40w% aqueous solution of NaOH (1000 equiv.) and the mixture was

stirred at 90 °C overnight. After cooling to rt, the organic solvents were removed under reduced

pressure and the resulting water suspension was acidified with hydrochloric and acetic acid to

pH = 3. The mixture was cooled to 4 °C and subsequently filtrated. The precipitate was washed

with hot water and DCM and the remaining solid was dissolved in MeOH/EtOH/DMF/NEt3.

The solvents were removed under reduced pressure and the residual solid was dried in vacuo.

In all saponification reactions the determination of the respective yield resulted in >100% due

to possible formation of carboxylate salts. The obtained solids 1-Br, 5 and 10 were further used

without additional purification.

Fabrication of multilayer heteroepitaxial SURMOF-2:

Figure: Out-of-plane XRD of the heteroepitaxial SURMOF structure 1-Br′/5′/10′ (green), and

simulated XRD of 1-Br′ (black).

3 6 9 12 15

(002)

2

(001)(003)

Page 151: Theoretical Investigations of the Photophysical Properties

145

Bibliography

1. Green, M. A.; Dunlop, E. D.; Hohl‐Ebinger, J.; Yoshita, M.; Kopidakis, N.; Hao, X., Solar cell efficiency

tables (version 56). Prog. Photovolt: Res. Appl. 2020, 28 (7), 629-638.

2. Hagfeldt, A.; Boschloo, G.; Sun, L.; Kloo, L.; Pettersson, H., Dye-sensitized solar cells. Chem. Rev.

2010, 110 (11), 6595-6663.

3. Dou, L.; You, J.; Hong, Z.; Xu, Z.; Li, G.; Street, R. A.; Yang, Y., 25th anniversary article: a decade of

organic/polymeric photovoltaic research. Adv. Mater 2013, 25 (46), 6642-6671.

4. Saparov, B.; Mitzi, D. B., Organic–inorganic perovskites: structural versatility for functional materials

design. Chem. Rev. 2016, 116 (7), 4558-4596.

5. Tang, C. W., Two‐layer organic photovoltaic cell. Appl. Phys. Lett 1986, 48 (2), 183-185.

6. Seo, D.-K.; Hoffmann, R., Direct and indirect band gap types in one-dimensional conjugated or stacked

organic materials. Theor. Chem. Acc 1999, 102 (1-6), 23-32.

7. Gust, D.; Moore, T. A.; Moore, A. L., Mimicking photosynthetic solar energy transduction. Acc. Chem.

Res 2001, 34 (1), 40-48.

8. Gust, D.; Moore, T. A.; Moore, A. L., Molecular mimicry of photosynthetic energy and electron transfer.

Acc. Chem. Res 1993, 26 (4), 198-205.

9. Campbell, W. M.; Burrell, A. K.; Officer, D. L.; Jolley, K. W., Porphyrins as light harvesters in the dye-

sensitised TiO2 solar cell. Coord. Chem. Rev. 2004, 248 (13-14), 1363-1379.

10. Martinez-Diaz, M. V.; de la Torre, G.; Torres, T., Lighting porphyrins and phthalocyanines for molecular

photovoltaics. Chem. Commun. 2010, 46 (38), 7090-7108.

11. Biesaga, M.; Pyrzyńska, K.; Trojanowicz, M., Porphyrins in analytical chemistry. A review. Talanta

2000, 51 (2), 209-224.

12. Milgrom, L. R.; Warren, M. J., The colours of life: an introduction to the chemistry of porphyrins and

related compounds. 1997.

13. Gouterman, M., Spectra of Porphyrins. J. Mol. Spectrosc. 1961, 6, 138.

14. Hashimoto, T.; Choe, Y.-K.; Nakano, H.; Hirao, K., Theoretical study of the Q and B bands of free-base,

magnesium, and zinc porphyrins, and their derivatives. J. Phys. Chem. A 1999, 103 (12), 1894-1904.

15. Rosa, A.; Ricciardi, G.; Baerends, E.; van Gisbergen, S. A., The optical spectra of NiP, NiPz, NiTBP,

and NiPc: electronic effects of meso-tetraaza substitution and tetrabenzo annulation. J. Phys. Chem. A

2001, 105 (13), 3311-3327.

16. Zhou, H.; Li, X.; Fan, T.; Osterloh, F. E.; Ding, J.; Sabio, E. M.; Zhang, D.; Guo, Q., Artificial inorganic

leafs for efficient photochemical hydrogen production inspired by natural photosynthesis. Adv. Mater

2010, 22 (9), 951-956.

17. Huijser, A.; Suijkerbuijk, B. M.; Klein Gebbink, R. J.; Savenije, T. J.; Siebbeles, L. D., Efficient exciton

transport in layers of self-assembled porphyrin derivatives. J. Am. Chem. Soc. 2008, 130 (8), 2485-2492.

18. Martin, K. E.; Wang, Z.; Busani, T.; Garcia, R. M.; Chen, Z.; Jiang, Y.; Song, Y.; Jacobsen, J. L.; Vu, T.

T.; Schore, N. E., Donor− acceptor biomorphs from the ionic self-assembly of porphyrins. J. Am. Chem.

Soc. 2010, 132 (23), 8194-8201.

19. Ng, K. K.; Lovell, J. F.; Vedadi, A.; Hajian, T.; Zheng, G., Self-assembled porphyrin nanodiscs with

structure-dependent activation for phototherapy and photodiagnostic applications. ACS Nano 2013, 7 (4),

3484-3490.

20. Zhou, H.-C.; Long, J. R.; Yaghi, O. M., Introduction to metal–organic frameworks. Chem. Rev. : 2012;

Vol. 112, p 673.

Page 152: Theoretical Investigations of the Photophysical Properties

146

21. Zhao, M.; Ou, S.; Wu, C.-D., Porous metal–organic frameworks for heterogeneous biomimetic catalysis.

Acc. Chem. Res 2014, 47 (4), 1199-1207.

22. Lu, W.; Wei, Z.; Gu, Z.-Y.; Liu, T.-F.; Park, J.; Park, J.; Tian, J.; Zhang, M.; Zhang, Q.; Gentle III, T.,

Tuning the structure and function of metal–organic frameworks via linker design. Chem. Soc. Rev 2014,

43 (16), 5561-5593.

23. Kinoshita, Y.; Matsubara, I.; Higuchi, T.; Saito, Y., The crystal structure of bis (adiponitrilo) copper (I)

nitrate. Bull. Chem. Soc. Jpn. 1959, 32 (11), 1221-1226.

24. Yaghi, O.; Li, H., Hydrothermal synthesis of a metal-organic framework containing large rectangular

channels. J. Am. Chem. Soc. 1995, 117 (41), 10401-10402.

25. Yaghi, O. M.; Li, G.; Li, H., Selective binding and removal of guests in a microporous metal–organic

framework. Nature 1995, 378 (6558), 703-706.

26. Li, H.; Eddaoudi, M.; O'Keeffe, M.; Yaghi, O. M., Design and synthesis of an exceptionally stable and

highly porous metal-organic framework. nature 1999, 402 (6759), 276-279.

27. Moghadam, P. Z.; Li, A.; Wiggin, S. B.; Tao, A.; Maloney, A. G.; Wood, P. A.; Ward, S. C.; Fairen-

Jimenez, D., Development of a Cambridge Structural Database subset: a collection of metal–organic

frameworks for past, present, and future. Chem. Mater 2017, 29 (7), 2618-2625.

28. Chen, Z.; Li, P.; Anderson, R.; Wang, X.; Zhang, X.; Robison, L.; Redfern, L. R.; Moribe, S.; Islamoglu,

T.; Gómez-Gualdrón, D. A., Balancing volumetric and gravimetric uptake in highly porous materials for

clean energy. Science 2020, 368 (6488), 297-303.

29. Rowsell, J. L.; Spencer, E. C.; Eckert, J.; Howard, J. A.; Yaghi, O. M., Gas adsorption sites in a large-

pore metal-organic framework. Science 2005, 309 (5739), 1350-1354.

30. Bourrelly, S.; Llewellyn, P. L.; Serre, C.; Millange, F.; Loiseau, T.; Férey, G., Different adsorption

behaviors of methane and carbon dioxide in the isotypic nanoporous metal terephthalates MIL-53 and

MIL-47. J. Am. Chem. Soc. 2005, 127 (39), 13519-13521.

31. Haque, E.; Lee, J. E.; Jang, I. T.; Hwang, Y. K.; Chang, J.-S.; Jegal, J.; Jhung, S. H., Adsorptive removal

of methyl orange from aqueous solution with metal-organic frameworks, porous chromium-

benzenedicarboxylates. J. Hazard. Mater. 2010, 181 (1-3), 535-542.

32. Haque, E.; Jun, J. W.; Jhung, S. H., Adsorptive removal of methyl orange and methylene blue from

aqueous solution with a metal-organic framework material, iron terephthalate (MOF-235). J. Hazard.

Mater. 2011, 185 (1), 507-511.

33. Jhung, S. H.; Lee, J. H.; Yoon, J. W.; Serre, C.; Férey, G.; Chang, J. S., Microwave synthesis of chromium

terephthalate MIL‐101 and its benzene sorption ability. Adv. Mater 2007, 19 (1), 121-124.

34. Achmann, S.; Hagen, G.; Kita, J.; Malkowsky, I. M.; Kiener, C.; Moos, R., Metal-organic frameworks

for sensing applications in the gas phase. Sensors 2009, 9 (3), 1574-1589.

35. Kumar, P.; Deep, A.; Kim, K.-H., Metal organic frameworks for sensing applications. Trends Anal. Chem

2015, 73, 39-53.

36. Bhattacharjee, S.; Yang, D.-A.; Ahn, W.-S., A new heterogeneous catalyst for epoxidation of alkenes via

one-step post-functionalization of IRMOF-3 with a manganese (II) acetylacetonate complex. Chem.

Commun. 2011, 47 (12), 3637-3639.

37. Kim, J.; Kim, S.-N.; Jang, H.-G.; Seo, G.; Ahn, W.-S., CO2 cycloaddition of styrene oxide over MOF

catalysts. Appl. Catal. 2013, 453, 175-180.

38. Lee, J.; Farha, O. K.; Roberts, J.; Scheidt, K. A.; Nguyen, S. T.; Hupp, J. T., Metal–organic framework

materials as catalysts. Chem. Soc. Rev. 2009, 38 (5), 1450-1459.

39. Huxford, R. C.; Della Rocca, J.; Lin, W., Metal–organic frameworks as potential drug carriers. Curr.

Opin. Chem. Biol. 2010, 14 (2), 262-268.

40. Mingabudinova, L.; Vinogradov, V.; Milichko, V.; Hey-Hawkins, E.; Vinogradov, A., Metal–organic

frameworks as competitive materials for non-linear optics. Chem. Soc. Rev. 2016, 45 (19), 5408-5431.

41. Lustig, W. P.; Mukherjee, S.; Rudd, N. D.; Desai, A. V.; Li, J.; Ghosh, S. K., Metal–organic frameworks:

functional luminescent and photonic materials for sensing applications. Chem. Soc. Rev. 2017, 46 (11),

3242-3285.

Page 153: Theoretical Investigations of the Photophysical Properties

147

42. Nguyen, T. N.; Ebrahim, F. M.; Stylianou, K. C., Photoluminescent, upconversion luminescent and

nonlinear optical metal-organic frameworks: From fundamental photophysics to potential applications.

Coord. Chem. Rev. 2018, 377, 259-306.

43. Stavila, V.; Talin, A. A.; Allendorf, M. D., MOF-based electronic and opto-electronic devices. Chem.

Soc. Rev. 2014, 43 (16), 5994-6010.

44. Stassen, I.; Burtch, N.; Talin, A.; Falcaro, P.; Allendorf, M.; Ameloot, R., An updated roadmap for the

integration of metal–organic frameworks with electronic devices and chemical sensors. Chem. Soc. Rev.

2017, 46 (11), 3185-3241.

45. Dolgopolova, E. A.; Galitskiy, V. A.; Martin, C. R.; Gregory, H. N.; Yarbrough, B. J.; Rice, A. M.;

Berseneva, A. A.; Ejegbavwo, O. A.; Stephenson, K. S.; Kittikhunnatham, P., Connecting Wires:

Photoinduced Electronic Structure Modulation in Metal–Organic Frameworks. J. Am. Chem. Soc. 2019,

141 (13), 5350-5358.

46. Kurmoo, M., Magnetic metal–organic frameworks. Chem. Soc. Rev. 2009, 38 (5), 1353-1379.

47. Shao, Y.; Zhou, L.; Bao, C.; Ma, J.; Liu, M.; Wang, F., Magnetic responsive metal–organic frameworks

nanosphere with core–shell structure for highly efficient removal of methylene blue. Chem. Eng. J. 2016,

283, 1127-1136.

48. Espallargas, G. M.; Coronado, E., Magnetic functionalities in MOFs: from the framework to the pore.

Chem. Soc. Rev. 2018, 47 (2), 533-557.

49. Betard, A.; Fischer, R. A., Metal–organic framework thin films: from fundamentals to applications.

Chem. Rev. 2012, 112 (2), 1055-1083.

50. Shekhah, O.; Wang, H.; Kowarik, S.; Schreiber, F.; Paulus, M.; Tolan, M.; Sternemann, C.; Evers, F.;

Zacher, D.; Fischer, R. A., Step-by-step route for the synthesis of metal− organic frameworks. J. Am.

Chem. Soc. 2007, 129 (49), 15118-15119.

51. Shekhah, O.; Wang, H.; Strunskus, T.; Cyganik, P.; Zacher, D.; Fischer, R.; Wöll, C., Layer-by-layer

growth of oriented metal organic polymers on a functionalized organic surface. Langmuir 2007, 23 (14),

7440-7442.

52. Eddaoudi, M.; Kim, J.; Rosi, N.; Vodak, D.; Wachter, J.; O'Keeffe, M.; Yaghi, O. M., Systematic design

of pore size and functionality in isoreticular MOFs and their application in methane storage. Science

2002, 295 (5554), 469-472.

53. Shekhah, O.; Eddaoudi, M., The liquid phase epitaxy method for the construction of oriented ZIF-8 thin

films with controlled growth on functionalized surfaces. Chem. Commun. 2013, 49 (86), 10079-10081.

54. Jia, H.; Yao, Y.; Zhao, J.; Gao, Y.; Luo, Z.; Du, P., A novel two-dimensional nickel phthalocyanine-

based metal–organic framework for highly efficient water oxidation catalysis. J. Mater. Chem. A 2018,

6 (3), 1188-1195.

55. Nagatomi, H.; Yanai, N.; Yamada, T.; Shiraishi, K.; Kimizuka, N., Synthesis and Electric Properties of

a Two‐Dimensional Metal‐Organic Framework Based on Phthalocyanine. Chem. Eur. J 2018, 24 (8),

1806-1810.

56. Meng, Z.; Aykanat, A.; Mirica, K. A., Welding metallophthalocyanines into bimetallic molecular meshes

for ultrasensitive, low-power chemiresistive detection of gases. J. Am. Chem. Soc. 2018, 141 (5), 2046-

2053.

57. Munuera, C.; Shekhah, O.; Wang, H.; Wöll, C.; Ocal, C., The controlled growth of oriented metal–

organic frameworks on functionalized surfaces as followed by scanning force microscopy. Phys. Chem.

Chem. Phys. 2008, 10 (48), 7257-7261.

58. Petkov, P. S.; Vayssilov, G. N.; Liu, J.; Shekhah, O.; Wang, Y.; Wöll, C.; Heine, T., Defects in MOFs:

a thorough characterization. ChemPhysChem 2012, 13 (8), 2025-2029.

59. Wang, Z.; Liu, J.; Lukose, B.; Gu, Z.; Weidler, P. G.; Gliemann, H.; Heine, T.; Woll, C., Nanoporous

designer solids with huge lattice constant gradients: multiheteroepitaxy of metal–organic frameworks.

Nano Lett. 2014, 14 (3), 1526-1529.

60. Liu, J.; Lukose, B.; Shekhah, O.; Arslan, H. K.; Weidler, P.; Gliemann, H.; Bräse, S.; Grosjean, S.; Godt,

A.; Feng, X., A novel series of isoreticular metal organic frameworks: realizing metastable structures by

liquid phase epitaxy. Sci. Rep. 2012, 2, 921.

Page 154: Theoretical Investigations of the Photophysical Properties

148

61. Bae, T. H.; Lee, J. S.; Qiu, W.; Koros, W. J.; Jones, C. W.; Nair, S., A high‐performance gas‐separation

membrane containing submicrometer‐sized metal–organic framework crystals. Angew. Chem. Int. Ed.

2010, 122 (51), 10059-10062.

62. Talin, A. A.; Centrone, A.; Ford, A. C.; Foster, M. E.; Stavila, V.; Haney, P.; Kinney, R. A.; Szalai, V.;

El Gabaly, F.; Yoon, H. P., Tunable electrical conductivity in metal-organic framework thin-film devices.

Science 2014, 343 (6166), 66-69.

63. Kreno, L. E.; Hupp, J. T.; Van Duyne, R. P., Metal− organic framework thin film for enhanced localized

surface plasmon resonance gas sensing. Anal. Chem. 2010, 82 (19), 8042-8046.

64. Abrahams, B. F.; Hoskins, B. F.; Robson, R., A new type of infinite 3D polymeric network containing

4-connected, peripherally-linked metalloporphyrin building blocks. J. Am. Chem. Soc. 1991, 113 (9),

3606-3607.

65. Ohmura, T.; Usuki, A.; Fukumori, K.; Ohta, T.; Ito, M.; Tatsumi, K., New Porphyrin-Based Metal−

Organic Framework with High Porosity: 2-D Infinite 22.2-Å Square-Grid Coordination Network. Inorg.

Chem. 2006, 45 (20), 7988-7990.

66. Gao, W.-Y.; Chrzanowski, M.; Ma, S., Metal–metalloporphyrin frameworks: a resurging class of

functional materials. Chem. Soc. Rev. 2014, 43 (16), 5841-5866.

67. Stassen, I.; Burtch, N.; Talin, A.; Falcaro, P.; Allendorf, M.; Ameloot, R., An updated roadmap for the

integration of metal–organic frameworks with electronic devices and chemical sensors. Chem. Soc. Rev

2017, 46 (11), 3185-3241.

68. Liu, J.; Wöll, C., Surface-supported metal–organic framework thin films: fabrication methods,

applications, and challenges. Chem. Soc. Rev 2017, 46 (19), 5730-5770.

69. Makiura, R.; Motoyama, S.; Umemura, Y.; Yamanaka, H.; Sakata, O.; Kitagawa, H., Surface nano-

architecture of a metal–organic framework. Nat. Mater. 2010, 9 (7), 565-571.

70. Motoyama, S.; Makiura, R.; Sakata, O.; Kitagawa, H., Highly crystalline nanofilm by layering of

porphyrin metal− organic framework sheets. J. Am. Chem. Soc. 2011, 133 (15), 5640-5643.

71. So, M. C.; Jin, S.; Son, H.-J.; Wiederrecht, G. P.; Farha, O. K.; Hupp, J. T., Layer-by-layer fabrication

of oriented porous thin films based on porphyrin-containing metal–organic frameworks. J. Am. Chem.

Soc. 2013, 135 (42), 15698-15701.

72. Liu, J.; Zhou, W.; Liu, J.; Howard, I.; Kilibarda, G.; Schlabach, S.; Coupry, D.; Addicoat, M.; Yoneda,

S.; Tsutsui, Y. S., Tsuneaki; Seki, S.; Wang, Z.; Lindemann, P.; Redel, E.; Heine, T.; Wöll, C.,

Photoinduced Charge‐Carrier Generation in Epitaxial MOF Thin Films: High Efficiency as a Result of

an Indirect Electronic Band Gap? Angew. Chem. Int. Ed. 2015, 54 (25), 7441-7445.

73. Liu, J.; Zhou, W.; Liu, J.; Fujimori, Y.; Higashino, T.; Imahori, H.; Jiang, X.; Zhao, J.; Sakurai, T.;

Hattori, Y.; Matsuda, W.; Seki, S.; Garlapati, S. K.; Dasgupta, S.; Redel, E.; Sun, L.; Wöll, C., A new

class of epitaxial porphyrin metal–organic framework thin films with extremely high photocarrier

generation efficiency: promising materials for all-solid-state solar cells. J. Mater. Chem. A 2016, 4 (33),

12739-12747.

74. Schrödinger, E., An undulatory theory of the mechanics of atoms and molecules. Phys. Rev. 1926, 28

(6), 1049.

75. Born, M.; Oppenheimer, R., Zur quantentheorie der molekeln. Ann. Phys. 1927, 389 (20), 457-484.

76. Hartree, D. In The Wave Mechanics of an Atom with a Non-Coulomb Central Field. Part I. Theory and

Methods, Math. Proc. Cambridge Philos. Soc., Cambridge University Press: 1928; pp 89-110.

77. Parr, R. G., Density functional theory of atoms and molecules. In Horizons of quantum chemistry,

Springer: 1980; pp 5-15.

78. Slater, J. C., The theory of complex spectra. Phys. Rev. 1929, 34 (10), 1293.

79. Hohenberg, P.; Kohn, W., Inhomogeneous electron gas. Phys. Rev. 1964, 136 (3B), B864.

80. Kohn, W.; Sham, L. J., Self-consistent equations including exchange and correlation effects. Phys. Rev.

1965, 140 (4A), A1133.

81. Schmider, H.; Becke, A., Chemical content of the kinetic energy density. J. Mol. Struct. (Theochem)

2000, 527 (1-3), 51-61.

Page 155: Theoretical Investigations of the Photophysical Properties

149

82. Mori-Sánchez, P.; Cohen, A.; Yang, W., Many-electron self-interaction error in approximate density

functionals. J. Chem. Phys. 2006, 125 (20), 201102.

83. Becke, A. D., A new mixing of Hartree–Fock and local density‐functional theories. J. Chem. Phys. 1993,

98 (2), 1372-1377.

84. Harris, J.; Jones, R., The surface energy of a bounded electron gas. J. Phys. F: Metal Phys. 1974, 4 (8),

1170.

85. Gunnarsson, O.; Lundqvist, B. I., Exchange and correlation in atoms, molecules, and solids by the spin-

density-functional formalism. Phys. Rev. B 1976, 13 (10), 4274.

86. Langreth, D. C.; Perdew, J. P., Exchange-correlation energy of a metallic surface: Wave-vector analysis.

Phys. Rev. B 1977, 15 (6), 2884.

87. Harris, J., Adiabatic-connection approach to Kohn-Sham theory. Phys. Rev. A 1984, 29 (4), 1648.

88. Mardirossian, N.; Head-Gordon, M., Thirty years of density functional theory in computational

chemistry: an overview and extensive assessment of 200 density functionals. Mol. Phys. 2017, 115 (19),

2315-2372.

89. Casida, M. E.; Huix-Rotllant, M., Progress in time-dependent density-functional theory. Annu. Rev. Phys.

Chem. 2012, 63, 287-323.

90. Yabana, K.; Bertsch, G., Time-dependent local-density approximation in real time. Phys. Rev. B 1996,

54 (7), 4484.

91. Marques, M. A.; Castro, A.; Bertsch, G. F.; Rubio, A., octopus: a first-principles tool for excited

electron–ion dynamics. Comput. Phys. Commun. 2003, 151 (1), 60-78.

92. Casida, M. E., Time-Dependent Density Functional Response Theory for Molecules. In Recent Advances

In Density Functional Methods: (Part I), World Scientific: 1995; p 155.

93. Runge, E.; Gross, E. K., Density-Functional Theory for Time-Dependent Systems. Phys. Rev. Lett. 1984,

52, 997.

94. Paier, J.; Marsman, M.; Kresse, G., Dielectric properties and excitons for extended systems from hybrid

functionals. Phys. Rev. B 2008, 78 (12), 121201.

95. Castro, A.; Marques, M. A.; Alonso, J. A.; Bertsch, G. F.; Yabana, K.; Rubio, A., Can optical

spectroscopy directly elucidate the ground state of C 20? J. Chem. Phys. 2002, 116 (5), 1930-1933.

96. Marques, M. A.; López, X.; Varsano, D.; Castro, A.; Rubio, A., Time-dependent density-functional

approach for biological chromophores: the case of the green fluorescent protein. Phys. Rev. Lett. 2003,

90 (25), 258101.

97. Dreuw, A.; Head-Gordon, M., Single-reference ab initio methods for the calculation of excited states of

large molecules. Chem. Rev. 2005, 105 (11), 4009-4037.

98. Hirata, S.; Head-Gordon, M., Time Dependent Density Functional Theory within the Tamm–Dancoff

Approximation. Chem. Phys. Lett. 1999, 314, 291.

99. Sears, J. S.; Koerzdoerfer, T.; Zhang, C.-R.; Brédas, J.-L., Communication: Orbital instabilities and

triplet states from time-dependent density functional theory and long-range corrected functionals. J.

Chem. Phys. 2011, 135 (15), 151103.

100. Furche, F., On the density matrix based approach to time-dependent density functional response theory.

J. Chem. Phys. 2001, 114 (14), 5982-5992.

101. Cai, Z.-L.; Sendt, K.; Reimers, J. R., Failure of density-functional theory and time-dependent density-

functional theory for large extended π systems. J. Chem. Phys 2002, 117 (12), 5543-5549.

102. Dreuw, A.; Weisman, J. L.; Head-Gordon, M., Long-range charge-transfer excited states in time-

dependent density functional theory require non-local exchange. J. Chem. Phys 2003, 119 (6), 2943-

2946.

103. Dreuw, A.; Head-Gordon, M., Failure of Time-Dependent Density Functional Theory for Long-Range

Charge-Transfer Excited States: The zincbacteriochlorin-bacteriochlorin and bacteriochlorophyll-

spheroidene complexes. J. Am. Chem. Soc. 2004, 126, 4007.

Page 156: Theoretical Investigations of the Photophysical Properties

150

104. Grimme, S.; Parac, M., Substantial errors from time‐dependent density functional theory for the

calculation of excited states of large π systems. ChemPhysChem 2003, 4 (3), 292-295.

105. Laurent, A. D.; Jacquemin, D., TD‐DFT benchmarks: a review. Int. J. Quantum Chem 2013, 113 (17),

2019-2039.

106. Grimme, S., A simplified Tamm-Dancoff Density Functional Approach for the Electronic Excitation

Spectra of Very Large Molecules. J. Chem. Phys. 2013, 138, 244104.

107. Bannwarth, C.; Grimme, S., A simplified Time Dependent Density Functional Theory Approach for

Electronic Ultraviolet and Circular Dichroism Spectra of Very Large Molecules. Comput. Theor. Chem.

2014, 1040, 45.

108. Lowdin, P. O., On the non‐orthogonality problem connected with the use of atomic wave functions in

the theory of molecules and crystals. J. Chem. Phys. 1950, 18 (3), 365-375.

109. Nishimoto, K.; Mataga, N., Electronic structure and spectra of some nitrogen heterocycles. Z. Phys.

Chem. 1957, 12 (5_6), 335-338.

110. Ohno, K., Some remarks on the Pariser-Parr-Pople method. Theor. Chim. Acta 1964, 2 (3), 219-227.

111. Klopman, G., A semiempirical treatment of molecular structures. II. Molecular terms and application to

diatomic molecules. J. Am. Chem. Soc. 1964, 86 (21), 4550-4557.

112. Seifert, G.; Eschrig, H.; Bieger, W., An approximation variant of LCAO-X-ALPHA methods. Z. Phys.

Chem. 1986, 267 (3), 529-539.

113. Foulkes, W. M. C.; Haydock, R., Tight-binding models and density-functional theory. Phys. Rev. B 1989,

39 (17), 12520.

114. Seifert, G., Tight-binding density functional theory: an approximate Kohn− Sham DFT scheme. J. Phys.

Chem. A 2007, 111 (26), 5609-5613.

115. Elstner, M.; Porezag, D.; Jungnickel, G.; Elsner, J.; Haugk, M.; Frauenheim, T.; Suhai, S.; Seifert, G.,

Self-consistent-charge density-functional tight-binding method for simulations of complex materials

properties. Phys. Rev. B 1998, 58 (11), 7260.

116. Köhler, C.; Seifert, G.; Gerstmann, U.; Elstner, M.; Overhof, H.; Frauenheim, T., Approximate density-

functional calculations of spin densities in large molecular systems and complex solids. Phys. Chem.

Chem. Phys. 2001, 3 (23), 5109-5114.

117. Niehaus, T. A.; Suhai, S.; Della Sala, F.; Lugli, P.; Elstner, M.; Seifert, G.; Frauenheim, T., Tight-binding

approach to time-dependent density-functional response theory. Phys. Rev. B 2001, 63 (8), 085108.

118. Elstner, M.; Frauenheim, T.; Kaxiras, E.; Seifert, G.; Suhai, S., A self‐consistent charge density‐

functional based tight‐binding scheme for large biomolecules. Phys. Status Solidi B 2000, 217 (1), 357-

376.

119. Frauenheim, T.; Seifert, G.; Elsterner, M.; Hajnal, Z.; Jungnickel, G.; Porezag, D.; Suhai, S.; Scholz, R.,

A self‐consistent charge density‐functional based tight‐binding method for predictive materials

simulations in physics, chemistry and biology. Phys. Status Solidi B 2000, 217 (1), 41-62.

120. Mulliken, R. S., Electronic population analysis on LCAO–MO molecular wave functions. I. J. Chem.

Phys 1955, 23 (10), 1833-1840.

121. Grimm, U., Aperiodic crystals and beyond. Acta Cryst. B 2015, 71 (3), 258-274.

122. Monkhorst, H. J.; Pack, J. D., Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13 (12),

5188.

123. Bloch, F., Über die quantenmechanik der elektronen in kristallgittern. Z. Phys 1929, 52 (7-8), 555-600.

124. Te Velde, G. t.; Bickelhaupt, F. M.; Baerends, E. J.; Fonseca Guerra, C.; van Gisbergen, S. J.; Snijders,

J. G.; Ziegler, T., Chemistry with ADF. J. Comput. Chem. 2001, 22 (9), 931-967.

125. Addicoat, M. A.; Coupry, D. E.; Heine, T., AuToGraFS: automatic topological generator for framework

structures. J. Phys. Chem. A 2014, 118 (40), 9607-9614.

126. Rappé, A. K.; Casewit, C. J.; Colwell, K.; Goddard III, W. A.; Skiff, W. M., UFF, a full periodic table

force field for molecular mechanics and molecular dynamics simulations. J. Am. Chem. Soc. 1992, 114

(25), 10024-10035.

Page 157: Theoretical Investigations of the Photophysical Properties

151

127. R. Rüger, A. Y., P. Philipsen, S. Borini, P. Melix, A. F. Oliveira, M. Franchini, T. van Vuren, T. Soini,

M. de Reus, M. Ghorbani Asl, T. Q. Teodoro, D. McCormack, S. Patchkovskii, T. Heine, AMS DFTB

2019.3, SCM, Theoretical Chemistry, Vrije Universiteit, Amsterdam, The Netherlands,

http://www.scm.com.

128. Ahlrichs, R.; Bär, M.; Häser, M.; Horn, H.; Kölmel, C., Electronic Structure Calculations on Workstation

Computers: The Program System Turbomole. Chem. Phys. Lett. 1989, 162, 165.

129. Neese, F., The ORCA Program System. WIREs: Comput Mol Sci 2012, 2, 73.

130. Dovesi, R.; Erba, A.; Orlando, R.; Zicovich‐Wilson, C. M.; Civalleri, B.; Maschio, L.; Rérat, M.;

Casassa, S.; Baima, J.; Salustro, S., Quantum‐mechanical condensed matter simulations with CRYSTAL.

Wiley Interdiscip. Rev. Comput. Mol. Sci. 2018, 8 (4), e1360.

131. Li, L.-L.; Diau, E. W.-G., Porphyrin-Sensitized Solar Cells. Chem. Soc. Rev. 2013, 42, 291.

132. Ishihara, S.; Labuta, J.; Van Rossom, W.; Ishikawa, D.; Minami, K.; Hill, J. P.; Ariga, K., Porphyrin-

Based Sensor Nanoarchitectonics in Diverse Physical Detection Modes. Phys. Chem. Chem. Phys. 2014,

16, 9713.

133. Xu, J.; Wu, J.; Zong, C.; Ju, H.; Yan, F., Manganese Porphyrin-dsDNA Complex: A Mimicking Enzyme

for Highly Efficient Bioanalysis. Anal. Chem. 2013, 85, 3374.

134. Bonnett, R., Photosensitizers of the Porphyrin and Phthalocyanine Series for Photodynamic Therapy.

Chem. Soc. Rev. 1995, 24, 19.

135. Haldar, R.; Batra, K.; Marschner, S. M.; Kuc, A. B.; Zahn, S.; Fischer, R. A.; Bräse, S.; Heine, T.; Wöll,

C., Bridging the Green Gap: Metal–Organic Framework Heteromultilayers Assembled from Porphyrinic

Linkers Identified by Using Computational Screening. Chem. Eur. J 2019, 25, 7847-7851.

136. Kandambeth, S.; Shinde, D. B.; Panda, M. K.; Lukose, B.; Heine, T.; Banerjee, R., Enhancement of

chemical stability and crystallinity in porphyrin‐containing covalent organic frameworks by

intramolecular hydrogen bonds. Angew. Chem. Int. Ed. 2013, 52 (49), 13052-13056.

137. Christiansen, O.; Koch, H.; Jørgensen, P., The Second-Order Approximate Coupled Cluster Singles and

Doubles Model CC2. Chem. Phys. Lett. 1995, 243, 409.

138. Hättig, C.; Weigend, F., CC2 Excitation Energy Calculations on Large Molecules Using the Resolution

of the Identity Approximation. J. Chem. Phys. 2000, 113, 5154.

139. Schreiber, M.; Silva-Junior, M. R.; Sauer, S. P.; Thiel, W., Benchmarks for Electronically Excited States:

CASPT2, CC2, CCSD, and CC3. J. Chem. Phys. 2008, 128, 134110.

140. Schirmer, J., Beyond the random-phase approximation: A new approximation scheme for the

polarization propagator. Phys. Rev. A 1982, 26 (5), 2395.

141. Andersson, K.; Malmqvist, P. Å.; Roos, B. O., Second‐Order Perturbation Theory with a Complete

Active Space Self‐Consistent Field Reference Function. J. Chem. Phys. 1992, 96, 1218.

142. Serrano-Andrés, L.; Merchán, M.; Rubio, M.; Roos, B. O., Interpretation of the electronic absorption

spectrum of free base porphin by using multiconfigurational second-order perturbation theory. Chem.

Phys. Lett. 1998, 295, 195.

143. Tokita, Y.; Hasegawa, J.; Nakatsuji, H., SAC-CI Study on the Excited and Ionized States of Free-Base

Porphin: Rydberg Excited States and Effect of Polarization and Rydberg Functions. J. Phys. Chem. A

1998, 102, 1843.

144. Gwaltney, S. R.; Bartlett, R. J., Coupled-Cluster Calculations of the Electronic Excitation Spectrum of

Free Base Porphin in a Polarized Basis. J. Chem. Phys. 1998, 108, 6790.

145. Angeli, C.; Cimiraglia, R.; Evangelisti, S.; Leininger, T.; Malrieu, J.-P., Introduction of n-Electron

Valence States for Multireference Perturbation Theory. J. Chem. Phys. 2001, 114, 10252.

146. Angeli, C.; Pastore, M.; Cimiraglia, R., New perspectives in Multireference Perturbation Theory: The n-

Electron Valence State Approach. Theor. Chem. Acc. 2007, 117, 743.

147. Gross, E.; Kohn, W., Time-dependent density-functional theory. In Advances in quantum chemistry,

Elsevier: 1990; Vol. 21, pp 255-291.

Page 158: Theoretical Investigations of the Photophysical Properties

152

148. Van Leeuwen, R., Key concepts in time-dependent density-functional theory. Int. J. Mod. Phys. A 2001,

15, 1969-2023.

149. Marques, M. A.; Gross, E. K., Time-dependent density functional theory. Annu. Rev. Phys. Chem. 2004,

55, 427-455.

150. Casida, M. E., Time-dependent density-functional theory for molecules and molecular solids. J. Mol.

Struct. (Theochem) 2009, 914 (1-3), 3-18.

151. Adamo, C.; Jacquemin, D., The calculations of excited-state properties with Time-Dependent Density

Functional Theory. Chem. Soc. Rev. 2013, 42 (3), 845-856.

152. Chantzis, A.; Laurent, A. D.; Adamo, C.; Jacquemin, D., Is the Tamm-Dancoff Approximation Reliable

for the Calculation of Absorption and Fluorescence Band Shapes? J. Chem. Theory. Comput. 2013, 9,

4517.

153. Wang, Y. L.; Wu, G. S., Improving the TDDFT Calculation of Low‐Lying Excited States for Polycyclic

Aromatic Hydrocarbons using the Tamm–Dancoff approximation. Int. J. Quantum Chem 2008, 108, 430.

154. Tozer, D. J.; Amos, R. D.; Handy, N. C.; Roos, B. O.; Serrano-Andrés, L., Does Density Functional

Theory Contribute to the Understanding of Excited States of Unsaturated Organic Compounds? Mol.

Phys. 1999, 97, 859.

155. Bauernschmitt, R.; Ahlrichs, R., Treatment of Electronic Excitations within the Adiabatic Approximation

of Time Dependent Density Functional Theory. Chem. Phys. Lett 1996, 256, 454.

156. Tian, B.; Eriksson, E. S.; Eriksson, L. A., Can Range-Separated and Hybrid DFT Functionals Predict

Low-Lying Excitations? A Tookad Case Study. J. Chem. Theory Comput. 2010, 6, 2086.

157. Eriksson, E. S.; Eriksson, L. A., Predictive Power of Long-Range Corrected Functionals on the

Spectroscopic Properties of Tetrapyrrole Derivatives for Photodynamic Therapy. Phys. Chem. Chem.

Phys. 2011, 13, 7207.

158. Lee, M.-J.; Balanay, M. P.; Kim, D. H., Molecular Design of Distorted Push–Pull Porphyrins for Dye-

Sensitized Solar Cells. Theor. Chem. Acc 2012, 131, 1269.

159. Winter, N. O.; Graf, N. K.; Leutwyler, S.; Hättig, C., Benchmarks for 0–0 Transitions of Aromatic

Organic Molecules: DFT/B3LYP, ADC (2), CC2, SOS-CC2 and SCS-CC2 Compared to High-

Resolution Gas-Phase Data. Phys. Chem. Chem. Phys. 2013, 15, 6623.

160. Fang, C.; Oruganti, B.; Durbeej, B., How Method-Dependent are Calculated Differences Between

Vertical, Adiabatic, and 0–0 Excitation Energies? J. Phys. Chem. A 2014, 118, 4157.

161. Send, R.; Kuhn, M.; Furche, F., Assessing Excited State Methods by Adiabatic Excitation Energies. J.

Chem. Theory Comput. 2011, 7, 2376.

162. Theisen, R. F.; Huang, L.; Fleetham, T.; Adams, J. B.; Li, J., Ground and Excited states of Zinc

Phthalocyanine, Zinc Tetrabenzoporphyrin, and Azaporphyrin Analogs using DFT and TDDFT with

Franck-Condon Analysis. J. Chem. Phys. 2015, 142, 094310.

163. Grimme, S.; Bannwarth, C., Ultra-fast computation of electronic spectra for large systems by tight-

binding based simplified Tamm-Dancoff approximation (sTDA-xTB). J. Chem. Phys. 2016, 145 (5).

164. Becke, A. D., Density Functional Exchange Energy Approximation with Correct Asymptotic Behavior.

Phys. Rev. A 1988, 38, 3098.

165. Lee, C.; Yang, W.; Parr, R. G., Development of the Colle-Salvetti Correlation Energy Formula into a

Functional of the Electron Density. Phys. Rev. B 1988, 37, 785.

166. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H., A Consistent and Accurate ab initio Parametrization of

Density Functional Dispersion Correction (DFT-D) for the 94 Elements H-Pu. J. Chem. Phys 2010, 132,

154104.

167. Baerends, E.; Ellis, D.; Ros, P., Self-Consistent Molecular Hartree-Fock-Slater Calculations I. The

Computational Procedure. Chem. Phys. 1973, 2, 41.

168. Dunlap, B. I.; Connolly, J.; Sabin, J., On Some Approximations in Applications of X α Theory. J. Chem.

Phys. 1979, 71, 3396.

Page 159: Theoretical Investigations of the Photophysical Properties

153

169. Eichkorn, K.; Weigend, F.; Treutler, O.; Ahlrichs, R., Auxiliary Basis Sets for Main Row Atoms and

Transition Metals and their Use to Approximate Coulomb Potentials. Theor. Chem. Acc 1997, 97, 119.

170. Schäfer, A.; Huber, C.; Ahlrichs, R., Fully Optimized Contracted Gaussian Basis Sets of Triple Zeta

Valence Quality for Atoms Li to Kr. J. Chem. Phys. 1994, 100, 5829.

171. Becke, A. D., Density Functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys.

1993, 98, 5648.

172. Perdew, J. P., Density Functional Approximation for the Correlation Energy of the Inhomogeneous

Electron Gas. Phys. Rev. B 1986, 33, 8822.

173. Perdew, J. P.; Burke, K.; Ernzerhof, M., Generalized Gradient Approximation Made Simple. Phys. Rev.

Lett. 1996, 77, 3865.

174. Perdew, J. P.; Wang, Y., Accurate and Simple Analytic Representation of the Electron-Gas Correlation

Energy. Phys. Rev. B 1992, 45, 13244.

175. Tao, J.; Perdew, J. P.; Staroverov, V. N.; Scuseria, G. E., Climbing the Density Functional Ladder:

Nonempirical meta–Generalized Gradient Approximation Designed for Molecules and Solids. Phys. Rev.

Lett. 2003, 91, 146401.

176. Zhao, Y.; Truhlar, D. G., A New Local Density Functional for Main-Group Thermochemistry, Transition

Metal Bonding, Thermochemical Kinetics, and Non-Covalent Interactions. J. Chem. Phys. 2006, 125,

194101.

177. Perdew, J. P.; Ernzerhof, M.; Burke, K., Rationale for Mixing Exact Exchange with Density Functional

Approximations. J. Chem. Phys. 1996, 105, 9982.

178. Becke, A. D., A New Mixing of Hartree–Fock and Local Density Functional Theories. J. Chem. Phys.

1993, 98, 1372.

179. Tao, J.; Perdew, J. P.; Staroverov, V. N.; Scuseria, G. E., Climbing the Density Functional Ladder: Non-

Empirical meta-Generalized Gradient Approximation Designed for Molecules and Solids. Phys. Rev.

Lett. 2003, 91, 146401.

180. Zhao, Y.; Truhlar, D. G., The M06 suite of Density Functionals for Main Group Thermochemistry,

Thermochemical Kinetics, Non-Covalent Interactions, Excited States, and Transition Elements: Two

New Functionals and Systematic Testing of Four M06-Class Functionals and 12 Other Functionals.

Theor. Chem. Acc. 2008, 120, 215.

181. Chai, J.-D.; Head-Gordon, M., Systematic Optimization of Long-Range Corrected Hybrid Density

Functionals. J. Chem. Phys. 2008, 128, 084106.

182. Tawada, Y.; Tsuneda, T.; Yanagisawa, S.; Yanai, T.; Hirao, K., A Long Range Corrected Time

Dependent Density Functional Theory. J. Chem. Phys. 2004, 120, 8425.

183. Yanai, T.; Tew, D. P.; Handy, N. C., A New Hybrid Exchange–Correlation Functional Using the

Coulomb Attenuating Method (CAM-B3LYP). Chem. Phys. Lett. 2004, 393, 51.

184. Grimme, S., Semiempirical Hybrid Density Functional with Perturbative Second-Order Correlation. J.

Chem. Phys. 2006, 124, 034108.

185. Karton, A.; Tarnopolsky, A.; Lamére, J.-F.; Schatz, G. C.; Martin, J. M., Highly Accurate First-Principles

Benchmark Data-Sets for the Parametrization and Validation of Density Functional and Other

Approximate Methods. Derivation of a Robust, Generally Applicable, Double-Hybrid Functional for

Thermochemistry and Thermochemical Kinetics. J. Phys. Chem. A 2008, 112, 12868.

186. Schwabe, T.; Grimme, S., Towards Chemical Accuracy for the Thermodynamics of Large Molecules:

New Hybrid Density Functionals Including Non-Local Correlation Effects. Phys. Chem. Chem. Phys.

2006, 8, 4398.

187. Weigend, F.; Ahlrichs, R., Balanced Basis Sets of Split Valence, Triple Zeta Valence and Quadruple

Zeta Valence Quality for H to Rn: Design and Assessment of Accuracy. Phys. Chem. Chem. Phys. 2005,

7, 3297.

188. Rappoport, D.; Furche, F., Property-optimized Gaussian basis sets for molecular response calculations.

J. Chem. Phys. 2010, 133 (13), 134105.

Page 160: Theoretical Investigations of the Photophysical Properties

154

189. Feyereisen, M.; Fitzgerald, G.; Komornicki, A., Use of Approximate Integrals in Abinitio Theory - An

Application in MP2 Energy Calculations. Chem. Phys. Lett. 1993, 208, 359.

190. Weigend, F.; Haser, M.; Patzelt, H.; Ahlrichs, R., RI-MP2: Optimized Auxiliary Basis Sets and

Demonstration of Efficiency. Chem. Phys. Lett. 1998, 294, 143.

191. Neese, F.; Wennmohs, F.; Hansen, A.; Becker, U., Efficient, approximate and parallel Hartree–Fock and

hybrid DFT calculations. A ‘chain-of-spheres’ algorithm for the Hartree–Fock exchange. Chem. Phys.

2009, 356 (1-3), 98-109.

192. Foresman, J. B.; Head-Gordon, M.; Pople, J. A.; Frisch, M. J., Toward a Systematic Molecular-Orbital

Theory for Excited-States. J. Phys. Chem. 1992, 96, 135.

193. Head-Gordon, M.; Rico, R. J.; Oumi, M.; Lee, T. J., A Doubles Correction to Electronic Excited-States

from Configuration-Interaction in the Space of Single Substitutions. Chem. Phys. Lett. 1994, 219, 21.

194. Head-Gordon, M.; Maurice, D.; Oumi, M., A Perturbative Correction to Restricted Open-Shell

Configuration-Interaction with Single Substitutions for Excited-States of Radicals. Chem. Phys. Lett.

1995, 246, 114.

195. Eisner, U.; Linstead, R., Chlorophyll and related substances. Part II. The dehydrogenation of chlorin to

prophin and the number of extra hydrogen atoms in the chlorins. J. Chem. Soc. (Resumed) 1955, 3749-

3754.

196. Du, H.; Fuh, R. C. A.; Li, J.; Corkan, L. A.; Lindsey, J. S., PhotochemCAD: A computer‐aided design

and research tool in photochemistry. Photochem. Photobiol 1998, 68, 141-142.

197. Dixon, J. M.; Taniguchi, M.; Lindsey, J. S., PhotochemCAD 2: a refined program with accompanying

spectral databases for photochemical calculations. Photochem. Photobiol. 2005, 81, 212-213.

198. Eisner, U.; Lichtarowicz, A.; Linstead, R., 142. Chlorophyll and related compounds. Part VI. The

synthesis of octaethylchlorin. J. Chem. Soc. (Resumed) 1957, 733-739.

199. Zass, E.; Isenring, H. P.; Etter, R.; Eschenmoser, A., Der Einbau von Magnesium in Liganden der

Chlorophyll‐Reihe mit (2, 6‐Di‐t‐butyl‐4‐methylphenoxy) magnesiumjodid. Helv. Chim. Acta 1980, 63,

1048-1067.

200. Buchler, J. W.; Puppe, L., Metallkomplexe mit Tetrapyrrol‐Liganden, II1) Metallchelate des α. γ‐

Dimethyl‐α. γ‐dihydro‐octaäthylporphins durch reduzierende Methylierung von Octaäthylporphinato‐

zink. Eur. J. Org. Chem. 1970, 740, 142-163.

201. Barnett, G. H.; Hudson, M. F.; Smith, K. M., Concerning meso-tetraphenylporphyrin purification. J.

Chem. Soc., Perkin Trans. 1 1975, (14), 1401-1403.

202. Miller, J.; Dorough, G., Pyridinate Complexes of Some Metallo-derivatives of Tetraphenylporphine and

Tetraphenylchlorin1. J. Am. Chem. Soc. 1952, 74, 3977-3981.

203. Collman, J. P.; Gagne, R. R.; Reed, C.; Halbert, T. R.; Lang, G.; Robinson, W. T., Picket fence

porphyrins. Synthetic models for oxygen binding hemoproteins. J. Am. Chem. Soc. 1975, 97, 1427-1439.

204. Segawa, H.; Takehara, C.; Honda, K.; Shimidzu, T.; Asahi, T.; Mataga, N., Photoinduced electron-

transfer reactions of porphyrin heteroaggregates: energy gap dependence of an intradimer charge

recombination process. J. Phys. Chem. A 1992, 96, 503-506.

205. Son, H.-J.; Jin, S.; Patwardhan, S.; Wezenberg, S. J.; Jeong, N. C.; So, M.; Wilmer, C. E.; Sarjeant, A.

A.; Schatz, G. C.; Snurr, R. Q., Light-harvesting and ultrafast energy migration in porphyrin-based

metal–organic frameworks. J. Am. Chem. Soc. 2013, 135, 862-869.

206. Bhyrappa, P.; Krishnan, V., Octabromotetraphenylporphyrin and its metal derivatives: electronic

structure and electrochemical properties. Inorg. Chem. 1991, 30, 239.

207. Iikura, H.; Tsuneda, T.; Yanai, T.; Hirao, K., A long-range correction scheme for generalized-gradient-

approximation exchange functionals. J. Chem. Phys 2001, 115 (8), 3540-3544.

208. Gill, P. M.; Adamson, R. D.; Pople, J. A., Coulomb-attenuated exchange energy density functionals. Mol.

Phys. 1996, 88 (4), 1005-1009.

209. Leininger, T.; Stoll, H.; Werner, H.-J.; Savin, A., Combining long-range configuration interaction with

short-range density functionals. Chem. Phys. Lett 1997, 275 (3-4), 151-160.

Page 161: Theoretical Investigations of the Photophysical Properties

155

210. Risthaus, T.; Hansen, A.; Grimme, S., Excited states using the simplified Tamm-Dancoff-Approach for

range-separated hybrid density functionals: development and application. Phys Chem Chem Phys 2014,

16 (28), 14408-14419.

211. Casida, M. E.; Jamorski, C.; Casida, K. C.; Salahub, D. R., Molecular excitation energies to high-lying

bound states from time-dependent density-functional response theory: Characterization and correction

of the time-dependent local density approximation ionization threshold. J. Chem. Phys. 1998, 108 (11),

4439-4449.

212. Haug, F.-J.; Ballif, C., Light management in thin film silicon solar cells. Energy Environ Sci. 2015, 8

(3), 824-837.

213. Gao, P.; Yang, Z.; He, J.; Yu, J.; Liu, P.; Zhu, J.; Ge, Z.; Ye, J., Dopant‐free and carrier‐selective

heterocontacts for silicon solar cells: recent advances and perspectives. Adv. Sci. 2018, 5 (3), 1700547.

214. Bella, F.; Gerbaldi, C.; Barolo, C.; Grätzel, M., Aqueous dye-sensitized solar cells. Chem. Soc. Rev.

2015, 44 (11), 3431-3473.

215. Peter, L. M., The gratzel cell: where next? J. Phys. Chem. Lett. 2011, 2 (15), 1861-1867.

216. Yang, D.; Yang, R.; Priya, S.; Liu, S., Recent advances in flexible perovskite solar cells: fabrication and

applications. Angew. Chem. Int. Ed. 2019, 58 (14), 4466-4483.

217. Li, Z.; Klein, T. R.; Kim, D. H.; Yang, M.; Berry, J. J.; van Hest, M. F.; Zhu, K., Scalable fabrication of

perovskite solar cells. Nat. Rev. Mater. 2018, 3 (4), 1-20.

218. Cheng, P.; Zhan, X., Stability of organic solar cells: challenges and strategies. Chem. Soc. Rev. 2016, 45

(9), 2544-2582.

219. Stoltzfus, D. M.; Donaghey, J. E.; Armin, A.; Shaw, P. E.; Burn, P. L.; Meredith, P., Charge generation

pathways in organic solar cells: assessing the contribution from the electron acceptor. Chem. Rev. 2016,

116 (21), 12920-12955.

220. Wang, G.; Melkonyan, F. S.; Facchetti, A.; Marks, T. J., All‐Polymer Solar Cells: Recent Progress,

Challenges, and Prospects. Angew. Chem. Int. Ed. 2019, 58 (13), 4129-4142.

221. Xu, W.; Gao, F., The progress and prospects of non-fullerene acceptors in ternary blend organic solar

cells. Mater. Horiz. 2018, 5 (2), 206-221.

222. Kim, K.-H.; Yu, H.; Kang, H.; Kang, D. J.; Cho, C.-H.; Cho, H.-H.; Oh, J. H.; Kim, B. J., Influence of

intermolecular interactions of electron donating small molecules on their molecular packing and

performance in organic electronic devices. J. Mater. Chem. A. 2013, 1 (46), 14538-14547.

223. Friederich, P.; Gómez, V.; Sprau, C.; Meded, V.; Strunk, T.; Jenne, M.; Magri, A.; Symalla, F.;

Colsmann, A.; Ruben, M., Rational in silico design of an organic semiconductor with improved electron

mobility. Adv. Mater. 2017, 29 (43), 1703505.

224. Yi, A.; Chae, S.; Hong, S.; Lee, H. H.; Kim, H. J., Manipulating the crystal structure of a conjugated

polymer for efficient sequentially processed organic solar cells. Nanoscale 2018, 10 (45), 21052-21061.

225. Hutter, E. M.; Gélvez-Rueda, M. C.; Osherov, A.; Bulović, V.; Grozema, F. C.; Stranks, S. D.; Savenije,

T. J., Direct–indirect character of the bandgap in methylammonium lead iodide perovskite. Nat. Mater.

2017, 16 (1), 115-120.

226. Kirchartz, T.; Rau, U., Decreasing radiative recombination coefficients via an indirect band gap in lead

halide perovskites. J. Phys. Chem. Lett. 2017, 8 (6), 1265-1271.

227. Wang, T.; Daiber, B.; Frost, J. M.; Mann, S. A.; Garnett, E. C.; Walsh, A.; Ehrler, B., Indirect to direct

bandgap transition in methylammonium lead halide perovskite. Energy. Environ. Sci. 2017, 10 (2), 509-

515.

228. Kitagawa, S.; Kitaura, R.; Noro, S. i., Functional porous coordination polymers. Angew. Chem. Int. Ed.

2004, 43 (18), 2334-2375.

229. Furukawa, H.; Cordova, K. E.; O’Keeffe, M.; Yaghi, O. M., The chemistry and applications of metal-

organic frameworks. Science 2013, 341 (6149).

230. Keller, N.; Calik, M.; Sharapa, D.; Soni, H. R.; Zehetmaier, P. M.; Rager, S.; Auras, F.; Jakowetz, A. C.;

Gorling, A.; Clark, T., Enforcing extended porphyrin J-aggregate stacking in covalent organic

frameworks. J. Am. Chem. Soc. 2018, 140 (48), 16544-16552.

Page 162: Theoretical Investigations of the Photophysical Properties

156

231. Aziz, A.; Ruiz-Salvador, A. R.; Hernández, N. C.; Calero, S.; Hamad, S.; Grau-Crespo, R., Porphyrin-

based metal-organic frameworks for solar fuel synthesis photocatalysis: band gap tuning via iron

substitutions. J. Mater. Chem. A. 2017, 5 (23), 11894-11904.

232. Yuan, S.; Liu, T.-F.; Feng, D.; Tian, J.; Wang, K.; Qin, J.; Zhang, Q.; Chen, Y.-P.; Bosch, M.; Zou, L.,

A single crystalline porphyrinic titanium metal–organic framework. Chem. Sci. 2015, 6 (7), 3926-3930.

233. Peters, M. V.; Goddard, R.; Hecht, S., Synthesis and characterization of azobenzene-confined

porphyrins. J. Org. Chem. 2006, 71 (20), 7846-7849.

234. A structure search using the porphyrin-core in SciFinder results ~ 116, p. s.

235. Aratani, N.; Kim, D.; Osuka, A., Discrete cyclic porphyrin arrays as artificial light-harvesting antenna.

Acc. Chem. Res. 2009, 42 (12), 1922-1934.

236. Urbani, M.; Grätzel, M.; Nazeeruddin, M. K.; Torres, T., Meso-substituted porphyrins for dye-sensitized

solar cells. Chem. Rev. 2014, 114 (24), 12330-12396.

237. Mahmood, A.; Hu, J.-Y.; Xiao, B.; Tang, A.; Wang, X.; Zhou, E., Recent progress in porphyrin-based

materials for organic solar cells. J. Mater. Chem. A. 2018, 6 (35), 16769-16797.

238. Wu, S.-L.; Lu, H.-P.; Yu, H.-T.; Chuang, S.-H.; Chiu, C.-L.; Lee, C.-W.; Diau, E. W.-G.; Yeh, C.-Y.,

Design and characterization of porphyrin sensitizers with a push-pull framework for highly efficient dye-

sensitized solar cells. Energy. Environ. Sci. 2010, 3 (7), 949-955.

239. Senge, M. O.; Shaker, Y. M.; Pintea, M.; Ryppa, C.; Hatscher, S. S.; Ryan, A.; Sergeeva, Y., Synthesis

of meso-substituted ABCD-type porphyrins via functionalization reactions. Eur. J. Org. Chem 2010,

237-258.

240. Heitmann, G.; Dommaschk, M.; Low, R.; Herges, R., Modular Synthetic Route to Monofunctionalized

Porphyrin Architectures. Org. Lett. 2016, 18 (20), 5228-5231.

241. Bengasi, G.; Baba, K.; Frache, G.; Desport, J.; Gratia, P.; Heinze, K.; Boscher, N. D., Conductive fused

porphyrin tapes on sensitive substrates by a chemical vapor deposition approach. Angew. Chem. Int. Ed.

2019, 131 (7), 2125-2130.

242. Boscher, N. D.; Wang, M.; Gleason, K. K., Chemical vapour deposition of metalloporphyrins: a simple

route towards the preparation of gas separation membranes. J. Mater. Chem. A 2016, 4 (46), 18144-

18152.

243. Medforth, C. J.; Wang, Z.; Martin, K. E.; Song, Y.; Jacobsen, J. L.; Shelnutt, J. A., Self-assembled

porphyrin nanostructures. Chem. Commun. 2009, (47), 7261-7277.

244. Grimme, S.; Ehrlich, S.; Goerigk, L., Effect of the damping function in dispersion corrected density

functional theory. J. Comput. Chem. 2011, 32 (7), 1456-1465.

245. Weigend, F., Accurate Coulomb-fitting basis sets for H to Rn. Phys. Chem. Chem. Phys. 2006, 8 (9),

1057-1065.

246. Coupry, D. E.; Addicoat, M. A.; Heine, T., Extension of the universal force field for metal–organic

frameworks. J. Chem. Theory Comput. 2016, 12 (10), 5215-5225.

247. Peintinger, M. F.; Oliveira, D. V.; Bredow, T., Consistent Gaussian basis sets of triple‐zeta valence with

polarization quality for solid‐state calculations. J. Comput. Chem. 2013, 34 (6), 451-459.

248. Fukuda, R.; Ehara, M., Mechanisms for solvatochromic shifts of free-base porphine studied with

polarizable continuum models and explicit solute–solvent interactions. J. Chem. Theory. Comput. 2013,

9 (1), 470-480.

249. Ghosh, M.; Sinha, S., Solvatochromic Stokes shift and determination of excited state dipole moments of

free base and zinc octaethylporphyrin. Spectrochim. Acta. Part A: Mol. Biomol. Spectrosc. 2015, 150,

959-965.

250. Gliemann, H.; Wöll, C., Epitaxially grown metal-organic frameworks. Mater. Today. 2012, 15 (3), 110-

116.

251. Shekhah, O.; Liu, J.; Fischer, R.; Wöll, C., MOF thin films: existing and future applications. Chem. Soc.

Rev. 2011, 40 (2), 1081-1106.

Page 163: Theoretical Investigations of the Photophysical Properties

157

252. Falcaro, P.; Okada, K.; Hara, T.; Ikigaki, K.; Tokudome, Y.; Thornton, A. W.; Hill, A. J.; Williams, T.;

Doonan, C.; Takahashi, M., Centimetre-scale micropore alignment in oriented polycrystalline metal–

organic framework films via heteroepitaxial growth. Nat. Mater. 2017, 16 (3), 342-348.

253. Wannapaiboon, S.; Sumida, K.; Dilchert, K.; Tu, M.; Kitagawa, S.; Furukawa, S.; Fischer, R. A.,

Enhanced properties of metal–organic framework thin films fabricated via a coordination modulation-

controlled layer-by-layer process. J. Mater. Chem. A. 2017, 5 (26), 13665-13673.

254. Haldar, R.; Jakoby, M.; Mazel, A.; Zhang, Q.; Welle, A.; Mohamed, T.; Krolla, P.; Wenzel, W.; Diring,

S.; Odobel, F., Anisotropic energy transfer in crystalline chromophore assemblies. Nat. Commun. 2018,

9 (1), 1-8.

255. Rowsell, J. L.; Yaghi, O. M., Metal–organic frameworks: a new class of porous materials. Microporous

Mesoporous Mater. 2004, 73 (1-2), 3-14.

256. Long, J. R.; Yaghi, O. M., The pervasive chemistry of metal–organic frameworks. Chem. Soc. Rev. 2009,

38 (5), 1213-1214.

257. Batten, S. R.; Neville, S. M.; Turner, D. R., Coordination polymers: design, analysis and application.

Royal Society of Chemistry: 2008.

258. Cote, A. P.; Benin, A. I.; Ockwig, N. W.; O'Keeffe, M.; Matzger, A. J.; Yaghi, O. M., Porous, crystalline,

covalent organic frameworks. Science 2005, 310 (5751), 1166-1170.

259. Diercks, C. S.; Yaghi, O. M., The atom, the molecule, and the covalent organic framework. Science 2017,

355 (6328).

260. Deng, H.; Grunder, S.; Cordova, K. E.; Valente, C.; Furukawa, H.; Hmadeh, M.; Gándara, F.; Whalley,

A. C.; Liu, Z.; Asahina, S., Large-pore apertures in a series of metal-organic frameworks. Science 2012,

336 (6084), 1018-1023.

261. Hönicke, I. M.; Senkovska, I.; Bon, V.; Baburin, I. A.; Bönisch, N.; Raschke, S.; Evans, J. D.; Kaskel,

S., Balancing mechanical stability and ultrahigh porosity in crystalline framework materials. Angew.

Chem. Int. Ed. 2018, 57 (42), 13780-13783.

262. Furukawa, H.; Ko, N.; Go, Y. B.; Aratani, N.; Choi, S. B.; Choi, E.; Yazaydin, A. Ö.; Snurr, R. Q.;

O’Keeffe, M.; Kim, J., Ultrahigh porosity in metal-organic frameworks. Science 2010, 329 (5990), 424-

428.

263. Suh, M. P.; Park, H. J.; Prasad, T. K.; Lim, D.-W., Hydrogen storage in metal–organic frameworks.

Chem. Rev. 2012, 112 (2), 782-835.

264. Wahiduzzaman, M.; Walther, C. F.; Heine, T., Hydrogen adsorption in metal-organic frameworks: The

role of nuclear quantum effects. J. Chem. Phys. 2014, 141 (6), 064708.

265. Hu, Z.; Wang, Y.; Shah, B. B.; Zhao, D., CO2 capture in metal–organic framework adsorbents: an

engineering perspective. Adv. Sustainable Syst. 2019, 3 (1), 1800080.

266. Ding, M.; Flaig, R. W.; Jiang, H.-L.; Yaghi, O. M., Carbon capture and conversion using metal–organic

frameworks and MOF-based materials. Chem. Soc. Rev. 2019, 48 (10), 2783-2828.

267. Kim, J. Y.; Oh, H.; Moon, H. R., Isotope Separation: Hydrogen Isotope Separation in Confined

Nanospaces: Carbons, Zeolites, Metal–Organic Frameworks, and Covalent Organic Frameworks (Adv.

Mater. 20/2019). Adv. Mater. 2019, 31 (20), 1970147.

268. Weinrauch, I.; Savchenko, I.; Denysenko, D.; Souliou, S.; Kim, H.; Le Tacon, M.; Daemen, L. L.; Cheng,

Y.; Mavrandonakis, A.; Ramirez-Cuesta, A., Capture of heavy hydrogen isotopes in a metal-organic

framework with active Cu (I) sites. Nat. Commun. 2017, 8 (1), 1-7.

269. Wang, Z.; Knebel, A.; Grosjean, S.; Wagner, D.; Bräse, S.; Wöll, C.; Caro, J.; Heinke, L., Tunable

molecular separation by nanoporous membranes. Nat. Commun. 2016, 7 (1), 1-7.

270. Knebel, A.; Geppert, B.; Volgmann, K.; Kolokolov, D.; Stepanov, A.; Twiefel, J.; Heitjans, P.; Volkmer,

D.; Caro, J., Defibrillation of soft porous metal-organic frameworks with electric fields. Science 2017,

358 (6361), 347-351.

271. Batten, S. R.; Champness, N. R.; Chen, X.-M.; Garcia-Martinez, J.; Kitagawa, S.; Öhrström, L.;

O’Keeffe, M.; Suh, M. P.; Reedijk, J., Terminology of metal–organic frameworks and coordination

polymers (IUPAC Recommendations 2013). Pure Appl. Chem. 2013, 85 (8), 1715-1724.

Page 164: Theoretical Investigations of the Photophysical Properties

158

272. Hoffmann, R., Solids and Surfaces: A Chemist's View of Bonding in Extended Structures. (Print on

Demand Edition 1988), Wiley‐VCH, New York 2002. 1988.

273. Muller, K.; Fink, K.; Schottner, L.; Koenig, M.; Heinke, L.; Woll, C., Defects as Color Centers: The

Apparent Color of Metal–Organic Frameworks Containing Cu2+-Based Paddle-Wheel Units. ACS Appl.

Mater. Interfaces 2017, 9 (42), 37463-37467.

274. Jin, E.; Asada, M.; Xu, Q.; Dalapati, S.; Addicoat, M. A.; Brady, M. A.; Xu, H.; Nakamura, T.; Heine,

T.; Chen, Q., Two-dimensional sp2 carbon–conjugated covalent organic frameworks. Science 2017, 357

(6352), 673-676.

275. Dong, R.; Han, P.; Arora, H.; Ballabio, M.; Karakus, M.; Zhang, Z.; Shekhar, C.; Adler, P.; Petkov, P.

S.; Erbe, A., High-mobility band-like charge transport in a semiconducting two-dimensional metal–

organic framework. Nat. Mater. 2018, 17 (11), 1027-1032.

276. Dong, R.; Zhang, Z.; Tranca, D. C.; Zhou, S.; Wang, M.; Adler, P.; Liao, Z.; Liu, F.; Sun, Y.; Shi, W.,

A coronene-based semiconducting two-dimensional metal-organic framework with ferromagnetic

behavior. Nat. Commun. 2018, 9 (1), 1-9.

277. Zhuang, X.; Zhao, W.; Zhang, F.; Cao, Y.; Liu, F.; Bi, S.; Feng, X., A two-dimensional conjugated

polymer framework with fully sp 2-bonded carbon skeleton. Polym. Chem. 2016, 7 (25), 4176-4181.

278. Ding, X.; Chen, L.; Honsho, Y.; Feng, X.; Saengsawang, O.; Guo, J.; Saeki, A.; Seki, S.; Irle, S.; Nagase,

S., An n-channel two-dimensional covalent organic framework. J. Am. Chem. Soc. 2011, 133 (37),

14510-14513.

279. Heinke, L.; Woll, C., Surface‐Mounted Metal–Organic Frameworks: Crystalline and Porous Molecular

Assemblies for Fundamental Insights and Advanced Applications. Adv. Mater. 2019, 31 (26), 1806324.

280. Zacher, D.; Shekhah, O.; Wöll, C.; Fischer, R. A., Thin films of metal–organic frameworks. Chem. Soc.

Rev. 2009, 38 (5), 1418-1429.

281. Li, H.; Eddaoudi, M.; Groy, T. L.; Yaghi, O., Establishing microporosity in open metal− organic

frameworks: gas sorption isotherms for Zn (BDC)(BDC= 1, 4-benzenedicarboxylate). J. Am. Chem. Soc.

1998, 120 (33), 8571-8572.

282. Mueller, U.; Schubert, M.; Teich, F.; Puetter, H.; Schierle-Arndt, K.; Pastre, J., Metal–organic

frameworks—prospective industrial applications. J. Mater. Chem. 2006, 16 (7), 626-636.

283. Carson, C. G.; Hardcastle, K.; Schwartz, J.; Liu, X.; Hoffmann, C.; Gerhardt, R. A.; Tannenbaum, R.,

Synthesis and structure characterization of copper terephthalate metal–organic frameworks. Eur. J.

Inorg. Chem. 2009, 2009 (16), 2338-2343.

284. Clausen, H. F.; Poulsen, R. D.; Bond, A. D.; Chevallier, M.-A. S.; Iversen, B. B., Solvothermal synthesis

of new metal organic framework structures in the zinc–terephthalic acid–dimethyl formamide system. J.

Solid State Chem. 2005, 178 (11), 3342-3351.

285. Lever, A. B. P.; Leznoff, C. C., Phthalocyanines: Properties and Applications. New York: VCH, 1989-

c1996.: 1996.

286. Latos-Grazynski, L.; Kadish, K.; Smith, K.; Guilard, R., The Porphyrin Handbook. In:, Academic Press:

2000.

287. Petritsch, K.; Friend, R.; Lux, A.; Rozenberg, G.; Moratti, S.; Holmes, A., Liquid crystalline

phthalocyanines in organic solar cells. Synth. Met. 1999, 102 (1-3), 1776-1777.

288. Zhou, R.; Josse, F.; Gopel, W.; Özturk, Z.; Bekaroğlu, Ö., Phthalocyanines as sensitive materials for

chemical sensors. O. Appl. Organomet. Chem. 1996, 10 (8), 557-577.

289. Guillaud, G.; Simon, J.; Germain, J., Metallophthalocyanines: gas sensors, resistors and field effect

transistors. Coord. Chem. Rev. 1998, 178, 1433-1484.

290. De La Torre, G.; Vazquez, P.; Agullo-Lopez, F.; Torres, T., Role of structural factors in the nonlinear

optical properties of phthalocyanines and related compounds. Chem. Rev. 2004, 104 (9), 3723-3750.

291. Claessens, C. G.; Hahn, U.; Torres, T., Phthalocyanines: From outstanding electronic properties to

emerging applications. Chem. Rec. 2008, 8 (2), 75-97.

292. Hassan, A.; Gould, R., Structural studies of thermally evaporated thin films of copper phthalocyanine.

Phys. Stat. Sol. A 1992, 132 (1), 91-101.

Page 165: Theoretical Investigations of the Photophysical Properties

159

293. Nešpůrek, S.; Podlesak, H.; Hamann, C., Structure and photoelectrical behaviour of vacuum-evaporated

metal-free phthalocyanine films. Thin Solid Films 1994, 249 (2), 230-235.

294. Rajesh, K.; Menon, C., Estimation of the refractive index and dielectric constants of magnesium

phthalocyanine thin films from its optical studies. Mater. Lett. 2002, 53 (4-5), 329-332.

295. Xia, W.; Minch, B. A.; Carducci, M. D.; Armstrong, N. R., LB films of rodlike phthalocyanine

aggregates: Specular X-ray reflectivity studies of the effect of interface modification on coherence and

microstructure. Langmuir 2004, 20 (19), 7998-8005.

296. Oshiro, T.; Backstrom, A.; Cumberlidge, A.-M.; Hipps, K.; Mazur, U.; Pevovar, S.; Bahr, D.; Smieja, J.,

Nanomechanical properties of ordered phthalocyanine Langmuir–Blodgett layers. J. Mater. Res. 2004,

19 (5), 1461-1470.

297. Cook, M. J., Phthalocyanine thin films. Pure Appl. Chem. 1999, 71 (11), 2145-2151.

298. Calik, M.; Auras, F.; Salonen, L. M.; Bader, K.; Grill, I.; Handloser, M.; Medina, D. D.; Dogru, M.;

Lobermann, F.; Trauner, D.; Hartschuh, A.; Bein, T., Extraction of photogenerated electrons and holes

from a covalent organic framework integrated heterojunction. J Am Chem Soc 2014, 136 (51), 17802-7.

299. Zimmerman, S. S.; Khatri, A.; Garnier-Amblard, E. C.; Mullasseril, P.; Kurtkaya, N. L.; Gyoneva, S.;

Hansen, K. B.; Traynelis, S. F.; Liotta, D. C., Design, synthesis, and structure-activity relationship of a

novel series of GluN2C-selective potentiators. J Med Chem 2014, 57 (6), 2334-56.

300. Satake, A.; Shoji, O.; Kobuke, Y., Supramolecular array of imizazolylethynyl-zinc-porphyrin. J.

Organomet. Chem. 2007, 692 (1), 635-644.

301. Senge, M. O.; Fazekas, M.; Pintea, M.; Zawadzka, M.; Blau, W. J., 5,15‐A2B2‐ and 5,15‐A2BC‐Type

Porphyrins with Donor and Acceptor Groups for Use in Nonlinear Optics and Photodynamic Therapy.

Eur. J. Org. Chem. 2011, 2011 (29), 5797-5816.

302. Ogawa, K.; Hara, C.; Kobuke, Y., Syntheses and nonlinear absorption properties of conjugated porphyrin

supramolecules. J. Porphyr. Phthalocyanines 2007, 11 (05), 359-367.

303. Götz, D. C. G.; Bruhn, T.; Senge, M. O.; Bringmann, G., Synthesis and Stereochemistry of Highly

Unsymmetric β,Meso-Linked Porphyrin Arrays. J. Org. Chem. 2009, 74 (21), 8005-8020.

304. Zhai, B.; Shuai, L.; Yang, L.; Weng, X.; Wu, L.; Wang, S.; Tian, T.; Wu, X.; Zhou, X.; Zheng, C., Octa-

Substituted Anionic Porphyrins: Topoisomerase I Inhibition and Tumor Cell Apoptosis Induction.

Bioconjug. Chem. 2008, 19 (8), 1535-1542.

305. Zhou, W.; Begum, S.; Wang, Z.; Krolla, P.; Wagner, D.; Bräse, S.; Wöll, C.; Tsotsalas, M., High

Antimicrobial Activity of Metal–Organic Framework-Templated Porphyrin Polymer Thin Films. ACS

Appl. Mater. Interfaces 2018, 10 (2), 1528-1533.

Page 166: Theoretical Investigations of the Photophysical Properties

160

Page 167: Theoretical Investigations of the Photophysical Properties

161

Acknowledgements

Having concluded three and half years of research work, it comes with no surprise that there were many

helping hands and minds that steered me along this journey and a few words of appreciation are too less

but absolutely justified. I try my best to acknowledge people chronologically.

First and foremost, I would like to thank my parents Shri Yad Ram Badgujar and Shrimati Bimla Devi

back in India for every struggle they endured so that my path towards and along this journey remains

as smooth as possible. Also, I give my heartfelt thanks to my three beloved sisters (Sangeeta, Manjeeta,

and Kavita) for their endless love and support throughout my entire life.

No words suffice for my supervisor (Doktorvater in German) Prof. Dr. Thomas Heine. I am highly

grateful to you for offering me this lifetime opportunity. Before coming to your lab, I trained primarily

within the experimental periphery of chemistry and material science. Besides, a little basic textbook

knowledge, I had no hands-on experience in theoretical aspects of chemistry. It was, therefore, all very

overwhelming when I undertook my thesis in theoretical chemistry, this not only extended my research

periphery but also elevated my computational skill set. During these years, I have learned a lot from our

meetings, personal talks and by just observing how you conduct yourself in social and professional life.

Despite your busy schedule you always find time to resolve my silly questions and doubts. Also, special

thanks for trusting in and giving me a lot of freedom to collaborate with other groups which helped me

to navigate effectively through this journey and reach the finish line of my PhD. It was amazing to work

as I learnt the qualities of being humble and at the same time being a successful researcher.

Next, I am indebted to Dr. Stefan Zahn for introducing me to the density functional calculations and his

willingness to mentor a novice like me. I really admire his hard work and expertise. Also, I would like

to especially thank Dr. Agnieszka Kuc for giving me a hands-on experience in electronic band structure

calculations. For all the technical support and solving the computational bottlenecks, I am thankful to

Dr. Lyuben Zhechkov and Dipl. Knut Vietze. Further, Dr. Jan-Ole Joswig for having useful discussions

and introducing me to the interesting teaching assistant duties. Grateful acknowledgment is also made

Page 168: Theoretical Investigations of the Photophysical Properties

162

to Dr. Nina Vankova, who gave me continuous guidance in my projects and considerable help in my

thesis corrections by means of suggestions, comments, and fruitful discussions. I would like to express

my sincere gratitude towards Ms. Antje Völkel for all the help related to my administrative works and

organizing the group retreats and events. Hung-Hsuan for being my longest-running office mate and

sharing your ideas and thoughts in understanding chemistry. Hope you succeed in automating the whole

task of PhD itself. A big thanks also to all the present and past members of the theoretical chemistry

group in Leipzig and Dresden for their valuable discussions and brainstorms. Thank you all, for the nice

cakes, beers, barbecues, parties, and all the fun-adventurous biking-hiking trips.

Finally, the main part of the work on this thesis was performed while being employed in a position

funded by the DFG within the COORNETs project (SPP 1928). So, I want to thank the DFG and all its

representatives, as well as our principal investigators: Prof. Dr. Thomas Heine, Prof. Dr. Christof Wöll,

Prof. Dr. Stefan Bräse and Prof. Dr. Roland Fischer for providing and securing these funds and therefore

making this thesis possible. Further, ZIH Dresden is thanked for providing high-performance computing

resources and a big thank also to all the collaborators that participated in this wonderful project.

Page 169: Theoretical Investigations of the Photophysical Properties

163

Versicherung

Hiermit versichere ich, dass ich die vorliegende Arbeit ohne unzulässige Hilfe Dritter und ohne

Benutzung anderer als der angegebenen Hilfsmittel angefertigt habe; die aus fremden Quellen

direkt oder indirekt übernommen Gedanken sind als solche kenntlich gemacht. Die Arbeit

wurde bisher weder im Inland noch im Ausland in gleicher oder ähnlicher Form einer anderen

Prüfungsbehörde vorgelegt.

Datum Unterschrift

Erklärung

Die vorliegende Arbeit wurde in der Zeit von Mai 2017 bis Dezember 2020 an der Technischen

Universität Dresden im Rahmen des Projektes zum Thema: “ Theoretical Investigations of the

Photophysical Properties of Chromophoric Metal-Organic Frameworks” unter

wissenschaftlicher Betreuung von Herrn Prof. Dr. Thomas Heine durchgeführt.

Datum Unterschrift