You are on page 1of 196

University of Insubria

Department of Science and High Technology


PhD course in Chemical and Environmental Sciences (Chemical Science Curriculum)

BLUE EMISSIVE MATERIALS FOR


OPTOELECTRONIC DEVICES

Doctoral dissertation of:


Gioele Colombo

Supervisor: Prof. Stefano Brenna


Co-Supervisor: Prof. G.A. Ardizzoia

Academic Year 2020/2021


Index

Abbreviation 6
Chapter 1: Introduction 8
1.1 Photoluminescence Processes 9
1.1.1 Fluorescence and Phosphorescence 10
1.1.2 Luminescence Lifetime and Quantum Yield 12
1.1.3 Solvent Effect 14
1.1.4 Factors Affecting Fluorescence Intensity 15
1.1.5 Luminescence Quenching 17
1.2 Luminescence in First-Row Transition Metals
Coordination Compounds 18
1.2.1 Electronic Transitions in Metal Complexes 19
1.2.2 Fluorescence in Compounds of First-Row Transition Metal 23
1.3 Fluorescence in Action: Optoelectronic Devices 25
1.4 Imidazo[1,5-a]pyridines 28
1.4.1 Synthesis 28
1.4.2 Medical, Biological and Other Applications: a Short Overview 30
1.4.3 Imidazo[1,5-a]pyridines as Ligands 32
1.4.4 Luminescence in Imidazo[1,5-a]pyridine Derivatives 35
1.5 Fluorescence in BF2-Compounds 37
1.6 Chapters Outline 39

Chapter 2: Previous Results on Imidazo[1,5-a]


pyridines Among Our Research Group 41
2.1 Compounds with Neutral N,N-Bidentate
1-pyridylimidazo[1,5-a]pyridines as Ligands 42
2.1.1 Zinc(II) Halido Complexes 42
2.1.2 Zinc(II) Complexes with Polyatomic Anions 47
2.1.3 Heteroleptic Silver(I) Complexes 49
2.2 Compounds with Anionic N,O-Bidentate
2-(imidazo[1,5-a]pyridine-3-yl) Phenols 53
2.2.1 Catalytic Activity of Homoleptic Complexes of Divalent Metals 54
2.2.2 Tetrahydro-2-(imidazo[1,5-a]pyridine-3-yl) Phenols 54

Chapter 3: N,O-Homoleptic Zinc(II) Complexes


([Zn(LR)2]) 57
3.1 Synthesis and Characterization 58
3.2 Optical Properties in Solution 62
3.3 DFT Calculations 65

Chapter 4: Boron Difluoride Complexes of


Imidazo[1,5-a]pyridinyl Phenols (LR_BF2) 70
4.1 Synthesis and Characterization 71
4.2 Optical Properties in Solution 77
4.3 DFT Calculations 87

Chapter 5: Boron Difluoride Compounds


of Tetrahydroimidazo[1,5-a]pyridinyl
Phenols (hydrLR_BF2) 92
5.1 Synthesis and Characterization 93
5.2 Optical Properties 97
5.2.1 Optical Properties in Solution 97
5.2.2 Optical Properties in Solid State 101
5.2.3 Optical Properties in Polymeric Film 102
5.3 DFT Calculations 103
Chapter 6: Building OLEDs 106
6.1 Preliminary Screening 107
6.2 Basic Architecture of Devices 108
6.3 Energetic Issue 116
6.4 Film Morphology Issue 119
6.5 Possible Future Outlooks 123

Chapter 7: Conclusions 125

Chapter 8: Experimental Section 128


8.1 General Remarks 129
8.2 Synthesis of HLR Ligands 130
8.3 Synthesis of [Zn(LR)2] Compounds 131
8.4 Synthesis of LR_BF2 Compounds 136
8.5 Synthesis of hydroHLR Ligands 154
8.6 Synthesis of hydroLR_BF2 Compounds 154
8.7 Crystallographic Data 166
8.8 Computational Details 176

Appendix: DFT Calculations 177

References 190

5
Abbreviations

BODIHY Borondifluorohydrazone
BODIPY Borondipyrromethene
BOPHY Boronpyrrolylhydrazine
COSMO Conductor-Like Continuum Solvent Model
DSC-TGA Differential Scanning Calorimetry-Termogravimetric Analysis
DSSCs Dye-Sensitized Solar Cells
EDDM Electron Density Difference Map
GGA Generalized Gradient Approximation
GTO Gaussian Type Orbital
HF Hartree-Fock
HTL Hole Transporting Layer
HTM Hole Transporting Material
IC Internal Conversion
IR Infrared
ISC Intersystem Crossing
ITO Indium Tin Oxide
LC Ligand Centered (transition)
LCAO Ligand Combination of Atomic Orbitals
LDA Local Density Approximation
LED Light Emitting Diode
LGOs Ligand Group Orbitals
LLCT Ligand to Ligand Charge Transfer
LMCT Ligand to Metal Charge Transfer
LTT Long Time Term
MC Metal Centered (transition)
MLCT Metal to Ligand Charge Transfer
MO Molecular Orbital

6
NHC Nitrogen Heterocyclic Carbene
NTOs Natural Transition Orbitals
OLEDs Organic Light Emitting Diodes
PES Potential Energy Surface
PMMA Poly(Methyl MethAcrilate)
PVK Poly(VinylCarbazole)
SCF Self-Consistent Field
SIC Self-Interaction Correction
STO Slater Type Orbital
STT Short Time Term
TBAPF6 Tetrabutylammonium hexafluorophosphate
TD-DFT Time Dependent Density Functional Theory
TON Turnover Number
UV Ultraviolet
WOLEDs White Organic Light Emitting Diodes

7
Chapter 1

INTRODUCTION

8
In 2015, the United Nations published the “2030 agenda for sustainable development”,1 a
document reporting 17 social goals to be achieved by 2030. One of them concerns the
energy sustainability, since it is well known that the energy production (and therefore the
energy use) is the main cause of climate change, accounting for more than 70% of global
greenhouse gases emission. Continued consumption of fossil fuels is leading, as well as a
rapid depletion of the resources, to a higher CO2 level in the atmosphere, with all the
consequences we can hear daily. Therefore, science should try to afford solutions to this
problem, focusing on finding new devices that can increase the energy efficiency and the
“clean energy” production, such as solar cells, molecular sensors and alternative lighting
devices like Organic Light Emitting Diodes (OLEDs).

Figure 1.1: The 17 goals set by UN by 2030 (from 2030 agenda for sustainable development).1

In this context, this project had the goal to find and study new classes of molecules whose
photophysical properties are suitable to be exploited in this type of technologies.

1.1 PHOTOLUMINESCENCE PROCESSES2

Electrons in a molecule can have different energies, which determine the electronic state
of the whole molecule. When an electromagnetic radiation hits the molecule, if the energy
of the photon is equal to the energy difference between two different electronic states,
the radiation is absorbed and the molecule is excited to a higher energy state from which
it can relax by emitting another photon (radiative process) or in many other ways (non-

9
radiative processes). These processes can be very helpful both in practical applications and
for gaining information on the observed chemical species.
The more widespread spectroscopies do so, exploiting electromagnetic radiation with
wavelengths between 100 and 1000 nm, ranging from near UV to near IR, with
corresponding energies between 12.4 to 1.2 eV.

1.1.1 FLUORESCENCE AND PHOSPHORESCENCE

As briefly described above, when electrons go from a higher energy state to a lower energy
state, the “excess” energy can be re-emitted as electromagnetic radiation: this process is
known as fluorescence. Not all the transitions are permitted though, as the general
selection rule implies that the only transitions allowed are those between two electronic
states with the same spin multiplicity.
Generally, for organic molecules by absorbing of a photon (with an appropriate energy) a
transition from a singlet ground state to a singlet excited state occurs, in time spawn of just
1 femtosecond, while the re-emission of a photon (fluorescence) is a much slower process,
ranging from nanoseconds to microseconds. Right after the excitation though, before any
photoreaction or emission of radiation can occur, the molecule undergoes a series of
incredibly fast events. In the first place, a vibrational relaxation of the excited states occurs
(internal conversion, IC). During this process, another one can occur: the intersystem
crossing (ISC) is a process in which a singlet excited state can relax to a triplet excited state.
This is possible only if there are vibrionic isoenergetic states between singlet and triplet.
From this, the radiative relaxation mechanism is called phosphorescence.
The formally forbidden transition from a different multiplicity state is in this case allowed
by an interaction between the spin magnetic moment and the magnetic moment due to
the orbital motion of the electrons, especially in presence of heavy atoms such as S, Br or I
or transition metals with unpaired electrons in d orbitals.

10
Figure 1.2: Sequence of the events leading to the luminescence process.

Since this transition is formally forbidden, phosphorescence relaxation is way slower than
fluorescence relaxation and the triplet lifetime can be way longer, ranging from 10-3 to 103
seconds.

As fluorescence arises from relaxation from the lowest vibronic excited state to one of the
vibrational levels of the ground state (Kasha’s rule), the fluorescence emission spectra are
usually red-shifted with respect to the absorption spectra, maintaining similar profile. This
is because the energy of a photon is related to its vibrational frequency:
𝐸 = ℎ𝜈 eq. 1
Also, from

𝜈= eq. 2

it can be obtained

𝐸= eq.3

As h is the Planck’s constant and c is the velocity of light, the energy varies inversely as λ.
As mentioned above, the energy of the photon re-emitted in the fluorescence process must
be less than the energy of the photon absorbed and therefore from equation 3, the
wavelength of fluorescence must be longer than the one of the absorbed light (Figure 1.3).
The same applies for phosphorescence spectra, but with bigger shifts as the triplet excited
11
state is usually less energetic than the singlet excited state, since parallel arrangement of
spins in an electronic configuration lies lower in energy than the antiparallel arrangement.

Figure 1.3: Absorption and fluorescence emission spectra of anthracene.3

1.1.2 LUMINESCENCE LIFETIME AND QUANTUM YIELD

As described above, the singlet excited state (S1) can undergo three different relaxation
mechanisms (Figure 1.4): radiative (fluorescence) to the ground state (S0), ISC to triplet
excited state (T1) and non-radiative decay to the lowest-energy excited state (internal
conversion, IC). Each one of these processes is described by a first order kinetic equation
with its own kinetic constant.

Figure 1.4: Simplified Jablonski diagram depicting the various relaxation processes.

Therefore, we can define the rate of disappearing of S1, Jtot, as:


[ ]
𝐽 =− = (𝑘 +𝑘 +𝑘 )[𝑆 ] = 𝑘 [𝑆 ] eq. 4

12
The solution to this differential equation is

[𝑆 ] = [𝑆 ] 𝑒 eq. 5

This represents an exponential decay where [S1]0 is concentration of S1 at time 0, when the
first light pulse has reached the molecule. τ is the reciprocal of ktot and it is called lifetime
of S1.
When t = τ, the concentration will be
[ ]
[𝑆 ] = [𝑆 ] 𝑒 = [𝑆 ] 𝑒 = eq. 6

It can be concluded that τ is also the time needed for S1 concentration to become 1/e
(36.8%) of its starting value, not to be confused with the half-life time.

Experimentally, lifetimes are determined by sending a very short pulse of light and then
measuring the emission decay.

Figure 1.5: Exponential decay of emission intensity.4

Obviously, τ describes every decay process for S1, not only the radiative one, which is
described instead by τ0, which is equal to the reciprocal of kF; hence, as kF < ktot, τ0 > τ.
τ0 tends to τ the less ISC and non-radiative decay contribute to the general relaxation of S1.
Nevertheless, it is also important to remember that non-radiative relaxation processes, like
molecular collisions, are always present. For this reason, the number of photons re-emitted
is always lower than the number of those absorbed. The ratio between re-emitted and

13
absorbed photons is called fluorescence quantum yield (ΦF). It can also be defined as the
fraction of molecules in a sample relaxing with a radiative decay with respect to the total
number of excited molecules. This is equal to the quotient between the emission ad
absorption rates.

𝜙 = eq. 8

In a steady state condition, with continuous illumination of the sample, the rate of
formation of the excited state is the same of the rate of relaxation with every possible
decay path:
[ ]
𝜙 = = = = = eq. 9
[ ]

And from the definitions of τ and τ0:

𝜙 = = eq. 10

τ0 can be empirically estimated by the molar extinction coefficient εMAX of the absorption
spectrum.

𝜏 = eq. 11

with εMAX expressed in l ּ mol-1 ּ cm-1 and τ0 in seconds. A more accurate esteem can be
obtained from
. ⋅
𝜏 = eq. 12
⋅ ∙∆ /

with ν, maximum of absorption expressed in cm-1 and ∆𝜈 / the half-band width.

1.1.3 SOLVENT EFFECT

When a molecule is in vacuum has a characteristic energy given by its molecular and
electronic structure. For brevity, this energy will be called E(G) (Energy of the ground state).
On the other side, if the molecule is dispersed in a solvent, it is influenced by the
electrostatic fields of the solvent molecules, which are rapidly moving around it; it is like

14
the species is placed in a fluctuating electrostatic field, hence being in a non-equilibrium
state, with energy E(G)*. As the ground state has a virtually infinite lifetime, the molecule
has the time to adapt to the solvent, as well as the solvent molecules to reorganize
themselves around it, eventually reaching a new equilibrium condition. The energy of this
new system will be called E(G)S.
We can then define the difference between the starting and the ending states as
𝛥𝐸 = 𝐸( ) − 𝐸( ) eq.13

It is to be considered though, that ΔE is composed of two terms, the first one describing
the rapid immersion of the molecule in the solvent, reaching the energy E(G)* (Short Time
Term, STT) and the second one describing the slower rearrangement of the system reaching
E(G)S (Long Time Term, LTT). More in detail, the STT is an extremely rapid process regarding
the electrostatic interaction between the solvent and the electronic cloud of the solute.
This leads to a polarization of the molecule, thus changing the orbitals’ energies, without
changing the structure. The LTT instead regards the molecular rearrangement due to the
interaction solute-solvent and it is therefore a slow process (molecularly speaking) in which
the molecule can modify its geometry in order to reach a new energy minimum.

In the absorption process, the excited state geometry should not be different from the
ground state geometry, as the electronic transition is way faster than the solvent
rearrangement according to Franck-Condon principle. On the other hand, the interaction
solute-solvent will change in the excited state, leading to a non-equilibrium situation. In
this state, the orbitals’ energies will be perturbed with respect to those one can expect in
vacuum, hence leading to a shift of the electronic transitions. It can be concluded that this
shift is due only to dipolar interaction and not to a geometry change.

1.1.4 FACTORS AFFECTING FLUORESCENCE INTENSITY

Fluorescence intensity can be affected by numerous factors, such as concentration of the


sample, solvent used, pH, temperature as well as the stability of the compounds under
irradiation of the exciting beam.

15
One can be tricked to think that fluorescence intensity and concentration are directly
proportional, but this is true only for highly dilute solutions and therefore the
concentration of the fluorophore is a very important parameter to be considered,
especially for quantitative analyses. This is due to the re-absorption process: when the
excitation beam hits the sample, the latter is excited and then it emits photons that must
pass through the whole solution before reaching the detector. The higher the
concentration of the solution, the higher is the probability that the photons can be
absorbed again by another molecule, eventually resulting in a quenching of the emission.
Also, the choice of the solvent is a crucial decision, as not only intensity but also the
emission wavelengths can vary, often in an unpredictable way. It is to be noted though,
that usually a bathochromic shift occurs by increasing the dielectric constant of the solvent.
Hydrogen bond donor or acceptor solvents can also quench or intensify the fluorescence;
as an example, chlorophyll has a very poor fluorescence in non-polar solvents, while even
a small addition of protic solvents (methanol, ethanol or water) causes a remarkable
enhancement.5
Solvents can influence the fluorescence also by protonation or deprotonation of the
fluorophore (Figure 1.6). This can also be achieved by changing the pH of the solution.
Some compounds are fluorescent only over a short range of pH and some others, even if
fluorescent on a wider range, can see their fluorescence enhance noticeably at a certain
value. For example, salicylic acid, even if fluorescent from pH 0 to 14, at pH 1 to 4 it is 100
times as fluorescent as at pH 4 to 14.6 These principles are also applied for analytical
purposes, using pH sensitive fluorescent molecules as fluorescent indicators. Also, it must
be considered that excited molecules can have different pka from their ground state and
hence be ionized by the solvent. In this case, the emission of the ionized species is observed
instead of the emission of the neutral one, with the mechanism schematized below.

Figure 1.6: Schematised excitation-ionization-emission process.

16
The intensity of fluorescence tends to decrease with the temperature raising as well. This
is because by increasing the temperature, the motion of the molecules increases
consequently, resulting in a higher collision probability, hence to an increase of the non-
radiative decay paths. On the contrary, low temperatures usually lead to a fluorescence
emission increase.
Lastly, when studying the fluorescence properties of a molecule, its photostability must be
taken into account. There are several fluorescent compounds which photo-decompose in
very dilute solution, like the well-known quinine bisulphate. A compromise has to be found
between using a very intense excitation light source to improve the sensitivity of the
instruments and not causing an excessive decomposition of the sample. However, usually
the measurements are rapid enough to be performed without the compound being
decomposed. Worthy of note is the fact that photodecomposition can also enhance the
fluorescence in some cases: coumarin for example, is weakly fluorescent, but after
decomposition leads to a highly fluorescent product.7

1.1.5 LUMINESCENCE QUENCHING

As described above, there are many factors that can influence the luminescence of a
molecule. Generally speaking, a phenomenon which reduces or abolishes the fluorescence
of a system, is called quenching process. Basically, quenching processes are all ways the
system can relax in a non-radiative fashion.
The simplest way is collisional quenching, consisting in a loss of excitation due to an energy
transfer from the excited molecule to non-excited one via collision. Typical collisional
quenchers are oxygen (Figure 1.7), halogens, or electron-deficient molecules. This type of
quenching is called “quenching by energy degradation” and the mechanism is not always
the same, but it depends on the quencher. For example, oxygen acts as a quencher by
converting the singlet excited state of the fluorophore to the triplet state, hence
diminishing the fluorescence intensity.

Figure 1.7: Schematized oxygen quenching process.

17
Quenching can also occur by energy transfer. In this case, the energy level of the excited
quencher must be just below the energy level of the excited fluorophore and usually, the
quencher has a non-luminescence triplet state, so that when the fluorescent molecule is
excited, its energy passes to the quencher without being re-emitted. This kind of quenching
is more likely to happen in the solid state, where the molecular species are closer, while it
is less likely to occur in solution, unless high concentrations are used.
The degradation of energy process can also occur by electron transfer: the excited molecule
(but not the ground state one) may accept or donate an electron to a quencher. For
example, Fe2+ ions can donate electrons to an accepting fluorophore like excited methylene
blue, without reacting with the ground state.
There are also other quenching processes, like the formation of a non-fluorescent complex
between the fluorophore and the quencher, or the ionization of the fluorophore.
Ultimately the quenching process can also occur with non-molecular mechanisms like
attenuation of the incident light by the fluorophore itself or other absorbing species. In this
case, the quenching process is known as “inner filter effect”: the quencher can absorb the
excitation light, the fluorescence emission of the fluorophore, or even both of them. Some
solvents may have this effect, like acetone or benzene which absorb in the UV region.
Another instance of the inner filter effect is the concentration of the fluorophore itself, as
described above.

1.2 LUMINESCENCE IN FIRST-ROW TRANSITION METALS COORDINATION


COMPOUNDS8
Luminescent coordination compounds have been widely studied in material chemistry.
Many precious metals have been exploited in the preparation of luminescent complexes
used in optoelectronic devices such as OLEDs,9 in order to increase the efficiency of lighting
devices, or dye-sensitized solar cells (DSSCs),10 mainly with the aim to afford a sustainable
alternative to standard energy sources like coal, oil and gas. Unfortunately, these
applications often require the use of precious metals, i.e. Ru,11 Ir,12 Pt13 or Re,14 as well as
rare-earth elements. Of course, as they have a very low abundance, their costs are
extremely high and the energy transition to sustainable sources can only be reached
partially. Also, the use of rare-earth metals is geopolitically problematic, and the extraction
can be highly water and energy demanding.
18
For these reasons, earth-abundant metals, like first row transition metals, should be
preferred for these applications. The key challenge is the fact that a given coordination
environment imposes a considerably weaker ligand field on a 3d-metal than on a 4d- or 5d-
because of the more contracted orbitals and a weaker spatial overlap with the ligand
orbitals. Consequently, 3d-metal complexes have a multitude of metal-centred transitions,
resulting in non-radiative decay paths.

1.2.1 ELECTRONIC TRANSITIONS IN METAL COMPLEXES

In order to better describe the spectroscopic properties of transition metal complexes, it


can be useful to quickly describe their electronic states. In these regards, using the
Molecular Orbital Theory, both ground and excited states can be described with a good
approximation.
Starting from a general octahedral complex [ML6], with L being σ-donor ligand (for
simplicity we do not consider π-bonding ligands) and using the symmetry properties of the
orbitals, it is possible to find which are the combinations between the metal and the ligands
orbitals. Then, linear combination of ligands orbitals generates Ligand Group Orbitals
(LGOs) which must have the same symmetry of the metal orbitals in order to overlap and
generate a bond.
From character tables, we can define the symmetries of the metal s orbital as a a1g, the p
orbitals as t1u and divide the d orbitals in two sets, eg and t2g.
As the a1g orbital has a spherical symmetry, it can overlap with LGOs on all axes, as well as
t1u and eg, which have lobes along the bond directions, while the t2g orbitals have the lobes
directed between the bonding axes and therefore cannot properly overlap with the ligands
orbitals (Figure 1.8).

19
Figure 1.8: Overlap of ligands orbitals in the xy plane with metal dx2-y2 (a) and dxy (b) orbitals.

As the t2g orbitals do not have the appropriate symmetry to overlap with the ligands
orbitals, they are considered as nonbonding (but they can still be used for π bonds).

Figure 1.9: A σ-bond molecular orbital diagram for a complex of octahedral geometry.

Even though many approximations have been made, Figure 1.9 above shows qualitatively
that a1g and t1u molecular orbitals have the lowest energies, while their corresponding
antibonding orbitals, a1g* and t1u* are those with the most antibonding character. On the
other hand, eg and eg* are less displaced, due to the poor overlap with the LGOs, while t2g
result in being nonbonding.

20
Of course, the same procedure can be used for the generation of a MO diagram of
complexes with other geometries (Figure 1.10).

Figure 1.10: σ-bond MO diagrams for a complex of tetrahedral (a) and square planar (b) geometry.

If we then consider π bonding interaction, the situation of course becomes more complex.
It must be considered that metal and ligands orbitals involved in a π bond must lie
perpendicular to the bonding axes. Using the same process described above considering
the right geometries, the following diagram can be obtained.

Figure 1.11: MO diagram for an octahedral M(CO)6 complex including both σ and π interactions.

21
With this in mind, describing the transitions which can occur in a metal complex can be
easily done:
- Metal Centered (MC): usually they are represented by d → d transi ons. It must be
noted that these transitions are formally forbidden because of the Laporte rule
(parity forbidden, t2g  eg), thus the intensity of the bands due to these transitions
is very low. Nevertheless, they are still possible because this rule is relaxed by the
vibrations of the metal-ligand bonds, which temporarily lower the symmetry.
- Ligand Centered (LC): in this case the electron transition occurs withing the orbitals
localized on the ligands, generally leading to an emission in the visible range,
especially for π → π* transitions.
- Charge Transfer (CT): when the transition occurs from two elements of the complex,
we talk of Charge Transfer process. It can be from a metal orbital to a ligand one
(MLCT), vice versa (LMCT) or even between two different ligands (LLCT).
.
In general, for luminescent phenomena occurring for transition metal complexes the
involved mechanism belongs to the charge transfer type, particularly it is quite common
that electrons are promoted from metal centered molecular orbitals to orbitals localized
on the ligand.

Figure 1.12: Simplified MO diagram showing principal electronic transitions for a transition metal complex
with octahedral geometry.

22
1.2.2 FUORESCENCE COMPOUNDS OF FIRST-ROW TRANSITION METALS15

As stated above, first-row transition metal fluorescent compounds are gaining more and
more attention due to their cheapness and availability. Nevertheless, some issues need to
be tackled in order to reach the same performances of second or third-row metals, mainly
because of the more contracted orbitals.
For instance, Ru2+, Re+, Os2+ and Ir3+ are among the most widely exploited as MLCT emitters.
3d6 metals can be a valid alternative to these, starting from Fe2+ complexes. Unfortunately,
no emissive iron(II) complex has been prepared up to now.16
Starting from Mo0 coordination chemistry, Büldt17 prepared a Cr0 complex using chelating
diisocyanide ligands, with an emission from a MLCT state in solution at room temperature,
even if the emission quantum yield resulted to be very low, while Wegeberg18 managed to
enhance it by attaching a pyrene moiety on the ligand backbone. This large improvement
is due to the delocalization effect (Figure 1.13).

Figure 1.13: Diisocyanide ligands used in the synthesis of luminescent transition metal compounds.

Lately, chelating isocyanide ligands have been used for the synthesis of the first
luminescent manganese(I) complex by Herr19, with also the advantage to be more air-stable
thanks to the higher oxidation number with respect to the just cited chromium complex.
By swapping the roles, hence using electron-rich ligands and electron-deficient metals, it is
possible to obtain fluorescence from LMCT states. Specifically, d0 configuration seem to be
the more promising in these regards, as no MC states can be formed and deactivate non-
radiatively. Group III and IV metals, respectively in +3 and +4 oxidation state have been
used, like scandium(III)20 and titanium(IV)21 metallocenes.

23
More interestingly for this Thesis, 3d10 metal complexes showed the best performances,
thanks to LLCT states. In details, thanks to this electronic configuration, there are no
energetically low-lying MC states, and therefore MLCT and LLCT states can be achieved
more easily (Figure 1.14).

Figure 1.14: Comparison between electronic transition occurring in d10 metal complexes (a) and d6 metal
complexes (b).

Copper(I) has been the workhorse of this class of compounds, starting from tetrahedral
complexes bearing diamine and/or diphosphine ligands,22 leading to MLCT emitters. It has
been recently demonstrated by Gernert23 though, that two-coordinate linear copper(I)
complexes, are way better emitters, relying on a LLCT state. Among our research group, a
phosphorescent coordination polymer of copper pyrazolsulfonate has been reported as
well (Figure 1.15).24

Figure 1.15: a) Crystal structure of Cu(I) coordination polymer viewed down the b axis. Horizontal axis, c;
vertical axis, a. b) Schematic depiction of inter- and intra-chain Cu⋯Cu separations.

24
In some cases, quantum yields approaching 100% both in solution and in solid state at room
temperature have been reported, thanks to a “π push-pull” effect obtained by using NHCs
and N-bound amines (Figure 1.16).25

Figure 1.16: NHCs (top) and amines (bottom) used to achieve the “push-pull” effect.

These compounds have a key conceptual difference from the four-coordinated complexes,
as they do not undergo Jahn-Teller distortion and hence, a non-radiative decay path. Also,
variation of the ligands leads to a tuning of the emission color over the entire visible
spectrum.
Zinc has also received a lot of consideration the preparation of luminescent complexes,26
as it will also be discussed in the following chapters.

1.3 FLUORESCENCE IN ACTION: OPTOELECTRONIC DEVICES

OLEDs are a class of optoelectronic devices whose purpose is to convert electricity into
light. They are used to create television, computer or smartphone screens, and recently
also for handheld game consoles. More interestingly, White OLEDs (WOLEDs) are studied
for possible lighting applications, being usually more energy efficient than standard
lamps.27
OLEDs differ from standard LED, which are based on p-n junction, by their structure, made
of several layers, each one with a specific purpose (Figure 1.17). Shortly, an OLEDs works

25
upon the recombination of holes and electrons in the emissive layer, thus exciting the
emissive species which then relaxes with the emission of light.28

Figure 1.17: Functioning scheme of a LED (left) and an OLED (right).

Starting from the bottom, the first layer is the support, usually made of glass, quartz or
plastic. The main function of this is of course to support the whole system and determine
the mechanical properties of the device: if a foldable support is used, a foldable device can
be built.29 Support layers must usually be transparent, in order for the light to go through.
Then comes the anode, which is usually made of Indium Tin Oxide (ITO) because of its good
electrical conductivity and optical transparency, as well as the ease with which it can be
deposited as thin film.30
The third layer is actually the first to be organic: the hole-transporting layer facilitate the
injections of electronic holes from the anode into the emissive layer. For this reason, this
layer is usually made of organic species with electron-donating moieties, to stabilize the
forming cationic radicals. Phenylamine groups are the most widely used for this purpose.31
The hearth of the devices is the central layer, which is made of fluorescent materials. They
can be of various kind, ranging from metal complexes32 (as for the first properly working
OLED published by Tang in 1987),33 to polymers, to single organic molecules, either
fluorescent or phosphorescent.34 The wavelength of the emitted photons depends on the
HOMO-LUMO energy gap of the emissive species, hence there are virtually infinite
possibilities, as long as more and more fluorescent species are found. Unfortunately, not
every emissive molecule is suitable for these systems, as they need to possess various
properties like good solubility in organic solvents or thermal stability for processability (vide
infra), as well as excellent optical stability, film forming properties and good fluorescence
quantum yields. Also, the HOMO energy should be close to that of the underlying hole-
transporting material, in order to make the hole injection process efficient. Similarly, the

26
LUMO energy should be similar to that of the electron-transporting layer, the last organic
layer.
This is usually made of molecules bearing electron-accepting moieties, so that they can
form stable anion radicals when facilitating the injection of electrons from the cathode into
the emissive layer. Azole derivatives like oxadiazoles35 and triazoles36 are often used, as
well as pyridines.37
Finally, on top of the device, a metallic cathode can be found, usually made of aluminum,
silver or even calcium and magnesium, doped with inorganic salts (i.e. LiF, CsF) to lower the
Fermi energy and make the electron injection more efficient.
Not all these layers are always needed, in the simplest scenarios, OLEDs made only of the
emissive layers sandwiched between the two electrodes can be built.38

Figure 1.18: Examples of hole transporting materials (left) and electron transporting materials (right).

To fabricate OLEDs, two main methods can be followed:


- Evaporation process: in this case the organic species (as well as the metallic
cathode) are deposited one after the other by vacuum thermal evaporation. This
has the advantage of a fine control over the thickness and the homogeneity of the
layers, as well as the possibility to build very complex architectures. Unfortunately,
this method is much more expensive than the following one, and it cannot be
applied on large-area devices, limiting its applicability outside a laboratory scale.
- Solution process: with soluble organic species it is possible to form layers by spin
coating. Of course, the species must be soluble in organic solvents and they need
to have very good film forming properties in order to compete with the film quality
obtained by vapor deposition. In some cases, dispersing the emissive material in a
polymeric matrix (PMMA, PVK) can be an option to obtain better quality films.

27
This method of deposition is more cheap, simple and scalable than vacuum
deposition, with also the possibility to use printers to build large-scale devices. Still
there are some drawbacks: it can be very difficult to spin coat an organic film
without redissolving the underlying one; for this reason, every material should be
soluble in a specific organic solvent and insoluble in others, so that it would not be
washed away when the next layer is spin coated on top of it. Another solution to
this problem, as reported by Meerholz,39 can be to convert a soluble material into
an insoluble polymer film by crosslinking during the annealing process.

As stated above, the core of these systems is the emissive layer, made of fluorescent
organic or organometallic materials. For this reason, it can be important to continue this
introduction by considering some very important and well-studied classes of emissive
molecules, which have been the subjects of this PhD project.

1.4 IMIDAZO[1,5-a]PYRIDINES

Imidazo[1,5-a]pyridines are a class of heterocyclic compounds widely known in the


literature. From 1955, the year when the first paper citing this molecule was published,40
many other studies have been carried out on it, ranging from its application in the
biomedical field, to their use in the fabrication of optoelectronic devices or as ligands in the
synthesis of many coordination compounds.

Figure 1.19: imidazo[1,5-a]pyridine structure.

1.4.1 SYNTHESIS

The first synthetic method for imidazo[1,5-a]pyridine has been the one described in the
aforementioned paper by Bower,40 using 2-formamidemethylpyridine and phosphoryl
chloride in benzene. Luckily, several new synthetic methods have been developed over

28
time and nowadays there more ways to synthesize the original imidazo[1,5-a]pyridine core.
More interestingly, among others, there are four main ways to obtain substituted
derivatives, more useful in modern chemistry:

Figure 1.20: Main synthetic pathways for the imidazo[1,5-a]pyridine core.

The second method sees a pyridylketone reacting with a α-amino acid to give the desired
product,41 while the third one, reported by Hu, exploits a sp3 C-H amination reaction
mediated by molecular iodine.42 The same reaction has also been reported to proceed
using molecular sulphur instead of iodine.43
The last one instead, the one followed for the synthesis of the imidazo[1,5-a]pyridine
derivatives which will be discussed in this Thesis, is performed with pyridylketones,
aromatic aldehydes and ammonium acetate as source of nitrogen. To our goal products,
this seemed to be the cheapest and easiest way among all the others and therefore it has
been the only way we pursued, giving satisfying yields of over 70%.

29
Figure 1.21: reaction mechanism proposed for the synthesis of imidazo[1,5-a]pyridines.

The reaction mechanism (Figure 1.21), proposed by Wang,44 starts with the imine
formation from the aldehyde and ammonium acetate. Then, a nucleophilic attack to the
carbonylic carbon of the ketone occurs, with the following ring closure and aromatization
of the molecule. Moreover, it is also possible to have various substituents on the pyridinyl
ring by starting from different substituted pyridylketones and aldehydes.

1.4.2 MEDICAL, BIOLOGICAL AND OTHER APPLICATIONS: A SHORT OVERVIEW

As previously anticipated, imidazo[1,5-a]pyridine derivatives have found many different


medical applications. Interestingly, Kamal showed how this type of compounds can induce
apoptosis in human breast cancer cells by binding the DNA and therefore leading to an over
expression of two different tumor suppressor proteins.45 Other derivatives studied by the
same group, showed an even more wide range of applications, being effective against not
only human breast cancer cells, but also against melanoma, lung, colon, central nervous
system, ovarian, renal and prostate cancer, almost not affecting the non-cancerous cells
(Figure 1.22),46 which is a feature of great importance for the preparation of new
chemotherapy treatments.

30
Figure 1.22: two imidazo[1,5-a]pyridine derivates, studied for their antitumoral properties.

Imidazo[1,5-a]pyridine derivatives find applications also in other biomedical fields as


antibacterial,47,48 and for the treatment of many other diseases, among which it is worth
mentioning their use as effective brain penetrant drugs for the treatment of Alzheimer’s
disease.49
This class of compounds showed as well results concerning the possibility to be used as
Hg2+,50 H2S,51 and sulfides52 probes in living cells, as well as for the discrimination of
cysteine/homocysteine and glutathione.53 Also, it has been reported recently by Zhang a
coordination polymer derived from imidazo[1,5-a]pyridine capable of detecting Cu2+,
CrO42- and Cr2O72- ions in water solution.54 They can even be used as pH-sensors, noticeably
by changing the emission wavelength, quenching the fluorescence almost completely,55 or
ultimately varying the fluorescence quantum yield,56 with this feature being tested again
also in living cells.57
This class of molecules finds also application in the confocal microscopy field,58 a
microscopy technique which aims to obtain very high-resolution images. It can also obtain
3D representation of a sample by superposition of various 2D images taken at different
depths. This technique is often used in biological59 and materials sciences.60 To achieve this,
fluorophores are necessary in order to be excited by the light sent to the spot observed.
Quantum efficiency, brightness and excitation and emission spectra of the fluorescent
species are therefore key features to be considered, as well as solubility and
biocompatibility (especially in water for biological applications). As previously discussed,
imidazo[1,5-a]pyridine derivatives proved to possess all these properties in addition to the

31
possibility to modulate their emission wavelength by changing the substituents on the core
moiety. This led also to the possibility to easily obtain multi-colored bioimages.61

1.4.3 IMIDAZO[1,5-a]PYRIDINES AS LIGANDS

Imidazo[1,5-a]pyridines have been known for a long time for their ability to act as ligands
toward a large variety of metal centers, mainly to obtain luminescent coordination
compounds useful for the fabrication of optoelectronic devices, but also as catalysts. These
ligands, depending on the substituent on the main core, are capable of binding the
coordination center in various ways, starting from bidentate ligands,62 to tridentate63 and
even tetradentate64 (Figure 1.23). In some other cases, imidazo[1,5-a]pyridine based
carbenes can act as monodentate ligands (vide infra).

Figure 1.23: three examples of different binding modes displayed by imidazo[1,5-a]pyridine derivatives.

One of the most common substituents on the main core is the pyridinyl ring on the position
1: this permits a typical N,N bidentate coordination, which can be suitable for the
coordination of many transition metal ions, such as Mn2+,65 Ir3+, Cu+,66 Ag+, Zn2+, Ru2+ and
Os2+.67
On the other hand, a N,O coordination is possible when alcoholic68 or phenolic residues are
used as substituents. This, upon deprotonation of the hydroxyl group, has led to the
preparation of another series of coordination compounds, even with semimetals, like B(III)
centers.69

32
Usually, zinc and copper coordination compounds of imidazo[1,5-a]pyridines show
interesting luminescence properties. Volpi,70 recently reported a series of N∩N zinc
complexes of 1,3-substituted imidazo[1,5-a]pyridines with a maximum fluorescence
quantum yield of 33%, characterized by emission maxima in the blue region of the visible
spectrum (410-460 nm), noticing how the complexation enhanced the quantum yield itself
and resulted in a hypsochromic shift in the emission spectra with respect to the free
ligands. The authors proposed these compounds as tunable, air resistant low-cost emitting
materials for possible and desirable technological applications, as previously mentioned in
this introductory chapter.
Regarding copper compounds, it has been reported by Weber71 the synthesis of an
interesting complex displaying a blue fluorescence in solution and the largest shift ever
observed when built in a Luminescent Electrochemical Cell, obtaining a yellow emission.
Other metals have been used instead to obtain coordination compounds suitable to be
used as catalyst: nickel derivatives, like the one reported by Bluhm,72 proved to be active
in the dimerization of propene, with a 22% yield and a TON of 52.4ּ 103.
From 1968, when the first Nitrogen Heterocyclic Carbene (NHC) was prepared by Öfele,73
and from 1991, when Arduengo isolated the first stable carbene,74 a lot of efforts has been
devoted in the research for more of these ligands; as a matter of fact, their coordination
compounds demonstrated a huge potential as catalysts,75 finding applications in many
field, i.e. in Heck and Suzuki coupling, aryl amination, olefin metathesis, hydrosililation,
asymmetric catalysis and the preparation of supramolecular structures,76 among others.
As nitrogen heterocycles, also imidazo[1,5-a]pyridines have been used in the preparation
of several NHCs. This is because of the possibility to tune their electronic properties by
changing the residues on the main moiety.
There are two main different types of imidazo[1,5-a]pyridine based NHC: the first one
possesses the lone pair in position 3 (Figure 1.24-a), while the second, a mesoionic form,
has the pair in position 1 (Figure 1.24-b). Interestingly, this class of NHC showed a marked
π-accepting character with respect to other NHC ligands. This is due to the fact that in
addition to the empty p-orbital of the carbene itself, a π*-orbital located on the pyridinic
ring can act as an electronic density acceptor.77 The electronic properties can also be easily
tuned by changing the substituent in position 1,2 or 3 of the imidazopyridine ring.

33
Figure 1.24: synthetic pathways for the two NHCs derived from imidazo[1,5-a]pyridine.78

This category of NHC finds itself, as free molecules, many applications in organic chemistry,
like in the reaction, with aldehydes and dimethyl acetylenedicarboxylate or allenoates, to
give substituted furans,79 with phthalaldehydes and dimethyl acetylenedicarboxylate, to
obtain benzo[d]furo[3,2-b]azepines80 and, with ethyl propiolate and, again, dimethyl
acetylenedicarboxylate, to get respectively functionalized pyrroles and thiophenes.81
When used as ligands, imidazo[1,5-a]pyridine based NHCs form many complexes displaying
catalytic activity in various reactions. For example, Burstein82 reported a series of palladium
complexes capable of catalyze the Suzuki-Miyaura reaction of chloroarene, with yields of
up to 86% (Figure 1.25), while Chen83 reported the possibility of using nitro arenes as
starting material for the same reaction, by using sterically hindered and strong donating
NHC, in order to permit an oxidative addition at the arene-NO2 bond. They also
demonstrated to be efficient ligands in the preparation of palladium catalyst in Heck
reactions, managing to obtain quantitative yields.84,85

Figure 1.25: some of the compounds active in the Suzuki-Miyaura reaction.

34
Nickel complexes instead have been reported to be active in CO2 capture and utilization,
by C-H carboxylation of ethylene to obtain acrylates.86
Finally, gold(I) and silver(I) coordination compounds bearing this kind of NHC ligands
displayed in vitro anti-tumoral activities against ten different cancerous cell lines,87 as well
as fluorescence properties.88 Combining fluorescence and catalytic properties has also
been possible, as reported by Yagishita:89 an iridium complex has been used as
photocatalyst for the oxidative coupling reaction of benzylic amines to the corresponding
imines, obtained in quantitative yields.

1.4.4 LUMINESCENCE IN IMIDAZO[1,5-a]PYRIDINE DERIVATIVES

Imidazo[1,5-a]pyridine compounds are renowned for their luminescence properties


especially thanks to their usually large Stokes shifts,90 needed to avoid re-absorption, which
is an ideal feature for molecules to be used in optoelectronic devices or as down-shifting
materials.
Nevertheless, there are very few examples reporting imidazo[1,5-a]pyridines as emissive
layers in OLEDs,91 as well as in light-emitting electrochemical cells.92 However, many of
these derivatives showed high quantum yields and blue emission which, together with the
aforementioned large Stokes shift, are very important properties. Moreover, it is possible
to tune all these properties as well as their emission wavelength by changing the
substituents on the core moiety, and therefore their electronic properties. Shibahara93
reported a series of halogenated species with quantum yields up to 20%, but where the
correlation between substitution and electronic properties remained unclear, while in
many other cases and as it will be discussed in this Thesis, a correlation between
fluorescence properties and Hammet’s constants of the substituents has been reported.94
Mohbiya95 showed how the intramolecular charge transfer can be tuned as well, while
Salassa96 reported the possibility to tune the excited states of metal complexes by simply
changing the substituents on the imidazopyridine. In this case, more donor substituents
(i.e. dimethylamino group) resulted in an excited state with higher CT character.
On the other hand, extending the conjugation on the fluorescent core, usually results in a
fluorescence quantum yield increase, as reported by Albrecht,97 with imidazo[1,5-
a]quinolines up to three times more emissive than the respective imidazo[1,5-a]pyridines.
35
Also, the coplanarity between the imidazopyridine core and the substituents in position 1
and 3 can improve the quantum yield.
Steric effects have been considered as well in order to achieve highest fluorescence
quantum yields. Volpi98 reported the effect of the methoxy substituent in various position
of 1,3-disubstituted imidazo[1,5-a]pyridines, observing an increase up to 50% for those
species where the steric hindrance of the groups introduced allowed to block or at least
minimize the rotation of the rings attached to the core moiety.
The examples discussed above are reported in Figure 1.26.

Figure 1.26: Some of the imidazo[1,5-a]pyridine derivatives displaying fluorescence.

More information about fluorescence properties of imidazo[1,5-a]pyridine derivatives will


be discussed in Chapter 2.

36
1.5 FUORESCENCE IN BF2-CONTAINING COMPOUNDS99

One cannot write about BF2 luminescent compounds without considering the 4,4-difluoro-
4-bora-3a,4a-diaza-s-indacene, more known as boron dipyrromethene or, simply, as
BODIPY. This family of compounds has gained a lot of interest for being one of the most
applicable and versatile in chemistry, starting from 1968 when the first member of this
group was reported by Treibs and Kreuzer.100 Interestingly, these species have very high
quantum yields (up to 90%) and excellent chemical and photophysical stability in solution
as well as in solid state with fluorescence lifetime of 1 to 10 ns.
Basically, the high fluorescence, which is the main feature of this class of compounds, is
due to the very low flexibility of the system, with π-electrons conjugation running on the
whole organic backbone. Thanks to the latter, it is possible to finely tune the electronic
properties of the framework by adding suitable groups on one or both the pyrrole
fragments, as well as on position 8, called the meso-position; by tuning the electronic
properties in this way, of course absorption and emission of the species will vary as well,
being influenced by the extent of the electron delocalization.101,102 Generally speaking, by
extending the conjugation, a shift to lower energy of the emission wavelength is observed.
Ulrich103 reported near IR emission of BODIPY derivatives bearing in positions 3 and 5
anisole or ethylthiophene residues, with a corresponding decrease of the quantum yield
from 49 to 20%.
A large portion of these species finds its application as molecular sensor,104 in particular as
pH indicators in organic, aqueous and mixed aqueous-organic media. Usually, p-(N,N-
dialkyl)aniline subunit are contained in these species, leading to nonfluorescent derivatives
which develop strong emission upon protonation,105 paving the way to a new strategy for
in vivo tumour visualization: the probes are not fluorescent outside tumour cells at neutral
pH, but after being internalized by endocytosis in endosomes or lysosomes, the lower pH
(5-6) makes them highly fluorescent.106 BODIPY derivatives bearing phenols can be used as
pH sensors in the near-neutral pH range instead,107 while crown ether residues are used
for metal cation sensing, for example for potassium,108 sodium,109 calcium,110 zinc and
mercury,111 among others. Anions can be detected as well, like fluorides112 and cyanides113
and biomolecules like saxitoxin114 (component of shellfish poisons), peptides, proteins115
and monosaccharides.116 Very electron-perturbing substituent can be attached on the

37
BODIPY core, like N-oxides, used for the detection of oxyradicals in situ thank to the
decrease of the fluorescence after the reaction between the two.117
Of particular interest is the possibility to use modified BODIPY to localize viscosity on the
microscopic level, exploiting fluorescent molecular rotors based on this fluorophore: the
free rotation of an aromatic ring on the BODIPY backbone can influence the quantum yield
of the system by enhancing non-radiative deactivation of the excited state. As reported by
Kee,118 impeding the rotation of those rings would reduce the non-radiative decay,
ultimately leading to higher quantum yields.

Figure 1.27: Some of the previously mentioned BODIPY derivatives.

As previously mentioned, thanks to its outstanding properties, like high photo and chemical
stability, high absorption coefficients, high quantum yields and good solubility even in
aqueous media, this family of molecules is used in the fabrication of optoelectronic devices,
like OLEDs.119, 120 Unfortunately, in the solid state BODIPY derivatives show very small
Stokes shifts (5-20 nm), leading to serious self-quenching phenomena, as well as self-

38
aggregation, ultimately limiting their application. In order to solve this issues, bulky and
rigid side groups can be introduced on the core moiety.121 In some cases, white light-
emitting diodes have also been fabricated, exploiting ultra-thin layers of BODIPY
compounds, reaching a brightness of 6579 cd/m2 and an efficiency of 1.17%.122
Other classes of BF2 compounds have been investigated as well, like aza-BODIPY,123
pyrroles (BOPHY),124 hydrazones (BODIHY)125 and diketones,126 and, as it will be discussed
in this Thesis, also imidazo[1,5-a]pyridinyl phenols.

Figure 1.28: Other classes of BF2-containing compounds.

1.6 CHAPTERS OUTLINE

During the last decades, the coordination chemistry of imidazo[1,5-a]pyridines has been
widely investigated among our research group, starting from derivatives capable of binding
a metal center with a N,N fashion, to N,O monoanionic ligands. The fluorescent and
catalytic properties of these species have been studied and they will be briefly summarized
in Chapter 2.
As they looked more and more promising, the aim of this PhD project was to further extend
this class of compounds, focusing on their blue emission in order to find a suitable species
(or class of compounds) to be tested in the fabrication of OLEDs. This is because blue
emitters are still a major issue for OLED development, with current ones being not as
efficient as green or red counterparts.127 Zinc complexes and boron coordination
compounds will be presented and discussed respectively in Chapter 3, 4 and 5. As will be
discussed in the following chapters, all these species displayed bright blue fluorescence in
solution, and some of them also in the solid state and in polymeric films, with fluorescence
quantum yields reaching excellent values (up to almost 70%). Considering their properties,

39
these classes of molecules represented promising alternatives in the fabrication of blue
emissive OLEDs.
Finally, in Chapter 6 the fabrication and the properties of such devices built with the most
promising species will be discussed, with them being among the first examples reported in
the literature.

40
Chapter 2

PREVIOUS RESULTS ON IMIDAZO[1,5-a]


PYRIDINES AMONG OUR RESEARCH GROUP

41
2.1 COMPOUNDS WITH NEUTRAL N,N’-BIDENTATE 1-PYRIDYLIMIDAZO[1,5-
a]PYRIDINES AS LIGANDS

Among our research group, imidazo[1,5-a]pyridine derivatives have been largely studied.
Initially, a pyridyl substituent in position 1 on the heterocyclic backbone was used, in order
to obtain 1-pyridylimidazo[1,5-a]pyridines. The synthesis of these ones followed the
reaction path described in the previous chapter (Figure 1.20-d): two equivalents of
aromatic aldehydes (o-tolualdehyde or salicylaldehyde) were reacted with one equivalent
of di(2-pyridyl)ketone, using ammonium acetate as source of nitrogen in boiling acetic
acid.41 From this reaction, two different species, namely PIPMe and PIPOH, have been
prepared, which could act as N,N’-bidentate or N,N’,O-tridentate (R = OH) ligands (Figure
2.1). Using this ligands, Zn(II) and Ag(I) coordination compounds have been synthesized,
and their luminescence properties have been explored.

Figure 2.1: Synthesis of N,N’-bidentate 1-pyridylimidazo[1,5-a]pyridine ligands.

2.1.1 ZINC(II) HALIDO COMPLEXES


Species PIPOH and PIPMe have been reacted with one equivalent of ZnX2 (X = Cl, Br, I), in
methanol at room temperature to give respectively coordination compounds [ZnX2(PIPOH)]
and [ZnX2(PIPMe)] (Figure 2.2).128

42
R

R N R = Me: [Zn(PIPMe )X2 ]


N MeOH, r.t. N
N X
+ ZnX 2 R = OH: [Zn(PIPOH)X 2]
Zn
N X = Cl, Br, I
X
N

Figure 2.2: Synthesis of zinc(II) halido complexes with N,N-bidentate 1-pyridylimidazo[1,5-a]pyridines PIPMe
and PIPOH as ligands.

X-ray crystal structure analyses have been performed on [ZnCl2(PIPMe)], [ZnI2(PIPMe)] and
[ZnCl2(PIPOH)Cl2]: the results showed, as expected, the ligands displaying N,N’-bidentate
coordination mode, with the nitrogen atom of the pending pyridine (Npy) and the nitrogen
atom of the imidazolic ring (Nim) acting as coordination sites. The tetrahedral geometry of
the zinc atom is described by the τ4 parameters,129 whose calculated values are,
respectively, 0.87 for [ZnCl2(PIPMe)], 0.84 for [ZnI2(PIPMe)] and 0.89 for [ZnCl2(PIPOH)]. In
these three structures, the ligand molecules displayed a nearly planar configuration of the
1-pyridylimidazo[1,5-a]pyridine moiety (Figure 2.3).

Figure 2.3: Planes and main dihedral angles involving the 1-pyridylimidazo[1,5-a]-pyridine skeleton in
complexes [ZnCl2(PIPMe)], [ZnI2(PIPMe)] and [ZnCl2(PIPOH)] (a-c). Rotation of the phenolic residue in
[ZnCl2(PIPOH)], with O1 pointing away from Zn1 (c).

On the other hand, the plane defined by the aryl substituent in position 3 formed an angle
with the plane of the five-membered imidazole ring of 59.70° for [ZnCl2(PIPMe)], 65.35° for

43
[ZnI2(PIPMe)] and 47.27° for [ZnCl2(PIPOH)]. In the latter case, the rotation moves the oxygen
atom too far from the zinc center (Figure 2.3-c) to permit a coordination via hydroxyl group,
as also reflected by the sharp stretching band at 3280 cm-1 observed in the infrared spectra
of compounds [ZnX2(PIPOH)], attributed to the OH vibration. Also, the sharpness of this peak
let us exclude the presence of H-bond, in net contrast to what was observed for the
corresponding free ligands.128 The absence of H-bonding interaction in [ZnX2(PIPOH)]
species is critically important for the determination of their photophysical properties (vide
infra).
The absorption spectra recorded in dichloromethane solution (5 x 10-5 M) for compounds
[ZnX2(PIPMe)] and [ZnX2(PIPOH)] displayed a lower energy absorption band at about 370 nm,
with an intense shoulder in the range of 310-330 nm. The latter is characteristic of the free
ligands, thus suggesting a partial dissociation of the ligands in solution.
As no fluxional behavior has been observed, the possible temporary dissociation of the
pyridyl ring has been ruled out; for this reason, it is more likely that the dissociation
involved the whole ligand molecule, thus establishing an equilibrium with the starting
[ZnX2(PIPMe)] and [ZnX2(PIPOH)] complexes.
Using TD-DFT calculations,130 this absorption band has been determined to be due almost
exclusively to a HOMO-LUMO (>98%) transition. HOMO orbitals were mainly delocalized
on the imidazo[1,5-a]pyridine and pyridinic moieties, with lesser contribution of the aryl
substituent, while LUMO orbitals were spread almost only on the heterocyclic rings (Figure
2.4). The main absorption of these complexes has been attributed to intraligand (π–π*)
transitions. The energy of the frontier orbitals (HOMO = ca. -6.1 eV, LUMO = ca. -2.1 eV)
are very similar to those of the free ligands and among the series, with the HOMO-LUMO
gap almost constant throughout the two series of compounds.

Figure 2.4: Frontier molecular orbitals for compound [ZnCl2(PIPMe)].

44
As these compounds displayed partial dissociation in solution, their photophysical
properties have been only studied in the solid state and the data collected are summarized
in Table 2.1. All these complexes displayed blue emission in the solid state (Figure 2.5),
with emission maxima centered at about 443 nm for [ZnX2(PIPMe)], whereas [ZnX2(PIPOH)]
showed a bathochromic shifts up to 482 nm ([ZnCl2(PIPOH)]). The highest Stokes shifts
recorded have been 0.39 and 0.38 eV respectively for [ZnCl2(PIPOH)] and [ZnBr2(PIPOH)].

Figure 2.5: Solid-state fluorescence of compounds [Zn(PIPOH)X2] and ligand PIPOH. Inset: Blue fluorescence at
the solid state for [Zn(PIPMe)Cl2] (a) and [Zn(PIPOH)Cl2] (b).

λexc λem ΦPL


(nm) (nm)
PIPMe 418 442 0.66
PIPOH 400 452 0.15
PIPOD 400 430 0.38
[ZnCl2(PIPMe)] 398 442 0.40
[ZnBr2(PIPMe)] 413 443 0.21
[ZnI2(PIPMe)] 397 443 0.07
[ZnCl2(PIPOH)] 419 482 0.29
[ZnBr2(PIPOH)] 418 480 0.18
[ZnI2(PIPOH)] 423 444 0.10
Table 2.1: Solid-state luminescence properties of [ZnX2(PIPR)] compounds and ligands PIPMe and PIPOH used
in their synthesis.

Also, the fluorescence quantum yields linearly decreased with increasing molar mass of the
compounds (the latter being related to the atomic number of the halogen atom bound to

45
the zinc center) (Table 2.1 and Figure 2.6); this causes a spin–orbit coupling enhancement,
which in turn facilitates radiationless intersystem crossing.
[ZnX2(PIPMe)] displayed lower quantum yields with respect to those of the corresponding
free ligand (Table 2.1), while compounds [ZnX2(PIPOH)] were generally more emissive. This
behavior could be explained by considering the crystal structures of these derivatives:
ligand PIPOH showed intermolecular OH-N hydrogen bonds between the hydroxyl group on
the phenolic ring and the imidazolic nitrogen atom of an adjacent molecule. This interaction
surely favors non-radiative thermal relaxations, resulting in lower quantum yields when
compared to the coordination compounds in which no hydrogen bond can be formed.
Upon deuteration of PIPOH to give PIPOD, the fluorescence quantum yields rose again,
reaching similar values of those observed for [ZnX2(PIPOH)].
This trend is better visualized in Figure 2.16, where it is possible to see the clear
discordance in the behavior of PIPOH when compared with PIPMe. Indeed, the quantum yield
of free PIPMe aligns well with those of the corresponding complexes, and the ΦPL value of
free PIPOH does not follow this trend. A possible interpretation of the different behavior of
these two ligands can be found in their crystal structures: in the solid state, PIPOH can be
described as a catemeric 1D, whereas PIPMe shows a discrete molecular assembly. It is
possible that the large extent of H-bond connectivity in PIPOH favors thermal relaxation,
which causes a dramatic decrease of the fluorescence emission of PIPOH with respect to
that of PIPMe. In addition, in solution (CH2Cl2, 5x10–5 M), PIPOH shows a remarkable increase
in emission intensity and has a ΦPL value double that in the solid state (0.37 and 0.15
respectively).
Also, the deuterated species (PIPOD) showed a solid-state ΦPL value definitely higher than
that of PIPOH. Moreover, when the ΦPL value of PIPOD is correlated to the molar mass, a
good alignment with the [ZnX2(PIPOH)] complexes is noticed. The odd collocation of the
PIPOH quantum yields with respect to its zinc(II) halide adducts has to be ascribed to the
abovementioned intermolecular O–H···N interactions in the crystal structure, which result
in a non-radiative relaxation process. When PIPOH is coordinated to zinc, the capability to
produce H-bonds is hindered, and the spin–orbit coupling due to the heavy atom
predominantly controls the relaxation processes.
In conclusion, these classes of zinc(II) complexes, obtained using 1-pyridylimidazo[1,5-
a]pyridines as ligands, represent interesting examples of blue emissive materials in the

46
solid state, with photoluminescence performances (i.e. fluorescence quantum yield)
correlated to their structure.

Figure 2.6. Linear correlation between the absolute quantum yields and molar masses for compounds
[ZnX2(PIPMe)] and ligand PIPMe (blue circles) and [ZnX2(PIPOH)] and ligand PIPOD (red squares). Green triangle:
absolute quantum yield of PIPOH.

2.1.2 ZINC(II) COMPLEXES WITH POLYATOMIC ANIONS


Considering the abovementioned results, the exploration on zinc(II) luminescent
coordination compounds of pyridylimidazo[1,5-a]-pyridine ligands has been continued,
going from heteroleptic compounds to homoleptic complexes, reacting two equivalents of
PIPOH with one equivalent of ZnY2 (Y = polyatomic, low- or non-coordinating anion, ClO4-,
NO3-, BF4-).131 The syntheses have been performed by treatment of a solution of ZnY2 in
methanol with PIPOH, in a 1:2 metal:ligand molar ratio. Depending on the starting zinc salt,
three different compounds have been obtained, namely [Zn(PIPOH)2(ClO4)](ClO4)·H2O,
[Zn(PIPOH)2(NO3)](NO3)·H2O and [Zn(PIPOH)2(H2O)2](BF4)2·2H2O. A fourth complex,
[Zn(PIPOH)2(H2O)](PF6)2 has been prepared by reacting [Zn(PIPOH)2Cl2] with AgPF6 in
acetonitrile (Figure 2.7).
Similarly to the previously discussed class of compounds, the very broad band at 3300 cm-
1 in the infrared spectra was attributed to OH vibrations. Moreover, in this case a broad
stretching band appeared in the region between 1300 and 1000 cm-1, characteristic of the
polyatomic anion. The water content was assessed instead using thermogravimetric
analyses (DSC/TGA).

47
Figure 2.7: Synthesis of zinc(II) complexes with N,N’-bidentate 1-pyridylimidazo[1,5-a]pyridines PIPOH as
ligand.

The 1H-NMR spectra of these complexes displayed the expected pattern for coordinated
PIPOH, accompanied by some low-intensity peaks that have been attributed to the free
ligand: for this reason, in accordance with what observed for the previously studied
compounds, a possible lability of these complexes in solution has been theorized.
The proposed formulas have been confirmed by X-ray crystal structure analyses (Figure
2.8): in [Zn(PIPOH)2(ClO4)](ClO4)·H2O, the zinc center displayed a penta-coordinated
geometry in a N4O environment, with the two ligands coordinated in a N,N’-bidentate
mode and the fifth position occupied by an oxygen atom of the perchlorate. The geometry
is nearly trigonal bipyramidal, with a trigonality index τ = 0.73.132 Analogously to what has
been observed in the previous zinc(II) halido complexes, the 1-pyridylimidazo[1,5-
a]pyridine moiety of the ligands showed a planar conformation, forming an angle with the
plane of the phenolic ring of ca. 55° with the hydroxyl group directed far from the
coordination center.
Attempts to obtain single crystals of [Zn(PIPOH)2(H2O)2](BF4)2 from acetonitrile led to the
crystallization of the species [Zn(PIPOH)2(CH3CN)2](BF4)2. Also in this case, a nearly planar
arrangement of the 1-pyridylimidazo[1,5-a]pyridine moiety and absence of interaction
between zinc and the hydroxy group were observed (Figure 2.8).

48
Figure 2.8: Molecular representation of complexes and [Zn(PIPOH)2(CH3CN)2](BF4)2 (left) and
[Zn(PIPOH)2(ClO4)](ClO4)·H2O (right).

The photoluminescence properties in the solid state of these species have been
investigated as well. They all displayed blue to green fluorescence emission, with maxima
ranging from 468 to 500 nm (Table 2.2). A bathochromic shift of more than 40 nm has been
observed with respect to the free PIPOH ligand.
The main electronic transition responsible for the fluorescence properties has been defined
as ligand centered (LC) π-π*. All the excitation maxima are in the region between 450-470
nm, except for the nitrato derivative, for which a λexc at 421 nm has been recorded. The
fluorescence quantum yields are lower than those recorded for the previous class of
compounds.

λexc (nm) λem ΦPL


(nm)
[Zn(PIPOH)2(ClO4)](ClO4)·H2O 452 493 0.14
[Zn(PIPOH)2(NO3)](NO3)·H2O 421 468 0.16
[Zn(PIPOH)2(H2O)2](BF4)2·2H2O 451 491 0.09
[Zn(PIPOH)2(H2O)](PF6)2 452 494 0.12
Table 2.2: Solid-state luminescence properties of zinc(II) compounds with 1-pyridylimidazo[1,5-a]pyridine
ligands and polyatomic anions.

2.1.3 HETEROLEPTIC SILVER(I) COMPLEXES


Silver(I) compounds of N,N-bidentate 1-pyridylimidazo[1,5-a]pyridine of general formula
[Ag(PIPOH)(PR3)]NO3 have been synthesized as well, using phosphines as ligands.133 The
synthetic pathways followed the procedure originally developed by Hor134 and exploited by
Nierengarten135 to obtain similar complexes (Figure 2.9). One equivalent of AgNO3 has

49
been treated with one equivalent of PR3 in a mixed solvent solution
dichloromethane:methanol 5:1, affording the polymeric species {Ag(PR3)(NO3)}n, which,
upon addition of one equivalent of PIPOH, gave the desired products.

Figure 2.9: Synthesis of silver(I) complexes with N,N-bidentate 1-pyridylimidazo[1,5-a]pyridines PIPOH and
phosphinic ligands.

The infrared spectra (ATR) of these compounds displayed a broad band at ca. 1320 cm-1,
characteristic of the uncoordinated nitrate anions, while a second undefined band at 2800-
3000 cm-1, attributed to the free hydroxylic group, let us exclude the involvement of the
phenolic oxygen atom in the coordination to the metal center, as also corroborated by X-
ray structural analyses. In fact, single crystal analysis performed on [Ag(PIPOH)(PPh3)]NO3,
[Ag(PIPOH)(PMe2Ph)]NO3, [Ag(PIPOH)(PBu3)]NO3 and [Ag(PIPOH)(P{OPh}3)]NO3 allowed to
identify a trigonal geometry at the silver center in a N,N,P environment. In these structures,
the phenolic ring was tilted with respect to the plane of the heterocyclic moiety of an angle
of ca. 50°.
Among the others, the distance between the phosphorous atom and the plane defined by
the silver center and the two coordinating nitrogen atoms was a relevant parameter (Figure
2.10). Compound [Ag(PIPOH)(PPh3)]NO3 shows a higher value with respect to the others,
with the phosphorous atom out of such plane, as better described using the centroid-Ag-P
angle. Considering this aspect, it could be concluded that the geometry around the silver
in the PPh3 is more distorted from the ideal planar geometry than what has been found for
the other complexes.

50
Figure 2.10: Red dashed line: distance between the P atom and the Ag,N,N plane (Å); black dot: centroid of
the Ag,N,N plane.

31P{1H}-NMR experiments (CD2Cl2, 193 K) let us determine that the described geometries
were maintained also in solution. J(107Ag-31P) and J(109Ag-31P) have been determined to be
respectively around 641-655 Hz and 738-767 Hz, indicative of a trigonal geometry for the
silver atom.136 Conductivity measurements conducted in nitromethane allowed to confirm
the salt nature of this class of compounds, with a molar conductivity in the range of what
was expected for 1:1 electrolyte (75-90 ohm-1 cm2 mol-1).137
The photophysical properties of these species have been studied in dichloromethane
solution (10-5 M). Two main absorption bands in the UV-vis spectra have been recorded,
respectively at 325-330 nm and at 359 nm (Table 2.3). As for the previously discussed
classes of 1-pyridylimidazo[1,5-a]pyridine compounds, the higher energy absorption
strongly resembled that of the free ligand, thus suggesting a possible dissociation of PIPOH
in dilute solutions, as also later confirmed by fluorescence lifetime measurements (vide
infra). The low-energy absorption has been described using TD-DFT calculations as mostly
HOMO-LUMO transition (>90%), with an intra-ligand π–π* character. Frontier orbitals have
been determined to be spread over the whole molecules, with only a minor contribution
of the phenolic ring in the HOMO topology. Notably, the energy gap remained almost
constant throughout the series (ca. 2.50 eV).

51
Figure 2.11: Frontier molecular orbitals for the trigonal cation [Ag(PIPOH)(PPh3)]+.

All these compounds displayed a bright fluorescence in solution, with emission maxima in
the blue region, between 441 and 465 nm. Lifetimes of few nanoseconds have been
calculated, which, at a deeper analysis, resulted in biexponential decay curves with τ1 = 1.1-
1.4 ns and τ2 = 2.9-3.3 ns. The latter resulted to be very close to the values recorded for the
free ligands, further corroborating the hypothesis of a partial dissociation in solution. Good
quantum yields (0.30-0.38) have been measured, with [Ag(PIPOH)(PPh3)]NO3 even showing
a higher value (0.52).
Blue-green fluorescence in the solid state has been recorded as well, with the emission
maxima distributed over a wider range than those recorded in solution (449-498 nm).
Notably, the maximum of emission is correlated to the nature of the ancillary ligand (Figure
2.12), with a general bathochromic shift in the presence of electron-withdrawing
phosphines. PPh3-containing derivative displayed the highest absolute quantum yield
among the series also in the solid state, with a value of 0.68, while the other compounds
experienced a diminishing with respect of their performances in solution (ФPL = 0.14-0.20).
This behavior was related to the higher degree of distortion from planarity in
[Ag(PIPOH)(PPh3)]NO3: it is in fact known that tetrahedral silver(I) compounds usually
display better fluorescence properties than trigonal compounds. For this reason, the
absolute quantum yield recorded for [Ag(PIPOH)(PPh3)]NO3 was among the highest ever
recorded for mononuclear trigonal silver(I) compounds.

52
Figure 2.12: CIE 1931 chromaticity plots for the fluorescence emission of [Ag(PIPOH)(PR3)]NO3 derivatives in
dichloromethane solution (left) and in the solid state (right).

solution solid state


λexc λem ΦPL τ λexc λem ΦPL τ
(nm) (nm) (ns) (nm) (nm) (ns)
[Ag(PIPOH)(PPh3)]NO3 360 433 0.52 1.3, 3.3 360 473 0.68 1.9, 3.8
[Ag(PIPOH)(PMe2Ph)]NO3 357 461 0.38 1.4, 3.2 357 449 0.18 1.3, 2.6
[Ag(PIPOH)(PMePh2)]NO3 358 460 0.37 1.4, 3.2 358 490 0.18 1.6, 4.9
[Ag(PIPOH)(P(p-tolyl)3)]NO3 375 442 0.38 1.3, 3.2 375 471 0.20 1.0, 3.8
[Ag(PIPOH)(PBu3)]NO3 357 462 0.36 1.3, 3.2 357 463 0.14 1.5, 3.6
[Ag(PIPOH)(P(OPh)3)]NO3 377 441 0.30 1.1, 3.0 377 498 0.16 2.1, 4.1
[Ag(PIPOH)(P(OEt)3)]NO3 375 442 0.36 1.2, 2.9 275 480 0.19 2.2, 5.2
Table 2.3: Photophysical data for compounds [Ag(PIPOH)(PR3)]NO3.

2.2 COMPOUNDS WITH ANIONIC N,O-BIDENTATE 2-(IMIDAZO[1,5-


a]PYRIDIN-3YL) PHENOLS
As previously described, exploring the coordination chemistry of N,N’-bidentate
imidazo[1,5-a]pyridine derivatives resulting in the preparation of several luminescent
compounds, characterized by high Stokes shifts and good to high quantum yields. On the
other hand, these compounds displayed a ligand dissociation of the pendant pyridyl moiety
in solution. For this reason, we decided to substitute the pyridyl ring in position 1 using a
non-coordinating residue (Me, Ph). To achieve such compounds, di(2-pyridyl)ketone was
replaced by 2-acetylpyridine and 2-benzoylpiridine respectively, leading to two new classes
of 2-(imidazo[1,5-a]pyridin-3-yl)phenols (HLR and HLR-Ph). Upon deprotonation, these
compounds could act as N,O-bidentate monoanionic ligands.

53
2.2.1 CATALYTIC ACTIVITY OF HOMOLEPTIC COMPLEXES OF DIVALENT METALS
These new classes of compounds have been used in the preparation of coordination
compounds of divalent metals (cobalt(II), copper(II), nickel(II) and palladium(II)). All these
species did not displayed any dissociation of the ligands in dilute solution. Therefore, they
have been studied as potential catalysts for the Heck reaction, obtaining selectivity up to
100% and conversions up to 98% for the Ni(II) and Pd(II) compounds.138
As this Thesis deals with the synthesis and characterization of luminescence coordination
compounds, even if the catalytic properties of this class of derivatives were very
interesting, they will not be further discussed.

2.2.2 TETRAHYDRO-2-(IMIDAZO[1,5-a]PYRIDIN-3-YL) PHENOLS


As previously mentioned, the properties of imidazo[1,5-a]pyridine derivatives are widely
known in the literature. Despite this fact, no examples on the study of the photophysical
properties of their hydrogenated derivatives tetrahydroimidazo[1,5-a]pyridines have been
reported, even if they are already well-known both for their biological application139 and
for their role as synthetic intermediates.140
For this reason we decided to investigate the photophysical properties of the hydrogenated
counterparts of the previously described imidazo[1,5-a]pyridinyl phenols, obtained by
catalytic hydrogenation of HLR derivatives (Figure 2.15).141

Figure 2.15: Synthesis of hydrHLR derivatives from HLR. * NH2 obtained by hydrogenation of HLNO2.

This new class of compounds proved to be highly fluorescent in solution (dichloromethane


5 x 10-5 M). Their UV-vis spectra presented similar traces, displaying three different main
absorption bands at 240-245, 280-290 and 314-344 nm.

54
The substituent R in the para position to the hydroxyl group on the phenolic ring influenced
both the emission maxima and intensity in these compounds, moving from deep blue to
green with increasing the electron donating properties of the substituent (Table 2.4).

λabs λexc λem Stokes shift τ


R ΦPL
(nm) (nm) (nm) (eV) (ns)
H 314 314 452 1.21 0.63 3.5
Me 321 321 473 1.24 0.43 2.3
1.0 (95.8%)
OMe 332 333 503 1.26 0.11
3.8 (4.2%)
F 326 326 464 1.13 0.72 4.9
CF3 316 327 432 0.92 0.61 3.7
0.2 (80.8%)
NH2 344 343 537 1.31 <0.05
3.5 (19.2%)
Table 2.4. Photophysical data for compounds hydrHLR in solution (CH2Cl2, 5·10-5 M).

The fluorescence lifetimes are between 2.3 and 4.9 ns, with the compounds bearing the
most electron-donating substituents (i.e. OMe and NH2) displaying biexponential decays.
Also, the absolute fluorescence quantum yields of these two compounds are definitely
lower than those of the other compounds in this series, which displayed excellent values
instead.
As these compounds looked very promising, their photophysical properties have been
studied also in thin polymeric films, using poly(methylmethacrilate, PMMA) as host
polymeric material. The absorption and emission spectra of hydrHLR compounds perfectly
matched those recorded in solution.
Using TD-DFT calculations, the lower energy absorption has been defined as a HOMO-
LUMO transition of all these species (Figure 2.16). The orbital topology has been
determined, showing both HOMO and LUMO being spread over the whole aromatic parts
of the molecules. Notably, the HOMO orbitals showed some slight difference among the
series, thus influencing the emissive properties. In order to describe this influence, a linear
correlation between the electronic properties of R (described using Hammet’s σP) and the
emission maxima has been found, with a bathochromic shift moving to longer wavelength
with decreasing σP. Also, another linear correlation between σP and the HOMO energy
could be found, with the energy increasing with more electron donating groups.

55
As the energy is related to the emission, a further investigation was performed, calculating
the percentage distribution of the orbitals in the two moieties of the molecules, the
heterocyclic ring and the phenolic ring (Figure 2.16). Notably, the higher is σP, the more the
HOMO was found to be localized on the imidazo[1,5-a]pyridine moiety, demonstrating that
the different topology led to different fluorescence behavior.

Figure 2.16: Frontier molecular orbitals for compounds hydrHLOMe (top) and hydrHLH (center) and
contribution of phenolic and imidazopyridine rings to the orbital topology (bottom).

56
Chapter 3

N,O-HOMOLEPTIC ZINC(II) COMPLEXES


([Zn(LR)2])

57
As described in the previous chapter, zinc has always had an important role in our research,
as cheap, non-toxic, easily accessible. For this reason, this Thesis project started with the
synthesis and characterization of a series of homoleptic zinc(II) complexes using
imidazo[1,5-a]pyridine derivatives as ligands,142 since they proved to give fluorescent
coordination compounds.

3.1 SYNTHESIS AND CHARACTERIZATION

The synthesis of the ligands (imidazo[1,5-a]pyrid-3-yl)phenol, HLR (R = H, OMe, Me, F, Br,


CF3, NO2) followed the pathway depicted in Figure 1.20-d, namely the reaction of 2-
acetylpyridine with two equivalents of 5-substituted salicylaldehyde and five equivalents
of ammonium acetate in acetic acid, obtaining all the ligands in moderate to good yields
(47-68%). 5-(trifluoromethyl)salicylaldehyde, not commercially available, has been
prepared following a method reported by Geneste (Figure 3.2).143

Figure 3.1: Synthesis of the HLR ligands.

Figure 3.2: Synthesis of 5-(trifluoromethyl)salicylaldehyde.

58
The preparation of the zinc coordination compounds [Zn(LR)2] was carried out with two
equivalents of the ligand and one equivalent of zinc perchlorate hexahydrate in acetonitrile
with a small excess of triethylamine to deprotonate the phenolic ring of the
imidazopyridine derivatives (Figure 3.3).
R

R
O
N
2 N
OH Zn
Et3N N
N N
N + Zn(ClO4)2 6H2O O
CH3CN

Figure 3.3: Synthesis of [Zn(LR)2].

The yellow-brown crystalline products have been characterized by elemental analysis,


infrared, 1H-NMR and 13C-NMR (Figure 3.4-3.5).

Figure 3.4: 1H-NMR of [Zn(LH)2] in DMSO-d6.

59
130.91
126.54
122.77
122.23
119.66
118.48
115.05
113.88

Figure 3.5: 13C-NMR of [Zn(LH)2] in DMSO-d6.

The IR spectra showed the typical pattern of the ligands with intense and sharp bands
between 1515-1620 cm-1, attributed to the stretching of the C-N double bonds in the
heterocycles.
In the 1H-NMR spectra it is possible to observe the signals of the corresponding ligands,
with some shifts due to the coordination to the metal center (Figure 3.6). Unfortunately,
though, all these compounds were pretty insoluble, therefore in the 13C-NMR there are no
information about the quaternary carbons.

Figure 3.6: Comparison between the 1H-NMR spectra of free ligand HLH (above) and complex [Zn(LMe)2]
(below), enlargement of the aromatic region.

60
The expected tetrahedral geometry at the zinc center was eventually confirmed by the X-
ray crystal structure analysis performed on complex [Zn(LH)2] (Figure 3.7), which have been
the only compounds we managed to obtain crystals of, while, unfortunately, every other
attempt to grow single crystals failed because of their very low solubility. [Zn(LH)2]
crystallized in EtOH as a hemi-solvate compound, with the metal center bound via N-O
coordination by two ligand molecules. Noteworthy, [Zn(LH)2]·0.5EtOH showed two
different molecules in the asymmetric unit, with one of them displaying a hydrogen bond
with an ethanol molecule, with a distance between the phenolato and hydroxyl oxygen
atoms of 2.786 Å. (Figure 3.8).

Figure 3.7: ORTEP representation of [Zn(LH)2] at 50% probability level.

Figure 3.8: Detail of the coordination centers of the two molecules in the [Zn(LH)2] crystal cell.

61
Length Angle Angle
Atom1 Atom2 Atom1 Atom2 Atom3 Atom1 Atom2 Atom3
(Å) (°) (°)
N1 Zn1 1.982(7) N1 Zn1 N2 127.4(3) N5 Zn2 O4 118.1(3)
N2 Zn1 1.996(7) N1 Zn1 O1 113.4(3) N6 Zn2 O3 118.6(3)
O1 Zn1 1.905(6) N1 Zn1 O2 97.4(3) N6 Zn2 O4 95.4(3)
O2 Zn1 1.903(7) N2 Zn1 O1 96.6(3) O3 Zn2 O4 116.1(3)
N5 Zn2 1.991(7) N2 Zn1 O2 109.0(3)
N6 Zn2 1.965(8) O1 Zn1 O2 113.7(3)
O3 Zn2 1.927(6) N5 Zn2 N6 114.7(3)
O4 Zn2 1.916(6) N5 Zn2 O3 95.7(3)
Table 3.1: Bond lengths and angles around the metal centers for [Zn(LH)2] obtained by X-ray diffraction.

The zinc centers displayed distorted tetrahedral geometries with calculated τ4 being 0.843
and 0.874 respectively.129 Moreover, the ligands are not planar, with the phenolic ring
twisted with respect of the imidazopyridine moiety by an angle of 35-39°.

3.2 OPTICAL PROPERTIES IN SOLUTION

The photophysical properties of all these species have been studied in solution and
compared to those recorded for the free ligands (Table 3.2, Figure 3.9). 5 x 10-5 M solutions
in dichloromethane have been prepared and, firstly, UV-vis, emission and excitation
spectra were recorded (Figure 3.10).
The UV-vis spectra of these species are quite similar one to another, with two main
absorption bands in the intervals between 240-260 and 340-360 nm, resembling the
spectra of the respective free ligands (Figure 3.9).

λabs λex λem Stokes shift ΦPL


R τ (ns) ΦPL
-1
(nm) (nm) (nm) (cm ) (eV) (ligand)

H 340 356 490 7662 0.95 2.8 0.33 0.14


Me 354 353 496 8146 1.01 2.7 0.22 0.11
OMe 360 385 527 7017 0.87 2.2 0.10 0.07
F 349 367 491 6856 0.85 3.1 0.23 0.13
Br 362 367 484 6614 0.82 2.4 0.13 0.10
CF3 350 350 473 7420 0.92 2.8 0.21 0.12
NO2 350 397 615 8953 1.11 - < 0.05 < 0.05
Table 3.2: Photophysical data of compounds [Zn(LR )2] recorded in dichloromethane solution (5·10-5 M).

62
0.9
0.8
0.7

0
Absorbance/10
0.6
0.5
0.4
0.3
0.2
0.1

200 300 400 500 600 700


Wavelength/nm

Figure 3.9: Comparison between the UV-vis spectra of complex [Zn(LH )2] (blue) and free ligand HLH (red) in
dichloromethane solution (5·10-5 M).

Figure 3.10: Normalized UV-vis (black), excitation (red) and emission (blue) spectra of compounds [Zn(LR )2]
recorded in dichloromethane (5 x 10-5 M) solution.

63
Noteworthy, very high Stokes shift of 0.82-1.11 eV have been recorded, with a minimal
superposition of the emission and excitation spectra, a very important feature for species
that aim to be exploited in optoelectronic devices. These properties let us think it would
have been possible to transfer the electronic and photophysical properties of the ligands
to the zinc complexes, with also an enhancement of efficiency due to the coordination. For
this reason, we expected a possible influence of the substituent R on the phenolic ring on
the photophysical behavior of these species, influence that has been eventually found in
the emission spectra of the series: indeed, the electronic character of R influenced both the
intensity and the maxima of emission, the latter ranging from 456 to 615 nm (Figure 3.11).
The excitation spectra instead, faithfully reproduced the respective UV-vis spectra.

2.5 0.9
0.8
2.0 0.7
0
6

0.6
Counts/10
Counts/10

1.5
0.5
0.4
1.0
0.3

0.5 0.2
0.1
0.0 0.0
400 450 500 550 600 650 700 750 800 400 450 500 550 600 650 700 750 800
Wavelength/nm Wavelength/nm

Figure 3.11: Non-normalized (left) and normalized (right) emission spectra of compounds [Zn(LR )2] recorded
in dichloromethane (5 x 10-5 M) solution. Color code: R = H, black; R = Me, blue; R = OMe, green; R = F,
orange; R = Br, red; R = NO2, pink; R = CF3, purple.

Time-dependent analyses have been performed as well and we determined for all these
species a lifetime decay of few nanoseconds (Figure 3.12, Table 3.2), excepting for
[Zn(LNO2)2] for which it was not possible to measure because of its very poor emissive
response to the pulsed source. All the other compounds showed a mono-exponential decay
curve and lifetimes from 2.2 to 3.1 ns, similarly to those displayed by the free ligands. For
this reason, we hypothesized that the zinc atom was not involved in any electronic
transitions and therefore in any emissive process, as later confirmed by DFT calculations
(see Chapter 3.3). Indeed, the metal center only played a structural role in this type of
compounds, enhancing the photophysical properties by giving rigidity to the whole system
and preventing any possible rotational relaxation. This is confirmed by the fluorescence
quantum yields, which are greater than those of the respective ligands for all the

64
compounds, except for R = OMe and NO2. When R = OMe, a poor emission in solution was
recorded, while the nitro derivative displayed almost no emission at all as a consequence
of the introduction of the nitro group, which is known to act as a fluorescence quencher.144

103
Counts

102

101

100
0 5 10 15 20 25 30 35 40 45
Time/ns
Residuals

4.0
Figure 3.12: Fluorescence decay of [Zn(LH )2] recorded in CH2Cl2 (5 x 10-5 M) solution.

3.3 DFT CALCULATIONS

The electronic transitions responsible for the absorption and emission phenomena in this
class of compounds have been further studied using TD-DFT calculations. Firstly the
geometry of all the complexes have been optimized, taking as a starting point the structure
of compound [Zn(LH)2] obtained by the single crystal X-ray diffraction experiment. The
calculated geometries were always in agreement with the atom coordinates of the
experimental structure. Starting from this it has been possible to simulate the UV-vis
spectra for all the compounds and to calculate the frontier molecular orbitals (Figure 3.13).

Figure 3.13: Experimental (red) vs. calculated (dashed blue) UV-vis spectra of [Zn(LMe)2] calculated in
dichloromethane solution.

65
The molecular frontier orbitals’ shape is almost the same for every complex of this series,
with the HOMO orbital being delocalized over the entire molecule; on the contrary, LUMO
orbitals are distributed almost only on the imidazopyridine moiety (Figure 3.14). As
expected from the different luminescence properties, the [Zn(LNO2)2] complex displayed a
different orbital topology, with the LUMO being localized only on the phenolic ring with no
contribution of the heteroaromatic skeleton.

Figure 3.14: Calculated HOMO and LUMO orbitals of compounds [Zn(LR)2].

The vertical transition associated to the absorption process in these species has been
defined as a HOMO-LUMO transition in each complex, including the nitro substituted one.
This is better visualized using Natural Transition Orbitals (NTOs) (Figure 3.15), where the
donor orbital is superimposable to the HOMO orbital, being again distributed over the
whole molecule, while the virtual acceptor orbitals’ topology being the same of the LUMOs.

66
For [Zn(LNO2)2], the donor orbital followed the same trend of this series of compounds,
while the acceptor orbital is spread over the phenolic fragments of the two ligands.

Figure 3.15: NTOs associated to the main electronic transition (>95%) S0 → S1 in [Zn(LH)2] and [Zn(LNO2)2].

As previously theorized, the metal center is not involved in any electronic transition, thus
having a merely structural role.
Looking at the emission spectra and their variability in both intensity and emission maxima,
a correlation between the electronic properties of the ligands and the photophysical
properties could be expected, in particular between the calculated energies of the orbitals
involved in the electronic transitions and the electron acceptor/donor properties of the
substituents R, which are the only reason for the differentiation of the ligands. As
descriptor of the latter, we decided to use Hammett σp constants.145
As the HOMOs are the only orbitals with a clear contribution of the phenolic ring bearing
the R substituents (except for R = NO2), we started our investigation of this aspect looking
for a relationship between the HOMOs energy and σp, finding indeed a linear correlation,
with the energies increasing with increasing the donor properties of the substituents, or
with decreasing the Hammett constants. It can be concluded from this, that it is possible to
modulate the HOMO energy of these system by changing the electronic character of R
(Table 3.3).
Moreover, a similar emission maximum is displayed by the complexes bearing R = H, F and
Br, even if these substituents correspond to very different σp. The reason for this behavior

67
can be traced back to a linear correlation found between the HOMO-LUMO energy gap and
the λem. Indeed, the complex with R = H, F and Br possess very similar energy gaps.

R EHOMO (eV) ELUMO (eV) H-L gap (eV) σp

H -5.580 -1.353 4.227 0


OMe -5.403 -1.371 4.032 -0.27
Me -5.489 -1.395 4.194 -0.17

F -5.616 -1.395 4.221 0.06

Br -5.647 -1.422 4.225 0.23

CF3 -5.766 -1.442 4.324 0.54


NO2 -5.902 -2.467 3.435 0.78
Table 3.3: Calculated HOMO and LUMO energies for the compounds [Zn(LR)2].

To conclude, it can be said that the electronic nature of substituent R is responsible for the
differences in the HOMO energy, and hence in the HOMO-LUMO energy gap, ultimately
leading to different maxima of emission of these zinc compounds, which appear to possess
some grade of tunability (Figure 3.16).

Figure 3.16: Correlation between HOMO energies vs. Hammett constants (top) and maxima of emission vs.
HOMO-LUMO energy gap (bottom) (R = H, F, Br are highlighted in the red circle).

68
Considering the properties described in this chapter, compounds [Zn(LR)2] represent an
interesting class of zinc(II) coordination compounds, displaying a blue emission in solution
accompanied by good fluorescence quantum yields, high Stokes shifts and very low
superposition between the excitation and emission spectra. A possible use of these species
as emissive materials in optoelectronic application can be considered, especially, due to
their very low solubility in organic solvent, using the vapor deposition method.

69
Chapter 4

BORON DIFLUORIDE COMPOUNDS OF


IMIDAZO[1,5-a]PYRIDINYL PHENOLS

(LR_BF2)

70
With the properties of imidazo[1,5-a]pyridines derivatives being studied through the years
in our research group, and the outstanding properties displayed by BF2-containing species
(see Chapter 1.5), we decided to try to enhance the photophysical properties of HLR
compounds upon coordination to a BF2 center,146 similarly to what is done in the
preparation of BODIPY compounds.

4.1 SYNTHESIS AND CHARACTERIZATION

The synthesis of the ligands HLR (R = H, Me, OMe, F, Cl, Br, I, NO2) followed the procedure
described in the previous chapter. 5-Chloro and 5-Iodosalyciladehydes have been
synthesized using a slightly modified procedure reported by Bovosombat,147 by direct and
selective halogenation of salicylaldehyde, using N-chloro or N-iodosuccinimide as
halogenating agent, in acetonitrile in the presence of p-toluensulfonic acid (Figure 4.1).
Then, the boron compounds have been obtained by reacting the ligands with an excess of
boron trifluoride diethyletherate, using triethylamine to deprotonate the phenolic hydroxyl
group and neutralize the hydrogen fluoride formed (Figure 4.2).

Figure 4.1: Synthesis of 5-chloro and 5-iodosalycilaldehyde.

Figure 4.2: Synthesis of boron compounds LR_BF2.

71
All the new species have been characterized by elemental analysis, infrared spectroscopy,
as well as 1H-NMR, 13C-NMR, 19F-NMR and 11B-NMR (Figures 4.3-4.8).
In the IR spectra (Figure 4.3), of importance is the appearing of a series of signals between
1000 and 1100 cm-1 which is the typical range of a BF2 system, thus confirming the obtaining
of a boron difluoride coordination compound similar to those reported in the literature.
Also, the region between 2400 and 2800 cm-1 does not show the typical pattern of signals
traceable to the hydrogen bonding in the ligand (Figure 4.4).

Figure 4.3: Comparison between IR spectra (nujol mull) of LH (black) and HLH_BF2 (red), region between 650
and 1350 cm-1.

Figure 4.4: Comparison between IR spectra (nujol mull) of LH (black) and HLH_BF2 (red), region between
1500 and 3400 cm-1.

In the proton NMR spectra, a set of aromatic protons can be observed between 6.5 and 8.5
ppm, as well as a singlet around 2.5 ppm attributed to the methyl group on the imidazole

72
ring, with all these signals reflecting the expected pattern of the imidazo[1,5-a]pyridinyl
phenols (Figure 4.5). 13C-NMR allowed the complete attribution of the peaks, with the
aromatic carbon atoms ranging from 116 to 132 ppm and the methyl carbon at 9 ppm
(Figure 4.6). The 11B-NMR displayed in every case a triplet signal due to the 11B-19F coupling
at about 0.5-1.0 ppm (JB-F = 15 Hz) (Figure 4.7). The high symmetry of these signals gave us
information about the coordinating environment around the semi-metal center, which, as
later confirmed by structural investigation, is tetrahedral and highly symmetrical, with the
imidazopyridinyl phenol coordinating the boron in a bidentate N-O fashion. Ultimately, the
19F-NMR showed, at -140ppm, a non-binomial quartet, due to the coupling with 11B (JF-B =
15 Hz). Noteworthy, an unresolved shoulder on the left side of the signal appeared, which
has been attributed to the coupling with 10B (Figure 4.8). The other plausible explanation
would have been the presence of two magnetically inequivalent fluorine atoms because of
the two possible orientations of the phenol ring with respect to the imidazopyridine core.
Nevertheless, this option has been ruled out by a low temperature 19F-NMR experiment
which, even if a broadening of the signal occurred, showed no other peaks that would
indicate the presence of some kind of rotamers (Figure 4.9).

Figure 4.5: 1H-NMR of LH_BF2 in CH2Cl2-d2.

73
240 220 200 180 160 140 120 100 80 60 40 20 0 -10
ppm

Figure 4.6: 13C-NMR of LH_BF2 in CH2Cl2-d2.

5.0 4.5 4.0 3.5 3.0 2.5 2.0 1.5 1.0 0.5 0.0 -0.5 -1.0 -1.5 -2.0 -2.5 -3.0 -3.5
ppm

Figure 4.7: 11B-NMR of LH_BF2 in CH2Cl2-d2.

74
-138.5 -139.4 -140.3 -141.2
ppm

Figure 4.8: 19F-NMR of LH_BF2 in CH2Cl2-d2.

-137.5 -138.0 -138.5 -139.0 -139.5 -140.0 -140.5 -141.0 -141.5 -142.0 -142.5
ppm

Figure 4.9: Comparison between 19F-NMR spectra of LH_BF2 in CH2Cl2-d2, recorded at room temperature
(25°C, red) and at -25°C (black).

The compounds LH_BF2 and LOMe_BF2 have been crystallized by slow diffusion of a
dichloromethane solution into diethylether, and the crystals obtained have been used in X-
ray crystal structure analysis. As expected from the spectroscopic data, the boron atom
displayed a tetrahedral geometry, being coordinated by the nitrogen atom of the imidazolic
75
ring and the oxygen atom of the hydroxyl group on the phenolic moiety, with two fluorine
atoms completing the coordination sphere. LH_BF2 showed a nearly planar structure for the
imidazo[1,5-a]pyridine moiety, with the angle formed between the phenolic ring and the
heterocyclic core being 0.000(4)°. On the contrary, LOMe_BF2 was present as two
independent molecules, having the angle between the two rings respectively of 20.3° and
18.2° (Figure 4.10).

Figure 4.10: ORTEP representation of LH_BF2 (left) and LOMe_BF2 (right) at 50% probability level.

Length Angle
Atom1 Atom2 Atom1 Atom2 Atom3
(Å) (°)
B1 F1 1.337 F1 B1 F2 106.8
B1 N1 1.578 N1 B1 F1 109.3
B1 O1 1.440 N1 B1 F2 109.3
B1 F2 1.337 N1 B1 O1 108.5
F1 B1 O1 110.3
F2 B1 O1 110.3
H
Table 4.1: Bond lengths and angles around the boron center for L _BF2 obtained by X-ray diffraction.

Length Angle
Atom1 Atom2 Atom1 Atom2 Atom3
(Å) (°)
B1 F1 1.393 F1 B1 F2 109.7
B1 N1 1.577 N1 B1 F1 108.3
B1 O1 1.446 N1 B1 F2 109.6
B1 F2 1.370 N1 B1 O1 107.4
F1 B1 O1 109.8
F2 B1 O1 110.4

76
Length Angle
Atom1 Atom2 Atom1 Atom2 Atom3
(Å) (°)
B1’ F1’ 1.394 F1’ B1’ F2’ 109.7
B1’ N1’ 1.581 N1’ B1’ F1’ 109.4
B1’ O1’ 1.440 N1’ B1’ F2’ 109.1
B1’ F2’ 1.380 N1’ B1’ O1’ 107.2
F1’ B1’ O1’ 108.7
F2’ B1’ O1’ 112.6
Table 4.2: Bond lengths and angles around the boron center for the two molecules present in the
asymmetric units of LOMe_BF2 obtained by X-ray diffraction.

4.2 OPTICAL PROPERTIES IN SOLUTION

We started the investigation of their photophysical properties. As a starting point, we


evaluated the possible solvent effect on the photophysical properties of these system,
giving their high polarity due to their zwitterionic nature. Taking LH_BF2, LMe_BF2, LCl_BF2
and LNO2_BF2 as representative examples of the whole series, we recorded UV-vis, emission
and excitation spectra, as well as fluorescence quantum yields and lifetimes in different
solvents of different polarity. Considering LH_BF2, regardless of the solvent used, two main
absorption bands always appeared in the range of 234-245 and 345-350 nm (Figure 4.11).
As expected, the absorption spectra measured in acetone, ethyl acetate and N,N-
dimethylformamide did not show the first band because of the absorption due to the
carbonyl group present in these solvents.

0.8

0.6

0.4

0.2

200 250 300 350 400 450 500


Wavelength/nm

Figure 4.11: Normalized UV-vis spectra of LH_BF2 in various solvents. Color code: black, toluene; blue,
tetrahydrofuran; light blue, ethanol; dark green, chloroform; light green, ethyl acetate; orange, acetonitrile;
red, acetone; pink, N,N-dimethylformamide; purple, dichloromethane.

77
The emission spectra showed no relevant differences, with the most significant one being
the more structured band recorded in toluene because of the vibronic structure (Figure
4.12). The range of emission in different solvents is very narrow, from 441 to 450 nm.
Fluorescence intensity (i.e. fluorescence quantum yield) is not greatly affected as well, and
fluorescence lifetimes range from 3.2 to 4.4 ns. These data excluded the influence of the
solvents on the photoluminescence properties of LH_BF2 and, consequently, let us
hypothesize a ligand centered transition without any charge transfer phenomena. This
assignment was later confirmed by DFT calculations (vide infra).

0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0
400 450 500 550 600 650 700 750
Wavelength/nm

Figure 4.12: Normalized emission spectra of LH_BF2 in various solvents. Color code: black, toluene; blue,
tetrahydrofuran; light blue, ethanol; dark green, chloroform; light green, ethyl acetate; orange, acetonitrile;
red, acetone; pink, N,N-dimethylformamide; purple, dichloromethane.

Similar conclusions have been obtained for LMe_BF2 and LCl_BF2 (Tables 4.4-4.5).
On the other hand, LNO2_BF2 did display a great influence of the solvent on the emission
spectra (Table 4.6) This aspect has been subsequently rationalized thanks to the NTOs
topology (see Section 4.3).

λabs λex λem Stokes shift


R=H τ (ns) ΦPL
(nm) (nm) (nm) (eV)
CH2Cl2 347 348 446 0.78 3.5 0.22
Acetone 347 352 445 0.74 3.8 0.21
Acetonitrile 346 352 445 0.74 39 0.15
EtOH 346 347 441 0.76 3.9 0.22
THF 348 348 449 0.80 4.0 0.15
Toluene 351 352 450 0.77 3.8 0.19
AcOEt 347 347 446 0.79 3.7 0.13
Chloroform 348 348 448 0.80 3.2 0.11
DMF 349 353 446 0.73 4.4 0.24
Table 4.3: Photophysical data of LH_BF2 recorded in different solvents (5 x 10-5 M).

78
λabs λex λem Stokes shift
R = Me τ (ns) ΦPL
(nm) (nm) (nm) (eV)
CH2Cl2 352 351 452 0.79 3.5 0.19
Acetone 351 355 448 0.73 3.9 1.19
Acetonitrile 350 349 449 0.79 4.0 1.17
EtOH 349 354 446 0.72 3.9 0.15
THF 353 353 451 0.76 4.1 0.20
Toluene 355 355 454 0.76 3.8 0.17
AcOEt 351 352 449 0.77 3.8 0.16
Chloroform 352 353 451 0.77 3.3 0.15
DMF 353 355 448 0.73 4.4 0.22
Table 4.4: Photophysical data of LMe_BF2 recorded in different solvents (5 x 10-5 M).

λabs λex λem Stokes shift


R = Cl τ (ns) ΦPL
(nm) (nm) (nm) (eV)
CH2Cl2 358 356 442 0.68 4.1 0.22
Acetone 356 359 436 0.61 4.5 0.24
Acetonitrile 356 358 436 0.62 4.5 0.22
EtOH 356 357 436 0.63 4.4 0.27
THF 358 358 439 0.64 4.5 0.28
Toluene 360 362 444 0.63 4.4 0.23
AcOEt 357 357 440 0.66 4.3 0.23
Chloroform 358 357 441 0.66 3.8 0.20
DMF 359 364 433 0.54 5.0 0.35
Table 4.5: Photophysical data of LCl_BF2 recorded in different solvents (5 x 10-5 M).

λabs λex λem Stokes shift


R = NO2
(nm) (nm) (nm) (eV)
CH2Cl2 350 356 597 0.80
Acetone 340 366 623 1.41
Acetonitrile 337 354 657 1.39
EtOH 339 395 447 1.62
THF 350 357 560 0.37
Toluene / / / /
AcOEt 352 355 577 1.34
Chloroform 354 357 589 1.37
DMF 350 367 676 1.54
Table 4.6: Photophysical data of LNO2_BF2 recorded in different solvents (5 x 10-5 M).

Once it was assessed that the solvent effect on the photophysical properties of the vast
majority of these species was negligible (considering that LNO2_BF2 is the less emissive
compound among the series), we proceeded in the investigation using dichloromethane,
which is the best compromise between solvent power and photoluminescence
performances.

79
Absorption spectra of these compounds showed two main absorption bands, the first one
around 240 nm and the second one in the range between 345 and 365 nm, similar to those
of the free ligands. LNO2_BF2 behaved differently, showing one more absorption band at 307
nm (Figure 4.13). All the photophysical data are summarized in Table 4.7.

Figure 4.13: Normalized UV-vis (black), emission (blue) and excitation (red) spectra of compounds LR_BF2
recorded in CH2Cl2 (5 x 10-5 M) solution.

80
λabs λex λem Stokes shift
R τ (ns) ΦPL
(nm) (nm) (nm) (cm-1) (eV)
H 347 348 446 6314 0.78 0.22 3.53
Me 352 351 452 6366 0.79 0.19 3.54
OMe 362 365 463 5799 0.72 0.16 3.68
597 1.1 (95%)
NO2 350 356 5618 0.80 <0.05
(445) 3.0 (5%)
F 357 356 443 5516 0.68 0.22 3.55
Cl 358 356 442 5465 0.68 0.23 4.09
Br 358 357 430 5284 0.66 0.22 4.06
I 359 361 440 4974 0.62 0.09 1.46
Table 4.7: Photophysical data of compounds LR_BF2 recorded in CH2Cl2 (5 x 10-5 M) solution.

The emission spectra are all very similar, with the emission maxima differing of just 23 nm,
from 440 to 463 nm, being positioned in the deep blue region, with CIE 1931 plot (x,y) color
coordinates very close to those considered as “standard blue”148 (Table 4.8). The only the
only outlier was, as expected, LNO2_BF2 which displayed two maxima of emission, the first
one in the previous mentioned range, and the absolute maximum in the red-orange region
at 597 nm.

Figure 4.14: Normalized emission spectra of LR_BF2 in compounds in CH2Cl2 (5 x 10-5 M) solution. Color
code: black, LH_BF2; blue, LMe_BF2; light blue, LOMe_BF2; dark green, LF_BF2; light green, LCl_BF2; orange,
LBr_BF2; red, LI_BF2; pink, LNO2_BF2.

81
y 520
CIE 1931
530
0.8 540
510
550
0.7
560

0.6 570
500
580
0.5
590

0.4 NO2
600
610
620
0.3490 680

OMe
0.2 Me
H
F
Cl
I
Br
480
0.1
470
0.0 460
x
0.0 0.1 420 0.2 0.3 0.4 0.5 0.6 0.7

Figure 4.15: CIE1931 chromaticity plot of the emission in solution (CH2Cl2, 5 x 10-5 M) of compounds LR_BF2.

R x y

H 0.16145 0.15201
Me 0.16336 0.16613
OMe 0.17252 0.21003
NO2 0.37239 0.36685
F 0.15994 0.13593
Cl 0.15860 0.12244
Br 0.15842 0.12118
I 0.15874 0.12198
Standard
0.16 0.10
blue
Table 4.8: Color coordinates of the emission in solution (CH2Cl2, 5 x 10-5 M) of compounds LR_BF2 in the CIE
1931 chromaticity plot.

The excitation spectra at both this emission maxima were recorded for this species; we
observed their complete equality (Figure 4.16) which let us exclude the possibility that this
double emission was due to the presence of two different ground states. The motivation
behind this different behavior has been ascribed to two different transitions, being
respectively a 1π- π* at higher wavelength and a 1CT at shorter wavelength, the one at
higher frequencies and the other to lower frequencies respectively, with the one at 445 nm
being the most dominant, as also hypothesized by its dependence on the solvent polarity
and later confirmed by NTO analysis (see Section 4.3).

82
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0
250 300 350 400 450
Wavelength/nm

Figure 4.16: Normalized excitation spectra in solution (CH2Cl2, 5 x 10-5 M) of LNO2_BF2, recorded at 597 nm
(black) and 452 nm (blue).

For the rest of the series, the excitation spectra reproduced the corresponding absorption
traces.
Generally speaking, it can be observed how the emission maxima of the compounds in
these series depend on the substituent R on the phenolic ring, being the only discriminant
factor. In particular, the more is the electron donating character of the substituent, the
higher is the wavelength of emission.
Large Stokes shifts, from 0.62 to 0.80 eV, have been observed for these systems, as well as
good fluorescence quantum yields from 16% to 23%, with the only exception of LNO2_BF2
and LI_BF2, respectively experiencing the quenching effect of the nitro group and of the
heavy-atom effect. Fluorescence lifetimes are in the range of 3.5-4.1 ns, displaying a mono-
exponential decay curve, with the exceptions again of LNO2_BF2 which had a bi-exponential
decay characterized by τ1 = 1.1 ns and τ2 = 3.0 ns, coherently to what expected for a species
showing two emission maxima; LI_BF2 also represents an exception, with a lifetime of just
1.5 ns.
Because of their very promising fluorescence and solubility properties that let us think
about their possible application in optoelectronic devices, we also performed
photostability measurements. LH_BF2 was chosen as representative example of the series:
in dichloromethane solution, the compound has been subjected to a multiple emission scan
analysis, irradiating the sample at its absorption maximum for a total of two hours, after

83
which a total loss of emission intensity of just 0.84% has been observed, indicating a very
good photostability of these systems (Figure 4.17).

LH_BF2 photostability in CH2Cl2 solution


1,5
Emission Intensity (a.u)

1,3

1,1

0,9

0,7

0,5
0 20 40 60 80 100 120
Time (min)

Figure 4.17: Intensity of emission of LH_BF2 vs. time in dichloromethane solution (5 x 10-5 M).

Finally, it has been noted that compound LH_BF2 displayed different emission maxima
depending on the concentration of its solutions. In standard dilute solutions (5 x 10-5 M),
the abovementioned blue emission was observed, while in more concentrated solutions (5
x 10-3 M) an orange emission at 546 nm was observed instead, accompanied by the
previously observed blue fluorescence (Figure 4.18).

0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0
400 450 500 550 600 650 700
Wavelength/nm

Figure 4.18: Emission spectra of LH_BF2 recorded in dichloromethane solution (3 x 10-5 M), pink form (black)
and yellow form (blue).

84
Solvent polarity index vs Time of colour change
(min)
R² = 0,9168
70
60
50
40
Time (min)
30
20
10
0
-10 3,5 4 4,5 5 5,5 6 6,5 7
-20
Solvent Polarity Index

Figure 4.19: Time needed (min) for the LH_BF2 solution to change from orange-pink to yellow vs. solvent
polarity.

Also, the color of the more concentrated solutions changed over time, with a rate
depending on the solvent polarity149 (Figure 4.19): dissolving the pink-orange solid
obtained after synthesis, an orange solution could be obtained, which turned to yellow
over time. By evaporation of the solvent, a yellow solid was recovered, instead of the
starting pink-orange one.
To understand the differences between the two solids, several analyses have been
performed. Firstly, an attempt of crystallize the two forms has been done, but
unfortunately crystals leading to the same diffraction patterns were obtained.
Secondly, a multiple 19F-NMR spectra analysis (very sensitive to structural changes) has
been performed, revealing no structural differences occurring over time during color
changing of the solution (Figure 4.20).

Figure 4.20: 19F-NMR spectra of LH_BF2 (pink form) in CH2Cl2 over time.

85
A DSC-TGA analysis has been then performed on the pink solid, showing a mass loss of ca.
4% (Figure 4.21), letting us hypothesize the presence of water in the system. This is
coherent with the time needed for the color change in solution: the higher the solvent
polarity, the less time was needed to lose the water and hence to change the color. Also,
this could explain why in dilute solutions, only the “yellow” form is present. Also, when the
pink solid has been left at 100 °C over night, it turned to the yellow form, thus confirming
this hypothesis. Any attempt to reconvert the yellow form into the pink one has been
unsuccessful.

Figure 4.21: DSC/TGA analysis of LH_BF2.

An X-ray powder diffraction analysis was performed on the two solids, finally confirming
the fact that the yellow form is indeed the LH_BF2 compound, while the pink form is a
partially hydrated species of the same sample, being a mixture of hydrated and anhydrous
form, hence making the quantification of the exact water content not quantifiable (Figure
4.22).

Figure 4.22: Comparison between XRPD pattern of LH_BF2: hydrate form (green) vs anhydrous form (blue).

86
4.3 DFT CALCULATIONS

Starting from the LH_BF2 structure obtained experimentally, the ground state optimized
geometries for the whole series have been calculated at the DFT/PBE-D3 level of theory.
Comparison between experimental and calculated bond distances and angles for LH_BF2
and LOMe_BF2 proved the good agreement between the two and hence TD-DFT calculations
have been performed to better understand the nature of the transitions responsible for
the luminescence properties.
The calculated absorption spectra for LH_BF2 and LNO2_BF2 are in good agreement with the
experimental spectra (Figure 4.23).

Figure 4.23: Experimental (black line) and calculated (dashed blue line) spectra UV-vis spectra of LH_BF2
(left) and LNO2_BF2 (right).

It has been observed that almost for every compound of this series, the higher-wavelength
absorption corresponded to two electronic transitions, mainly being constituted by a
HOMO-LUMO and HOMO-LUMO+1 character respectively. In the case of LNO2_BF2, three
different transitions were responsible for the absorption process, with mainly a HOMO-
LUMO, HOMO-LUMO+1 and HOMO-LUMO+2 character.
From this, NTOs have been calculated: for all the compounds except the nitro derivative,
the occupied NTOs are spread over the whole molecule, while the virtual NTOs are localized
on the imidazopyridine moiety for NTO-1 or again over the entire molecule in NTO-2.
LNO2_BF2 occupied NTOs are also spread over the whole molecule, as well as two of the

87
virtual NTOs, while the one responsible for the first transition is localized over the phenolic
ring (Figure 4.24).

LR_BF2 NTOs
H
L _BF2 NTO1 HOMO → LUMO 81.3%
HOMO-1 → LUMO 7.5%
NTO2 HOMO → LUMO+1 5.3%
HOMO-2 → LUMO+1 1.3%
HOMO → LUMO+1 83.5%
HOMO → LUMO 8.1%
HOMO-1 → LUMO 2.1%
HOMO-1 → LUMO+1 2.0%
HOMO → LUMO+2 1.3%
LNO2_BF2 NTO1 HOMO → LUMO+1 98.0%
NTO2 HOMO-1 → LUMO 1.5%
HOMO → LUMO 93.2%
NTO3 HOMO-1 → LUMO 3.5%
HOMO-3 → LUMO+2 1.1%
HOMO → LUMO+2 89.7%
Table 4.9: Contributions of single orbital transitions for LH_BF2 and LNOw_BF2.

Figure 4.24: NTOs associated to the main electronic transition (>95%) S0 → S1 in LH_BF2 and LNO2_BF2.

88
HOMO and LUMO orbitals reproduced the topology displayed by NTOs, as they are the
orbitals contributing the most the electronic transitions. HOMO orbitals are in fact spread
over the whole molecules, while LUMO orbitals are localized on the imidazopyridine
portion. Regarding LNO2_BF2, the calculations showed the HOMO orbital to be localized on
the entire molecule, while LUMO is mainly centered on the phenolic ring (Figure 4.25).

Figure 4.25: HOMO and LUMO orbitals calculated for LMe_BF2 and LNO2_BF2.

In any case, as also observed in the previous Chapter 3.3, the coordination center is not
involved in the transition processes, thus only having a structural role.
Considering the emission properties of these species in solution, it can be said that for the
majority of these systems, the transition with the lower charge transfer character is the
dominating one; on the contrary, the dominant transition for the nitro derivative, is the
one with the highest charge transfer character.
Electron Difference Density Maps (EDDM) have been calculated as well, once again
confirming the conclusions derived from the NTOs analyses: the electron density moves
from the phenolic ring to the heterocyclic moiety, except for LNO2_BF2 in which the charge
transfer character is more distinct, and the density moves in the opposite way than for the
rest of the series (Figure 4.26).

89
Figure 4.26: EDDMs associated to the lowest energy singlet electronic transition calculated (black →
decreasing electron density, blue → increasing electron density).

As it has been calculated that the main contribution to the transitions came from HOMO,
LUMO and LUMO+1 orbitals, a dependence of their energies from the electronic properties
of the ligands has been investigated. Electron donor groups enhanced the HOMO energy,
while electron withdrawing group led to the opposite direction, with LUMO energies
following the same trend. Thanks to this similarity between HOMO and LUMO energies
behavior, the HOMO-LUMO energy gap remained almost constant throughout the whole
series, ranging from 2.087 to 2.472 eV.
Using σp Hammett constants of R, a linear correlation between the orbital energies and the
constants themselves could be found (Figure 4.27), describing the influence of the
electronic features of the imidazopyridine ligands on the emissive properties, which are in
turn related to the HOMO and LUMO energies. Frontier orbitals energies decreased with
increasing the σp, or increasing the electron withdrawing character of the substituent R.
Moreover, the energy gap remained almost constant, possibly explaining the reason for
such a small range of emission wavelengths in this series of compounds (Table 4.10).
Therefore, it can be concluded that this new class of boron difluoride compounds can pave
the way to a whole series of blue emissive species, which can be freely functionalized
without drastically changing their emissive properties.

90
Figure 4.27: Correlation between the Hammett constants of the R substituents vs HOMO and LUMO
energies.

EHOMO ELUMO ΔH-L σP


R
(eV) (eV) (eV)
H -5.339 -2.976 2.363 0
Me -5.228 -2.940 2.288 -0.17
OMe -5.048 -2.961 2.087 -0.27
NO2 -5.860 -3.541 2.319 0.78
F -5.402 -3.082 2.320 0.06
Cl -5.414 -3.098 2.316 0.18
Br -5.445 -3.114 2.330 0.23
I -5.434 -3.105 2.472 0.23
Table 4.10: Calculated HOMO and LUMO energies of LR_BF2 compounds.

For this species, except for the case of LNO2_BF2 which displayed a different behavior, the
following empirical equation could be formulated, relating σp and the emission maxima:

1240
𝜆 = 𝑛𝑚
0.1318𝜎 + 2.7650

To conclude, a new series of BF2-containing compounds have been synthesized and


characterized. Moreover, their fluorescence properties in solution have been studied: blue
emission with good quantum yields has been observed for each compound of the series,
except for LNO2_BF2, which, as expected, behaved differently both in terms of fluorescence
and for what concerned its orbitals morphology. Nevertheless, considering the very narrow
range of emission of these compounds, they represent a very interesting class of blue
emissive materials that can also be further functionalized in order to be used for various
applications, without losing their characteristic emission.

91
Chapter 5

BORON DIFLUORIDE COMPOUNDS OF


TETRAHYDROIMIDAZO[1,5-a]PYRIDINYL PHENOLS

(hydrLR_BF2)

92
As discussed in Chapter 2.2.2, in the past years a series of tetrahydroimidazo[1,5-
a]pyridinyl phenols has been synthesized, characterized and the photophysical properties
of this new class of compounds have been investigated.142 Unexpectedly, the luminescence
performances of this species (i.e. fluorescence quantum yield) were better than the
corresponding non-hydrogenated derivatives. For this reason, analogously to what has
been discussed in the previous chapter, we decided to prepare and study boron difluoride
compounds using tetrahydroimidazopyridines as ligands, in order to further enhance their
luminescence properties.150

5.1 SYNTHESIS AND CHARACTERIZATION

The synthesis of the imidazo[1,5-a]pyridinyl phenols have been conducted using the same
reaction descried in Section 3.1. Then, a further step has been done, hydrogenating the
pyridinic ring of substituted imidazo[1,5-a]pyridinyl phenols (R = H, Me, OMe, F, OH) using
molecular hydrogen in methanol, with Pd/C as catalyst (see Section 2.2.2). Finally, the
complexation reaction with boron trifluoride diethyletherate has been done using
dichloromethane as solvent in the presence of triethylamine (see Section 4.1). In this case,
a simple purification process was needed, by filtering the product solution over a silica pad.
Unfortunately, the hydrogenation of HLBr resulted in the hydrogenolysis of the bromine
atom, obtaining the hydrobromide species, which has not been used in the following
investigations.

Figure 5.1: Synthetic pathway for compounds hydrLR_BF2.

All the compounds have been characterized using elemental analysis, infrared
spectroscopy, 1H-NMR, 13C-NMR, 11B-NMR and 19F-NMR.

93
The infrared spectra showed a series of broad signals in the region between 1000 and 1200
cm-1, indicative of the presence of the BF2 group (Figure 5.2).

Figure 5.2: IR spectrum of hydrLH_BF2 (nujol mull). The red circle highlights the region attributed to the BF2.

In the 1H-NMR spectra it was possible to observe a first set of signals in the aromatic region
of 7-8 ppm, while the methylene protons were found at 1.90, 2.15, 2.80 and 4.35 ppm, with
also the singlet relative to the methyl group at ca. 2.3 ppm (Figure 5.3). The 13C-NMR

showed the corresponding signals in the aliphatic carbon region (ca. 45 to 15 ppm), while
the aromatic carbon could be found in the range from 115 to 135 ppm (Figure 5.4).
The 11B-NMR spectra displayed a highly symmetrical triplet signal centered at about 0.5-
1.0 ppm, with a coupling constant JB-F of about 15 Hz (Figure 5.5). The shape of the signal
is an indication of the magnetic equivalency of the two fluorine atoms, coupled with the
boron center, as expected for the tetrahedral geometry later confirmed by structural
analysis.
Finally, 19F-NMR showed a non-binomial quartet centered at about -140 ppm, with the
same JF-B coupling constant of ca. 15 Hz. The coupling with 10B is also detectable in the
unresolved shoulder appearing on left side of the quartet (Figure 5.6).

94
Figure 5.3: 1H-NMR of hydrLH_BF2 in CH2Cl2-d2.
131.93

123.93

118.92

46.65

22.80
20.44
18.76

8.90

Figure 5.4: 13C-NMR of hydrLH_BF2 in CH2Cl2-d2.

95
0.77

Figure 5.5: 11B-NMR of hydrLH_BF2 in CH2Cl2-d2.


-139.03
-139.07
-139.11
-139.15

Figure 5.6: 19F-NMR of hydrLH_BF2 in CH2Cl2-d2.

The structure of hydrLOMe_BF2 has been resolved in a X-ray single crystal structure analysis,
showing the expected tetrahedral environment around the boron center, constituted of
the two fluorine atoms and of the ligand binding in a N-O fashion. The heterocyclic moiety

96
is no more planar, with the piperidinic ring assuming a half-chair conformation. Moreover,
the phenolic ring is twisted with respect to the imidazole plane of 9.11° (Figure 5.7).
Even with this deviation from planarity, hydrLOMe_BF2 form π- π head-to-tail stacks along
the a axis, with a distance of 3.88 Å.

Figure 5.7: ORTEP representation of hydrLOMe_BF2 (left) at 50% probability level and π- π interaction (right).
See Figure 8.63 for atom labelling.

Length Angle
Atom1 Atom2 Atom1 Atom2 Atom3
(Å) (°)
B1 F1 1.384(5) F1 B1 F2 109.2(3)
B1 F2 1.390(5) F1 B1 N1 110.3(3)
B1 N1 1.561(5) F1 B1 O1 108.8(3)
B1 O1 1.429(5) F2 B1 N1 107.8(3)
F2 B1 O1 111.8(3)
N1 B1 O1 108.9(3)
Table 5.1: Bond lengths and angles around the boron center for hydrLOMe_BF2 obtained by X-ray diffraction.

5.2 OPTICAL PROPERTIES

5.2.1 OPTICAL PROPERTIES IN SOLUTION


The investigation of the photophysical properties of these species started from the
evaluation of a possible solvent effect. Using hydrLH_BF2 as example of the whole series,
we recorded absorption, emission and excitation spectra, as well as fluorescence quantum
yield and lifetime in various solvents with different polarity.
Regardless of the solvent used, the UV-vis spectra displayed no significant difference, with
two main absorption bands, one at 270-280 nm and the second one at about 320 nm

97
(Figure 5.8). The only differences were detectable in acetone solution, where the first
absorption band did not appear because of the strong absorption of the solvent in that
range, and in cyclohexane, in which the product was poorly soluble and hence the second
absorption band is present only as a low intensity broad band. The excitation spectra
matched with the UV-vis spectra.
Emission spectra appeared very similar in shape, but with a slight bathochromic shift of the
maxima of emission with decreasing polarity of the solvent, contrary to what is usually
observed for similar compounds (Figure 5.9).

Figure 5.8: Normalized UV-vis spectra of hydrLH_BF2 recorded in different solvents (5 x 10-5 M). Color code:
blue, dichloromethane; red, acetone; black, acetonitrile; green, ethanol; purple, tetrahydrofuran; light blue,
toluene; orange, cyclohexane.

Figure 5.9: Normalized emission spectra of hydrLH_BF2 recorded in different solvents (5 x 10-5 M). Color
code: blue, dichloromethane; red, acetone; black, acetonitrile; green, ethanol; purple, tetrahydrofuran;
light blue, toluene; orange, cyclohexane.

98
λabs λex λem Stokes shift
R=H τ (ns) ΦPL
(nm) (nm) (nm) (eV)
CH2Cl2 319 319 357 0.41 1.3 0.54
Acetone 331 333 356 0.24 1.3 0.27
Acetonitrile 316 316 354 0.42 1.5 0.41
EtOH 317 318 356 0.42 1.4 0.34
THF 320 320 356 0.41 1.4 0.31
Toluene 322 323 360 0.39 1.3 0.46
CyH 323 324 362 0.40 1.2 0.17
Table 5.2: Photophysical data of hydrLH_BF2 recorded in different solvents (5 x 10-5 M).

In any case, no dramatic quenching of the fluorescence has been detected and mono-
exponential decay lifetimes between 1.2 and 1.5 ns have always been recorded. Even
considering the abovementioned slight bathochromic shift, it could be concluded that the
solvent only had a very minor effect on the fluorescence performances of these systems.
For this reason, we decided to use methylene chloride as solvent for performing the
following analysis, also in accordance with what has been done for the previous classes of
compounds (see Chapter 3.2 and 4.2).
Regarding all the other compounds, the absorption spectra are very similar to the one
described for hydrLH_BF2, being composed of two absorption bands, the first one at 240-
290 nm and the second at 320-340 nm, with the excitation spectra again largely
reproducing the traces of the UV-vis spectra (Figure 5.10).
The emission maxima instead fell in the near UV/blue region, in the range from 357 to 390
nm and were influenced by the nature of the R substituents on the phenolic ring of the
ligands: to higher electron donating properties of R, corresponded high emission maxima
(Figure 5.11).

Figure 5.10: Normalized UV-vis spectra of hydrLR_BF2 recorded in CH2Cl2 (5 x 10-5 M). Color code: red,
hydrLH_BF2; orange, hydrLH_BF2; green, hydrLOMe_BF2; blue, hydrLF_BF2; black, hydrLOH_BF2.

99
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0
350 400 450 500 550
Wavelength/nm

Figure 5.11: Normalized emission spectra of hydrLR_BF2 recorded in CH2Cl2 (5 x 10-5 M). Color code: red,
hydrLH_BF2; orange, hydrLH_BF2; green, hydrLOMe_BF2; blue, hydrLF_BF2; black, hydrLOH_BF2.

Figure 5.12: Normalized UV-vis (black), emission (blue) and excitation (red) spectra of compounds
hydrLR_BF2 recorded in CH2Cl2 (5 x 10-5 M) solution.

100
All the compounds displayed mono-exponential lifetime decays between 1.1 and 1.4 ns.
Noteworthy is the fact that fluorescence quantum yields recorded for hydrLOMe_BF2, was
significantly higher than that recorded for the free ligand (Table 5.3, see Table 2.4 for
comparison).
We also performed a photostability measurement on hydrLH_BF2, resulting in a loss of
emission intensity of just 3.49% after 100 minutes of irradiation at its absorption maximum.

λabs λex λem Stokes shift


R τ (ns) ΦPL
(nm) (nm) (nm) (cm-1) (eV)
H 319 319 357 3307 0.41 1.3 0.54
Me 326 326 367 3388 0.42 1.3 0.47
OMe 342 342 390 3629 0.45 1.4 0.49
F 329 329 369 3307 0.41 1.1 0.30
OH 342 341 388 3549 0.44 1.4 0.31
Table 5.3: Photophysical data of compounds hydrLR_BF2 recorded in CH2Cl2 solution (5 x 10-5 M).

5.2.2 OPTICAL PROPERTIES IN SOLID STATE


Contrary to what observed for the non-hydrogenated boron compounds, these new
materials displayed a bright blue fluorescence even in the solid state, with excitation and
emission maxima being red-shifted with respect to those in solution.
Fluorescence quantum yields are higher than those in solution for hydrLH_BF2, hydrLMe_BF2
and hydrLF_BF2, while hydrLOMe_BF2 and hydrLOH_BF2 saw a lowering in their fluorescence
intensity. Lifetimes are longer though, in the range between 6-7 ns, being multiexponential
in the case of the stronger electron donating R substituent (i.e. R = OMe, OH) (Table 5.4).
The bathochromic shift of the emission maxima in the solid state caused the λmax to fall in
the deep blue region, with CIE 1931 (x,y) color coordinates very close to those reported for
standard blue (Table 5.5).145
A photostability analysis has been conducted also in the solid state, revealing a higher
degree of degradation when compared to the results obtained in solution, with a loss in
emission intensity of 12.76% after 135 minutes.

101
λex λem Stokes shift
R τ (ns) ΦPL
(nm) (nm) (cm-1) (eV)
H 343 430 5888 0.73
6.1 0.68
Me 353 435 5404 0.67
6.9 0.63
1.7 (43%)
OMe 368 411 2823 0.35 0.35
5.4 (57%)
F 354 406 3629 0.45 5.6 0.47
0.5 (44%)
OH 364 388 1694 0.21 2.4 (25%) 0.15
7.5 (31%)
Table 5.4: Photophysical data of compounds hydrLR_BF2 recorded in solid state.

R x y

H 0.16398 0.06242
Me 0.15917 0.08492
OMe 0.15950 0.06810
F 0.16398 0.06242
OH 0.18361 0.15142
Standard
0.16 0.10
blue
Table 5.5: Color coordinates of the emission in the solid state of compounds hydrLR_BF2 in the CIE 1931
chromaticity plot.

5.2.3 OPTICAL PROPERTIES IN POLYMERIC FILM


Considering this deep blue emission with excellent quantum yields suggesting the possible
use of these materials in the fabrication of blue emissive optoelectronic devices, we further
investigated their photoluminescence properties in polymeric films, using poly(methyl
methacrylate) (PMMA) as host material.
Excitation and emission spectra profiles are similar to those recorded in solution and
lifetime decays are all biexponential, with τ1 = 1.0-1.3 ns and τ2 = 2.5-4.6 ns (Table 5.6). This
is probably due to some kind of interaction with the matrix, leading to different relaxation
pathways, as also indirectly confirmed by the lower fluorescence quantum yields compared
to those recorded in solution or in solid state. Self-absorption processes, not avoidable at
5% weight concentration, could also cause this behavior.

102
0.9
8.0 0.8
0.7
6.0 0.6
0.5
4.0 0.4
0.3
2.0 0.2
0.1
0.0 0.0
240 260 280 300 320 340 360 380 400 420 350 400 450 500 550 600 650
Wavelength/nm Wavelength/nm

Figure 5.13: Comparison between the normalized excitation (left) and emission (right) spectra of
compounds hydrLH_BF2 recorded in CH2Cl2 solution (5 x 10-5 M) (blue), in solid state (red) and in PMMA
films (5% w/w) (black).

λex λem Stokes shift


R τ (ns) ΦPL
(nm) (nm) (cm-1) (eV)
1.2 (85%)
H 318 358 3549 0.44 0.25
4.4 (15%)
1.0 (77%)
Me 325 369 3710 0.46 0.16
2.8 (23%)
1.3 (64%)
OMe 340 391 3871 0.48 0.16
2.9 (36%)
1.3 (83%)
F 328 371 3549 0.44 0.19
4.6 (17%)
1.2 (54%)
OH 343 394 3791 0.47 0.13
2.5 (46%)
Table 5.6: Photophysical data of compounds hydrLR_BF2 recorded in PMMA films (5% w/w).

5.3 DFT CALCULATIONS

The data obtained from the structural X-ray experiment on hydrLOMe_BF2 have been used
as a starting point for the optimization of the geometries of the compounds of the whole
series at a DFT/PBE-D3 level of theory. Moreover, to better understand the nature of the
electronic transitions responsible for the luminescence properties, TD-DFT calculations
have been performed. The calculated absorption spectra reproduced the experimental
ones, with the absorption at higher wavelength being composed of two main transitions,
HOMO-LUMO and HOMO-1-LUMO (Figure 5.14).

103
Figure 5.14: Experimental (blue line) and calculated (dashed red line) spectra UV-vis spectra of
hydrLOMe_BF2 (CH2Cl2 solution, 5 x 10-5 M).

HOMO and LUMO topologies are similar for every compound of this series, both being
delocalized over the aromatic portion of the molecules, excluding the piperdinic aliphatic
ring. The HOMO-1 are instead localized only on the imidazolic ring (Figure 5.15). In
accordance with results obtained for the previous series of boron difluoride compounds
presented in Chapter 4, no contribution from the BF2 fragment is observed. For this reason,
it can be concluded again that the boron center enhanced the photophysical properties by
giving more rigidity to the system, thus preventing rotational non-radiative decay paths.

Figure 5.15: Visualization of the orbitals involved in the transitions for hydrLOMe_BF2

104
Finally, the geometries of both ground and first excited states have been calculated in
dichloromethane solution for hydrLOMe_BF2: it has been possible to observe a distortion of
the excited state when compared to the ground state, in particular the angle between the
plane defined by the phenolic ring and the one defined by the heterocyclic moiety are
twisted of 15.5° in the ground state, with the same angle measuring 38.47° in the excited
state (Figure 5.16, see also Figure 8.63).

Figure 5.16: Visualization of the angle formed by the planes defined by the phenolic ring and the imidazolic
ring in hydrLOMe_BF2.

In conclusion, the preparation of BF2-containing compounds using tetrahydroimidazo[1,5-


a]pyridinyl phenols provided an enhancement of the luminescence performances (i.e.
fluorescence quantum yield) with respect to the corresponding non-hydrogenated
derivatives. The excellent photophysical properties of this class of compounds recorded in
solution are maintained also in the solid state, and a deep blue emission has also been
observed in polymeric films, suggesting a possible application of these compounds as
emissive materials in optoelectronic devices.

Figure 5.17 hydrLH_BF2 irradiated at 265 nm, when dispersed in PMMA (5% w/w).

105
Chapter 6

BUILDING OLEDS

106
Among various applications for luminescent compounds, one of the most studied topics is
their use as emissive materials in the fabrication of optoelectronic devices, such as Organic
Light Emitting Diodes (OLEDs). In particular, great interest is devoted to blue emissive
species, as the devices currently obtainable with blue emission are not as efficient as the
corresponding green or red OLED. Nowadays, in regular mobile phone or screens in
general, almost 50% of the whole pixel area is covered by blue pixels, in order to
compensate for their lack of efficiency.
With the OLED technology advancing more and more over the years, for example with new
devices exploiting this technology to achieve higher image resolution and brightness, the
research for new materials to be used as blue emissive layer is more important than ever.
With this been said, the previously described classes of imidazo[1,5-a]pyridine compounds
have been evaluated and screened, in order to find, among the three, the best series of
blue emissive materials among the three to be tested in the fabrication of new OLEDs.*
This part of work was conducted at the University of Cologne.

6.1 PRELIMINARY SCREENING


At first, the methodology to be used for the devices fabrication was chosen: as the spin
coating process is cheaper, easier, less time consuming and, most important, more scalable
than the vapor deposition process, we opted to process the emissive species in solution.
Furthermore, the thermal stability of these compounds should have been studied, while
their solubility properties were already known, as, for some of them (i.e. hydrLR_BF2), the
luminescence properties in polymeric films.
Therefore, the first variable to be considered has been the solvent of choice.
Based on previous experiences, toluene and chlorobenzene proved to be the most suitable
solvents in terms of film formation properties.
We then proceeded testing the solubility of each of the previous described class of
compounds: the zinc complexes actually displayed a very poor solubility, not only in these
two solvents, but also in every other tested, with dimethylsulfoxide being the only one

*
OLED manual fabrication is often affected by important batch-to-batch variations. For this reason, slightly
different values for the same device architecture can be found in the following tables.

107
capable of dissolving them appropriately. However, dimethylsulfoxide has a too high
boiling point to be used in this kind of fabrication method.
Boron difluoride compounds instead are more soluble in organic solvents, thus
representing the best candidates. Their solubility in both toluene and chlorobenzene
though, was not sufficient as well, and the only solvent with which we managed to obtain
solutions of these species concentrated enough (25 mg/mL) was N,N-dimethylformamide.
Unfortunately, despite they represent the best candidates in terms of photoluminescence
properties, hydrLR_BF2 derivatives were soluble enough only in hot DMF, which results in
very poor layer quality after the spin coating process, with defects visible even to the naked
eye. For these reasons, the non-hydrogenated boron difluoride compounds have been the
most tested during the following experiments, in particular LOMe_BF2 was chosen as the
best candidate because of its calculated HOMO energy being the highest among the series
and therefore being the closest to those of the hole transporting materials. Nonetheless,
some hydrogenated derivatives have been tested as well.

Class of compounds Solubility Quantum yields (in solution)

[Zn(LR)2] Only DMSO Up to 0.33

LR_BF2 Good, best in DMF Up to 0.23

hydrLR_BF2 Good, best in hot DMF Up to 0.54

Table 6.1: Comparison between solubility and luminescence properties of the previously described classes
of compounds.

6.2 BASIC ARCHITECTURE OF DEVICES


The first trial of devices was built using a standard architecture (Figure 6.1).
We opted to use glass covered with Indium Tin Oxide (ITO) as substrate and anode, on top
of which a water solution of PEDOT:PSS was spin coated at 4000 rpm for 30 seconds. The
substrates have been then heated at 150 °C for 5 minutes in order to remove any trace of
water, finally obtaining a layer of around 35 nm. This is needed to make the following layer

108
stick on the non-homogeneous surface of ITO and, therefore, to obtain better devices in
terms of layer thickness and performances.

Figure 6.1: Schematization of the basic architecture of the described devices.

Two different Hole Transporting Layers (HTLs) have been then deposited on top of the
PEDOT layer, the first one being made of N4,N4'-bis(4-((6-((3-ethyloxetan-3-
yl)methoxy)hexyl)oxy)phenyl)-N4,N4'-bis(4-methoxyphenyl)-[1,1'-biphenyl]-4,4'-diamine
(QUPD) and the second one made of N4,N4'-bis(4-(6-((3-ethyloxetan-3-
yl)methoxy)hexyl)phenyl)-N4,N4'-diphenyl-[1,1'-biphenyl]-4,4'-diamine (OTPD), whose
HOMO energies formed an energetically favorable step to make the hole injection in the
emissive layer efficient (Figure 6.3), considering the calculated frontier orbitals energy of
the emissive material. The two hole transporting materials have been deposited
subsequently using a 5 mg/mL toluene solutions and spin coated for 20 seconds at 1500
rpm. A photoacid have been added to these solutions in 2% molar concentration (OPPI, (4-
((2-hydroxytetradecyl)oxy)phenyl)(phenyl)iodonium hexafluoroantimonate(V)): treatment
of the spin coated solution with irradiation at 366 nm for 15 seconds followed by a curing
step of 1.5 minutes at 110 °C led to the polymerization of the hole transporting materials
making them insoluble in organic solvents. This is critically important as on top of those
layers, other materials had to been deposited from organic solutions, therefore a method
to prevent them to be washed away was needed.

109
CH3 H3 C
O O
O O
N N

N N

S
O O O
O

O
SO3 O O
O O
O

OH
F
O F F
Sb
F F
F
I

Figure 6.2: Top: from left to right, structures of PEDOT:PSS, QUPD and OTPD
Bottom: structure of OPPI.

Figure 6.3: Schematization of hole transporting mechanism from the anode (ITO) to the emissive layer with
(left) and without (right) hole transporting layers.

The emissive material has been deposited spin coating 190 μL of a DMF solution for 20
seconds at 1000 rpm, with an annealing process at 150 °C for 1.5 minutes. Also, we tested
the possibility to disperse the emissive material in a host polymeric matrix, using poly(vinyl
carbazole) (PVK) as the host matrix, starting with a 30% w/w concentration of LOMe_BF2.
Finally, the cathode has been deposited under vacuum: firstly, a thin layer of 2 nm of CsF
has been evaporated, then a 70 nm layer of aluminum has been deposited. All the
deposition conditions of the various layers are summarized in Table 6.2.

110
Layer Method Condition Curing Thickness

Glass/ITO / / / /
200 μL, 30 s, 150 °C,
PEDOT:PSS Spin coating 35 nm
4000 rpm 5 min
190 μL,
15 s UV (366 nm),
QUPD Spin coating 20 s, 25-30 nm
110 °C, 1.5 min
1500 rpm
190 μL,
15 s UV (366 nm),
OTPD Spin coating 20 s, 25-30 nm
110 °C, 1.5 min
1500 rpm
190 μL,
Emissive 150 °C,
Spin coating 20 s, 70-80 nm
material 10 min
1000 rpm
2 nm CsF,
Al/CsF Vapour deposition / /
70-100 nm Al
Table 6.2: Process conditions of the various layers.

The mechanism with which the HTM are polymerized is quite complex and worth of a
deeper description:151 OPPI generates electrophilic species under UV-irradiation which
then start the cross-linking process. The initiation step starts with the sensitization of the
photoacid, thanks to the hole transporting material itself, which has an absorption
wavelength longer than that of the photoacid. Upon photoexcitation, a photoinduced
electron transfer from the HTM to the photoacid occurs, forming a radical cation of the
HTM and an unstable photoacid radical, the latter further decomposing to an aryl iodide
and to a phenyl radical. At this point, the HTM radical cation and the phenyl radical may
recombine with each other and, in this case, a rearomatization of the system leads to the
generation of a free proton, initiating the polymerization reaction driven by the oxetane
ring opening (Figure 6.3).

111
Figure 6.4: Polymerization mechanism of QUPD and OTPD.

Once completed the fabrication process, we measured the electroluminescence properties


of the devices (Figure 6.5).

Brightness Efficiency Turn-on voltage


Emissive layer 2
(Cd /m ) (Cd/A) (V)

OMe
L _BF2 with no OTPD 2.1 0.054 12

OMe
L _BF2 7.4 0.0035 6

OMe
L _BF2 20% w/w in PVK 22.1 0.0039 8.5

H
L _BF2 21.7 0.0032 6.5

F
hydrL _BF2 16.5 0.0080 5

Table 6.3: Brightness, efficiency and turn-on voltage of the devices described so far.

112
Figure 6.5: Luminance current voltage measurements of the devices described so far. Color code for the
emissive layer: black, LOMe_BF2 with no OTPD; red, LOMe_BF2; green, LOMe_BF2 30% w/w in PVK; blue, LH_BF2;
light blue, hydrLF_BF2.

At first, a maximum brightness of 22 Cd/m2 at 20 V for the device built using LOMe_BF2
dispersed in PVK as emissive layer was observed. Unfortunately, the efficiencies recorded
are very low, with the highest among the working devices of just 0.0076 Cd/A. The turn-on
voltages attested between 5 and 8 V. All the data recorded in this experiment are
summarized in Table 6.3.
The layers thicknesses have been ultimately obtained using a profilometer (see Table 6.2).
In these measurements, thanks to the optical camera mounted on the instrument we could
closely observe the emissive layers, noticing a difference between those made out of pure
emissive material and the one made with PVK as matrix. In fact, the formation of small
agglomerates of material was observed for the pure emitter, while the layer made with
PVK resulted smooth and more homogeneous (Figure 6.6). This was also reflected by the
emission observed at naked eye of the devices, with the latter emitting sky blue light,
whereas the pure layer-devices (i.e. without PVK) displayed a blue light background with
red-pink dots that we hypothesized were due to some interaction inside those small
aggregates, plausibly tiny crystals.

113
Figure 6.6: Surface pictures of the device made using pure LOMe_BF2 (left) and LOMe_BF2 dispersed in PVK
(right) taken using the profilometer’s camera.

Considering this last aspect, we tried to enhance both efficiency and brightness of these
devices by increasing the concentration of emitter in PVK up to 45% w/w.
In this way, we managed to achieve a brightness of up to 18.9 Cd/m2, observing an increase
with increasing the emitter concentration in PVK, while the efficiency, notably higher than
for pure emitter, decreased with increasing the concentration (Figure 6.7 and Table 6.4).

Brightness Efficiency Turn-on voltage


Emissive layer 2
(Cd /m ) (Cd/A) (V)

OMe
L _BF2 14.6 0.0033 8

LOMe_BF2 30% w/w in PVK (no annealing) 3.42 0.006 18

LOMe_BF2 30% w/w in PVK 5.7 0.079 14.5

LOMe_BF2 40% w/w in PVK 12.6 0.069 10

LOMe_BF2 45% w/w in PVK 16.4 0.061 10

LH_BF2 30% w/w in PVK 10.5 0.092 10

hydrLF_BF2 30% w/w in PVK 20.1 0.013 10

Table 6.4: Brightness, efficiency and turn-on voltage of the devices made with PVK.

114
Figure 6.7: Luminance current voltage measurements of the devices made with PVK. Color code for the
emissive layer: blue, LOMe_BF2 no PVK; light blue, LOMe_BF2 30% w/w in PVK (no annealing); brown, LOMe_BF2
30% w/w in PVK; pink, LOMe_BF2 40% w/w in PVK; purple, LOMe_BF2 45% w/w in PVK; green, LH_BF2 30% w/w
in PVK; yellow, hydrLF_BF2 30% w/w in PVK.

Observing the devices with the profilometer’s camera provided a rationalization of this
aspect: in fact, at higher concentration, a phase separation occurred, leading to the
formation of crystals of the emissive species in the polymeric matrix, thus reducing the film
homogeneity and, consequently, its performances in terms of efficiency, with no net
enhancement of its brightness (Figure 6.8).
At this point, two main possible issues could be solved to improve the properties of the
devices, which will be discussed in the following sections: an “energetic” problem (Section
6.3) and a film quality problem (Section 6.4).

Figure 6.8: Surface pictures of the device made using LOMe_BF2 40% w/w in PVK (left) and LOMe_BF2 45% w/w
in PVK (right) taken using the profilometer’s camera.

115
6.3 ENERGETIC ISSUE
First, we wanted to evaluate the experimental HOMO energy of LOMe_BF2 to be sure that
the energetic step between the frontier orbitals of the HTM and emissive material was not
too wide, which could be an explanation for the poor efficiencies of the previously
fabricated devices. A cyclovoltammetric analysis have been performed, revealing an
experimental HOMO energy of -5.91 eV (using Equation 6.1) (calculated -5.339 eV using
PBE-D3 functional, see Section 4.3), too far from the HOMO energy of OTPD (-5.44 eV)
(Figure 6.9) (see Section 8.1 for experimental details).

/ /
𝐸 = −(𝐸 −𝐸 ) − 4.6 = −(0.756 − (−0.545)) − 4.6 = −5.91 𝑒𝑉

Equation 6.1

Figure 6.9: Cyclovoltammetric analysis of LOMe_BF2.

Starting from this experimental data, a benchmark of other DFT functionals (PBE0, B3LYP
and M06) have been performed, revealing PBE0 to be the most accurate for this system,
with a calculated HOMO energy of -5.843 eV (calculated with PBE-D3: -5.048 eV).
Also, we tried to build devices with additional HTL, in order to lessen the energetic step and
make the hole injection in the emissive layer more favorable. In these regards, two HTM
have been chosen, N4,N4'-bis(4-(6-((3-ethyloxetan-3-yl)methoxy)hexyl)phenyl)-N4,N4'-
bis(3,4,5-trifluorophenyl)-[1,1'-biphenyl]-4,4'-diamine (X2F6TPD) and N4,N4'-bis(3,4-
difluorophenyl)-N4,N4'-bis(4-(6-((3-ethyloxetan-3-yl)methoxy)hexyl)phenyl)-[1,1'-

116
biphenyl]-4,4'-diamine (X2F4TPD) (Figure 6.9), with HOMO energies of -5.66 eV and -5.56
eV, respectively. The first one to be tested, because of its closer HOMO energy to that of
LOMe_BF2, was X2F6TPD. 190 μL of a 5mg/mL toluene solution with 2% molar concentration
of OPPI was spin coated at 1500 rpm for 20 seconds, followed by a curing process identical
to the one described for the previous HTLs (UV and heat). Spin coating the emitter solution
on top of this HTL led to the formation of layers of just 2-5 nm, which are of course too thin
to make the hole-electron recombination happen inside the emissive layer itself, thus
lowering the efficiency of these devices: for instance, when compared to the efficiency of
the device built without X2F6TPD, a loss of around 75% was observed (see brown trace in
Figure 6.11 and Table 6.5) Also, trying to disperse the emissive material using X2F6TPD as
host matrix led to poor results. Also, brightness was not enhanced significantly by this new
architecture.

Figure 6.10: Structures of X2F4TPD and X2F6TPD.

117
Figure 6.11: Luminance current voltage measurements of the devices made with X2F6TPD. Color code for
the emissive layer: light pink, LOMe_BF2 (no X2F6TPD); pink, LOMe_BF2 30% w/w in X2F6TPD; red, LOMe_BF2;
brown, LOMe_BF2 45% w/w in PVK (no X2F6TPD); purple, LOMe_BF2 45% w/w in PVK; light blue, LH_BF2 (no
X2F6TPD); blue, LH_BF2; light green, hydrLF_BF2 (no X2F6TPD); green, hydrLF_BF2.

X2F6TPD Brightness Efficiency Turn-on voltage


Emissive material Matrix 2
(as HTL) (Cd /m ) (Cd/A) (V)

OMe
L _BF2 None No 18.5 0.0039 7

LOMe_BF2 30% w/w X2F6TPD No 14.6 0.0066 9

LOMe_BF2 None Yes 10.4 0.0003 9

LOMe_BF2 45% w/w PVK No 22.2 0.043 11

LOMe_BF2 45% w/w PVK Yes 11.77 0.012 12.5

LH_BF2 None No 20.3 0.0030 8

LH_BF2 None Yes 14.9 0.0016 9

hydrLF_BF2 None No Not working Not working Not working

hydrLF_BF2 None Yes 10.4 0.0017 9.5

Table 6.5: Brightness, efficiency and turn-on voltage of the devices made with X2F6TPD.

Using X2F4TPD provided similar results with emissive layers too thin to be tested.

118
6.4 FILM MORPHOLOGY ISSUE
To enhance the films quality we tested other matrixes, typically used as host materials in
the vapor deposition method: (oxybis(2,1-phenylene))bis(diphenylphosphine oxide)
(DPEPO), 1,3-di(9H-carbazol-9-yl)benzene (mCP), PVK doped with 7% w/w of 1,3-bis(5-
(tert-butyl)-1,3,4-oxadiazol-2-yl)benzene (OXD7) and, finally, 9-(4-(tert-butyl)phenyl)-3,6-
bis(triphenylsilyl)-9H-carbazole (CzSi) (Figure 6.12). Looking at the films with the
profilometer revealed that only CzSi led to the formation of homogenous films, The devices
made in this way were very similar in terms of brightness to those obtained with PVK
(Figure 6.13 and Table 6.6), but with lower efficiency. For this reason, and also because of
its extremely high cost, CzSi was discarded as well, leaving PVK the only matrix suitable to
be used for these devices.

Figure 6.12: Structures of the alternative host materials tested.

Figure 6.13: Luminance current voltage measurements of the devices made with CzSi and PVK. Color code
for the emissive layer: pink, LOMe_BF2 50% w/w in PVK; light blue, LOMe_BF2 50% w/w in CzSi.

119
Brightness Efficiency Turn-on voltage
Emissive layer 2
(Cd /m ) (Cd/A) (V)

OMe
L _BF2 50% w/w in PVK 17.2 0.011 13

LOMe_BF2 50% w/w in CzSi 13.6 0.006 12.5

Table 6.6: Brightness, efficiency and turn-on voltage of the devices made with CzSi and PVK.

Finally, we observed that by skipping the annealing process after the deposition of the
emissive layer, no phase separation occurred (Figure 6.14). With this shrewdness, it is very
improbable that all the solvent (DMF) was removed, because of its high boiling point, but
nevertheless, we managed to obtain the best performances among the various
experiments made on the fabrication of OLEDs with LOMe_BF2, reaching brightness more
than ten times higher than those obtained previously. We proceeded fabricating a series
of devices with the same architecture for all the LR_BF2 compounds, both as 50% w/w
concentration dispersion in PVK and as pure materials. Also, silver was used as cathode
instead of aluminum, due to some technical problems with the evaporation chamber.

Figure 6.14: Surface pictures of the device made using LOMe_BF2 50% w/w in PVK after (left) and before
(right) the annealing step, taken using the profilometer’s camera.
.

LMe_BF2, LOMe_BF2, LF_BF2 and LH_BF2 dispersed in PVK displayed a brightness from 59.8
(LH_BF2) to 298.8 (LMe_BF2) Cd/m2, comparable to commercially available OLEDs. The
efficiencies ranged from 0.15 Cd/A (LMe_BF2) to 0.012 Cd/A (LH_BF2) for these four devices.
Notably, heavy-atom substituted compounds (LBr_BF2 and LI_BF2), as well as, as expected,
the nitro derivative displayed no luminescence at all (Figure 6.15 and Table 6.7).

120
The devices built using pure materials as emissive layers displayed instead, in every case, a
luminescence brightness of 10 to 15 Cd/m2, even those made with materials which showed
no emission in PVK, nor in preliminary studies in solution (see Chapter 4.2). The very similar
emission displayed by these devices could be ascribed to the luminescence properties of
the materials below the emissive layer, which, even if poor, were still present and led to
this low-intensity blue electroluminescence.

Figure 6.15: Luminance current voltage measurements of the LR_BF2 devices made without the annealing
step. Color code for the emissive layer: black, LOMe_BF2 50% w/w in PVK; red, LOMe_BF2; light green, , LH_BF2
50% w/w in PVK; blue, LH_BF2; blue, LMe_BF 50% w/w in PVK2; pink, LMe_BF2; yellow, LF_BF2 50% w/w in PVK;
olive, LF_BF2; dark blue, LBr_BF2 50% w/w in PVK; purple, LBr_BF2; brown, LBr_BF2 50% w/w in PVK; dark
green, LI_BF2; blue-green, LNO2_BF2 50% w/w in PVK; navy blue, LNO2_BF2.

121
Brightness Efficiency Turn-on voltage
Emissive layer 2
(Cd /m ) (Cd/A) (V)

OMe
L _BF2 50% w/w in PVK 145.1 0.049 11

LOMe_BF2 10.7 0.0038 6

LH_BF2 50% w/w in PVK 59.8 0.012 12.5

LH_BF2 10.1 0.0024 7.5


Me
L _BF2 50% w/w in PVK 298.8 0.15 9.5

LMe_BF2 12.6 0.005 5.5

LF_BF2 50% w/w in PVK 109.9 0.063 11

LF_BF2 9.0 0.00087 6

LBr_BF2 50% w/w in PVK Not working Not working Not working

LBr_BF2 8.7 0.00085 6

LI_BF2 50% w/w in PVK 20.6 0.00084 20

LI_BF2 9.0 0.00080 6.5

LNO2_BF2 50% w/w in PVK Not working Not working Not working

LNO2_BF2 10.7 0.00081 7.5

Table 6.7: Brightness, efficiency and turn-on voltage of the LR_BF2 devices made without the annealing step

For the brightest device (the one with LMe_BF2 in PVK), an electroluminescence emission
spectrum has been recorded (Figure 6.16), revealing a maximum of emission at 516 nm in
the light blue-green region of the visible spectrum, using a voltage of 16 V at which the
maximum brightness value was observed. Unfortunately, due to the too low brightness of
the other devices, it has not been possible to record other electroluminescence emission
spectra. Nonetheless, in can be said that they all displayed a blue luminescence visible to
the naked eye.

122
Electroluminescence of LMe_BF2 (50% w/w in PVK)
1600
1400
1200
Intensity (a.u)
1000
800
600
400
200
0
300 400 500 600 700 800
Wavelength (nm)

Figure 6.16: Electroluminescence emission spectrum of the device build with LMe_BF2 50% w/w in PVK as
emissive layer, without the annealing step.

6.5 POSSIBLE FUTURE OUTLOOKS


Considering what has been described so far, LR_BF2 compounds represent a promising class
of blue emissive materials that could be used in the fabrication of blue emissive OLEDs. The
performances of these devices could be greatly improved, tackling the two main issues
affecting and limiting the brightness and, more importantly, the efficiency of the organic
LEDs build so far. On one side, other R groups can be used, in order to heighten the HOMO
energies of these compounds, making the hole injection from the HTL more efficient,
without losing the blue emission, typical of this family of molecules, as discussed in Chapter
4. On the other hand, having more soluble compounds, especially in toluene or
chlorobenzene, would in principle lead to better film-forming properties, thus ultimately
enhancing the properties of the devices. In these regards, long alkyl chains can be used
instead of the methyl group in position 1 on the imidazo[1,5-a]pyridine ring, or other non-
polar groups (i.e. tert-butyl groups) could be put on the phenolic ring for the same purpose.
Finally, other kind of functionalization can be used in order to improve the luminescence
properties (fluorescent quantum yields) of the materials themselves, for example
substituting the fluorine atoms of the BF2 fragments with acetylenic units or with a diphenyl
group (Figure 6.17).

123
R
R
R

O
N
N B O
O
N N O
N B N B
R' F N
F N B F
CH2 F
n
R' H 3C CH3

Figure 6.17: Possible new species to be tested in order to solve the “energetic” problem (left) and
to enhance the film forming properties of the system (right).

To conclude, the devices reported in this chapter are among the first examples of organic
light emitting diodes built using imidazo[1,5-a]pyridine derivatives as emissive layers. The
encouraging results, pave the way to further investigation on the possibility to obtain
efficient blue emissive devices from this cheap and easily accessible class of compounds.

Figure 6.18: Picture of an OLED device fabricated in this work (left) and electroluminescence of the same
OLED at work (right).

124
Chapter 7

CONCLUSIONS

125
In this Thesis work, three new classes of luminescent coordination compounds with
imidazo[1,5-a]pyridine ligands have been prepared and fully characterized, using zinc and
boron as coordination center. Moreover, their fluorescence behavior in solution and, in
some cases, in the solid state as well as in polymeric films have been evaluated, in order to
have a deeper understanding of their structure-properties correlation and of their potential
application as emissive layer in Organic Light Emitting Diodes.

In Chapter 3, a series of homoleptic zinc complexes with imidazo[1,5-a]pyridine-3-yl


phenols as N-O bidentate monoanionic ligands has been described. The fluorescence
properties of these species have been studied in solution, for which it could be observed a
transfer from the ligand to the complex, with an enhancement due to the coordination. In
fact, DFT calculations provided an understanding of the transitions responsible for the
emissive properties: the zinc center is not involved in the transitions, ultimately having a
structural role, and enhancing the fluorescence quantum yield by giving more rigidity to
the system. This series of zinc complexes represents a class of emissive materials whose
emission wavelengths range from blue to red, depending on the electronic properties of
the ligands, tuned by the substituent on the phenolic ring.

In Chapter 4, the same ligands described before have been used in the preparation of some
boron difluoride compounds. A full characterization of these species has been performed
using infrared and various NMR analysis (1H, 13C, 19F, 11B). The fluorescence properties of
LR_BF2 compounds have been studied in dichloromethane solution, defining this class of
boron compounds as blue emissive materials, with emissions very close to those expected
for standard blue. Noteworthy, this blue emission seemed to be almost independent from
the substituent on the organic ligands, giving access to multiple possible functionalizations
without losing this very interesting characteristic, thus paving the way to different
applications in the blue emissive devices technology. Finally, the different behavior of
compound LH_BF2 depending on its hydration has been evaluated.

In Chapter 5, another series of boron difluoride compounds has been investigated. This
time, the ligands were derived from the previously used imidazo[1,5-a]pyridines by
hydrogenation, leading to a class of imidazo[1,5-a]piperidines, which, from a work

126
previously done among our research group, displayed better photophysical properties if
compared to the corresponding non-hydrogenated derivatives. The luminescence
properties of these system have been evaluated in solution, in the solid state and, finally,
in polymeric films, using PMMA as host material. In every case a bright blue emission was
observed, with wavelength depending on the substituent on the phenolic ring of the ligand
itself. Furthermore, DFT calculations showed how the excited state geometries of these
compounds is greatly distorted if compared to the ground state geometries, thus giving a
possible explanation for their singular trend in terms of emission maxima vs. solvent
polarity (i.e. higher wavelength with decreasing the solvent polarity).

In Chapter 6, the work done at the Chemistry Department of the University of Cologne has
been presented. In particular, the second class of compounds, the boron difluoride
compounds of the non-hydrogenated ligands, has been tested in the fabrication of OLEDs.
This has been possible thanks to their good processability properties for the spin coating
fabrication method, despite their fluorescence properties are not the best among the three
series described in this Thesis. Many way to enhance the performances of the devices built
have been reported, starting from dispersing the emissive materials in PVK as polymeric
matrix, coming to the use of more Hole Transporting Layers and, ultimately, evaluating the
effect of the annealing process on the film quality. Finally, a series of OLEDs made using
LR_BF2 compounds dispersed in PVK as emissive layer has been obtained, with brightness
up to almost 300 Cd/m2, reaching values comparable to those of standard commercially
available devices.

In conclusion, despite additional studies are needed to fully optimize the performances of
the abovementioned OLEDs, both in terms of brightness and efficiency, this work
highlighted the possibility to utilize imidazo[1,5-a]pyridine derivatives as emissive layer for
optoelectronic application, with these examples being among the first reported in the
literature for this class of compounds.

127
Chapter 8

EXPERIMENTAL SECTION

128
8.1 GENERAL REMARKS

All the starting reagents have been purchased (TCI Chemicals, Fluorochem) and used without
further purifications. All reactions have been carried out under nitrogen atmosphere using the
conventional Schlenk technique, deoxygenating the solvents prior to use. Hydrogenation reactions
have been performed in a 100 mL PARR stainless stell autoclave reactor equipped with an Ashcroft
Duralife (3000 psi) pressure gauge. Infrared spectra (nujol mull) have been recorded on a Shimadzu
Prestige-21 spectro-photometer with a 1 cm-1 resolution, while ATR spectra have been recorded on
a Thermo Scientific Nicolet iS20 FTIR Spectrometer with a 1 cm-1 resolution. NMR spectra have been
recorded with an AVANCE III HD Bruker spectrometer at 400 MHz for 1H-NMR, 100 MHz for 13C{1H}-
NMR, 376 MHz for 19F-NMR and 128 MHz for 11B-NMR, giving chemical shifts in ppm relative to
residual solvent peaks as internal reference. J values are given in Hz. Elemental analyses have been
obtained with a Perkin-Elmer CHN Analyzer 2400 Series II. UV-vis, emission and excitation spectra
have been recorded on a Edinburgh Instruments FS5 fluorescence spectrometer equipped with a
150 W continuous Xenon lamp, and corrected for the wavelength response. Lifetime measurements
have been performed on the same instrument using an EPLED-320 as pulsed source. Absolute
fluorescence quantum yields in solution were determined with a PhotoMed GmbH K-Sphere
Integrating Sphere of 3.2 inch. diameter. Analysis of the lifetime decay curve and determination of
absolute quantum yields in solution have been done using Fluoracle Software package. Absolute
fluorescence quantum yields in the solid state and in polymeric films have been measured on a
Photon Technology International QuantaMaster QM-40 spectrometer equipped with a Xe arc lamp
(70 W), using the abovementioned integrating sphere. The integrate luminescence areas have been
obtained with the Felix32 analyses software and used to calculate the absolute photoluminescence
quantum yields with uncertainties of ±5% according to the literature152 with the equation below:

𝐸 (𝜆) − (1 − 𝑎) ⋅ 𝐸 (𝜆)
𝜙 =
𝐿 (𝜆) ⋅ 𝐴
where
𝐿 (𝜆) − 𝐿 (𝜆)
𝐴=
𝐿 (𝜆)
Ei(λ) = integrated emission under direct excitation of the sample;
E0(λ)= integrated emission under secondary excitation of the sample;
Li(λ) = integrated excitation under direct excitation of the sample;
L0(λ)= integrated excitation under secondary excitation of the sample;
Le(λ)= integrated excitation profile for the empty sphere.

129
Thermogravimetric analyses (TGA) were performed in a N2 stream (20 mL/min) on a Netzsch STA
409 PC Luxx (heating rate 20 °C/min). Structural analyses have been performed by Prof. Bruno
Therrien (University of Neuchâtel) and conducted using a Stoe Image Plate Diffraction system
equipped with a circle goniometer, using Mo-Kα graphite monochromated radiation (λ = 0.71073
Å) with φ range 0–200°. The structures have been solved by direct methods using SHELXS program,
refinement and further calculations using SHELXL. H-atoms have been included in calculated
positions and treated as riding atoms using SHELXL default parameters. Non-H atoms have been
refined anisotropically, using weighted full-matrix least-square on F2.
OLED devices have been fabricated at the University of Cologne, in Prof. Klaus Meerholz’s research
group. All the fabrications have been performed under nitrogen atmosphere using gloveboxes. The
glass substrates have been treated with ozone (10 min) prior to spin coating process. A Leybold
Univex 450 evaporation chamber has been used for CsF, which has been evaporated at a rate of 0.2
Å/s, subsequently Al or Ag has been evaporated at the same rate. Luminance current voltage
measurements have been performed using a Keithley 2004 source meter and the luminance values
have been measured with a photodiode calibrated with a Minolta Chroma meter CS-100.
Electroluminescence measurements have been performed under argon atmosphere using a
USB4000 Uv-vis-ES Ocean Optic spectrometer. Thickness analysis have been performed using a
DEKTAK XT stylus profilometer. Cyclovoltammetry has been carried out in a 0.1 M TBAPF6
dichloromethane solution with a three-electrode configuration (working electron: glassy carbon,
counter electrode: platinum, pseudo reference electrode: Ag/AgCl) and a BioLogic SP-150
potentiostat. The decamethylferrocene/decamethylferrocenium couple has served as internal
reference and the experiment has been run at a scan rate of 0.1 V/s.

8.2 SYNTHESIS OF HLR LIGANDS

2-acetylpyridine (1 equivalents), 5-substituted salicylaldehyde (2 equivalents) and ammonium


acetate (5 equivalents) are mixed in 40 mL of deoxygenated acetic acid. The reaction is left stirring
under inert atmosphere for one week. Then, the mixture is filtered and poured into 150 mL of water
and extracted with dichloromethane (3 x 75 mL). The recombined organic phase is washed with a
saturated NaHCO3 solution and dried over Na2SO4. The suspension is filtered and the solvent
evaporated under reduced pressure. Then, the crude product is triturated in hexane.
If R = NO2, the product is obtained with the first filtration, washed with acetic acid and water and
dried under vacuum.

130
8.3 SYNTHESIS OF [Zn(LR)2] COMPOUNDS

HLR (2 equivalents) are dissolved in 10 mL of deoxygenated acetonitrile and zinc perchlorate


hexahydrate (1 equivalent) is added to the solution. The reaction is stirred for 15 min, then
triethylamine (3.5 equivalents) is added. The suspension is left stirring for 12 h, then the solid is
filtered, washed with water, acetonitrile and diethyl ether and dried under vacuum.

8.3.1 [Zn(LH)2]
Yield 74 %. Elemental analysis (%) calcd. for C28H22N4O2Zn: C 65.70, H 4.33, N 10.95. Found C 65.60, H 4.48, N
10.74 %.1H-NMR (400 MHz, DMSO-d6, 25 °C): δ = 2.12 (s, 3H), 6.69 (d,J = 7.8 Hz, 1H), 6.85 (m, 3H), 7.20 (t, J=
7.9 Hz, 1H), 7.65 (t, J = 7.1 Hz, 2H), 8.56 (d, J = 7.4 Hz, 1H).

8.3.2 [Zn(LMe)2]
Yield 58 %. Elemental analysis (%) calcd. for C30H26NO2Zn: C 66.73, H 4.85, N 10.38. Found C 66.49, H 4.80, N
10.26 %. 1H-NMR (400 MHz, DMSO-d6, 25 °C): δ = 2.10 (s, 3H), 2.27 (s, 3H), 6.75 (d, J = 8.3 Hz, 1H), 6.84 (dt, J
= 17.6, 6.6 Hz, 2H), 7.01 (d, J = 8.4 Hz, 1H), 7.43 (s, 1H), 7.62 (d, J = 8.9 Hz, 1H), 8.59 (d, J = 7.3 Hz, 1H).

8.3.3 [Zn(LOMe)2]
Yield 68 %. Elemental analysis (%) calcd. for C30H26N4O4Zn: C 63.00, H 4.58, N 9.80. Found C 62.80, H 4.67, N
9.55 %. 1H-NMR (400 MHz, DMSO-d6, 25 °C): δ = 2.11 (s, 3H), 3.75 (s, 3H), 6.78 (d, J = 9.0 Hz, 1H), 6.88 (m,
3H), 7.16 (d, J = 3.1 Hz, 1H), 7.66 (d, J = 8.9 Hz, 1H), 8.64 (d, J= 7.3 Hz, 1H).

8.3.4 [Zn(LF)2]
Yield 56 %. Elemental analysis (%) calcd. for C28H20F2N4O2Zn: C 61.38, H 3.68, N 10.23. Found C 61.60, H 3.52,
N 10.11 %. 1H- NMR (400 MHz, DMSO-d6, 25 °C): δ = 2.14 (s, 3H), 6.87 (m, 3H), 7.05 (s, 1H), 7.46 (d, J = 8.9 Hz,
1H), 7.66 (d, J = 8.9 Hz, 1H), 8.60 (s, 1H).

8.3.5 [Zn(LBr)2]
Yield 47 %. Elemental analysis (%) calcd. for C28H20Br2N4O2Zn: C 50.22, H 3.01, N 8.37. Found C 50.45, H 3.12,
N 8.60 %. 1H-NMR (400 MHz, DMSO-d6, 25 °C): δ = 2.18 (s, 3H), 6.83(m, 3H), 7.29 (s, 1H), 7.59–7.85 (m, 2H),
8.50 (s, 1H).

131
8.3.6 [Zn(LCF3)2]
Yield 68 %. Elemental analysis (%) calcd. for C30H20F6N4O2Zn: C 55.62, H 3.11, N 8.65. Found C 55.33, H 3.18,
N 8.91 %. 1H-NMR (400 MHz, DMSO-d6, 25 °C): δ = 2.23 (s, 3H), 6.82–7.14 (m, 3H), 7.37–7.63 (m, 1H), 7.70 (d,
J = 9.0 Hz, 1H), 7.94 (s, 1H), 8.39–8.57 (m, 1H).

8.3.7 [Zn(LNO2)2]
Yield 60 %. Elemental analysis (%) calcd. for C28H20N6O6Zn: C 55.88, H 3.35, N 13.96. Found C 55.56, H 3.21, N
13.82 %. 1H-NMR (400 MHz, DMSO-d6, 25 °C): δ = 2.41 (s, 3H), 6.76 (d, J = 9.3 Hz, 1H), 6.88 (dt, J = 17.4, 6.6
Hz, 2H), 7.71 (d, J = 8.9 Hz, 1H), 8.04 (dd, J = 9.3, 3.0 Hz, 1H), 8.53 (t, J = 6.5 Hz, 2H).

Figure 8.1: 1H-NMR of [Zn(LH)2] in DMSO-d6.

132
Figure 8.2: 1H-NMR of [Zn(LMe)2] in DMSO-d6.

Figure 8.3: 1H-NMR of [Zn(LOMe)2] in DMSO-d6.

133
Figure 8.4: 1H-NMR of [Zn(LF)2] in DMSO-d6.

Figure 8.5: 1H-NMR of [Zn(LBr)2] in DMSO-d6.

134
Figure 8.6: 1H-NMR of [Zn(LCF3)2] in DMSO-d6.

Figure 8.7: 1H-NMR of [Zn(LNO2)2] in DMSO-d6.

135
8.4 SYNTHESIS OF LR_BF2 COMPOUNDS

HLR (1 equivalent) is dissolved in 6 mL of deoxygenated dichloromethane, then, and boron


trifluoride diethyletherate (2.5 equivalents) is added to the solution. Finally, triethylamine (1.2
equivalents) is added to the reaction mixture. The reaction is stirred for 1 hour, then the suspension
is filtered and the solid washed with cold dichloromethane and dried under vacuum.

8.4.1 LH_BF2
Yield 59.4 %. Elemental analysis (%) calcd. for C14H11BF2N2O: C 61.81, H 4.08, N 10.30. Found C 61.55, H 4.01,
N 10.18. 1H-NMR (400 MHz, CD2Cl2, 25 °C): δ = 8.52 – 8.41 (m, 1H), 7.82 (dd, J = 8.1, 1.5 Hz, 1H), 7.56 – 7.47
(m, 1H), 7.37-7.32 (m, 1H), 7.12 (dd, J = 8.3, 1.2 Hz, 1H), 7.03-6.98 (m, 1H), 6.94 – 6.80 (m, 2H), 2.62 (s, 3H).
13
C-NMR (100 MHz, CD2Cl2, 25 °C): δ = 154.92, 122.35, 122.18, 120.67, 120.10, 119.75, 119.01, 117.19, 9.34.
19
F- NMR (376 MHz, CD2Cl2, 25 °C): δ = -140.01 (q, J = 13.6 Hz). 11B-NMR (128 MHz, CD2Cl2, 25 °C): δ = 1.15 (t,
J = 14.9 Hz). FT-IR (ATR): (cm-1): ṽ = 1033 (s), 1061 (s) 1093 (s).

8.4.2 LMe_BF2
Yield 59.3 %. Elemental analysis (%) calcd. for C15H13BF2N2O: C 62.98, H 4.58, N 9.79. Found C 62.79, H 4.62,
N 9.73. 1H-NMR (400 MHz, CD2Cl2, 25 °C): δ = 8.84 – 8.43 (m, 1H), 7.86 – 7.67 (m, 1H), 7.67 – 7.49 (m, 1H),
7.38 – 7.23 (m, 1H), 7.13 (d, J = 8.4 Hz, 1H), 7.05 – 6.88 (m, 2H), 2.73 (s, 3H), 2.44 (s, 3H). 13C-NMR (100 MHz,
CD2Cl2, 25 °C): δ = 152.81, 132.74, 129.18, 127.55, 122.47, 122.14, 120.57, 119.79, 118.96, 117.04, 110.53,
20.61, 9.34. 19F-NMR (376 MHz, CD2Cl2, 25 °C): δ = -140.28 (q, J = 13.6 Hz). 11B-NMR (128 MHz, CD2Cl2, 25 °C,):
δ = 1.18 (t, J = 15.3 Hz). FT-IR (ATR): (cm-1): ṽ = 1036 (s), 1066 (s) 1090 (s).

8.4.3 LOMe_BF2
Yield 58.0 %. Elemental analysis (%) calcd. for C15H13BF2N2O2: C 59.64, H 4.42, N 9.27. Found C 59.38, H 4.31,
N 9.20. 1H-NMR (400 MHz, CD2Cl2, 25 °C): δ = 8.62 – 8.52 (m, 1H), 7.68 – 7.59 (m, 1H), 7.43 (d, J = 2.8 Hz, 1H),
7.18 (d, J = 9.0 Hz, 1H), 7.07 (dd, J = 9.0, 2.9 Hz, 1H), 7.05 – 6.95 (m, 2H), 3.90 (s, 3H), 2.74 (s, 3H). 13C-NMR
(100 MHz, CD2Cl2, 25 °C): δ = 152.76, 149.05, 127.71, 122.72, 122.29, 120.71, 120.61, 119.04, 117.53, 117.25,
110.82, 107.63, 56.06, 9.36. 19F-NMR (376 MHz, CD2Cl2, 25 °C): δ = -140.71 (q, J = 13.1 Hz). 11B-NMR (128 MHz,
CD2Cl2, 25 °C): δ = 1.15 (t, J = 14.8 Hz). FT-IR (ATR): (cm-1): ṽ = 1032 (s), 1055 (s) 1084 (s).

8.4.4 LF_BF2
Yield 61.7 %. Elemental analysis (%) calcd. for C14H10BF3N2: C 57.97, H 3.48, N 9.66. Found C 57.61, H 3.37, N
9.48. 1H- NMR (400 MHz, CD2Cl2, 25 °C): δ = 8.58 – 8.47 (m, 1H), 7.82 – 7.54 (m, 2H), 7.25 – 7.17 (m, 2H), 7.14
– 7.00 (m, 2H), 2.76 (s, 3H). 13C-NMR (100 MHz, CD2Cl2, 25 °C): δ = 157.02, 154.66, 128.00, 122.05, 121.06,
119.15, 118.68, 118.45, 117.68, 108.40, 108.14, 9.36. 19F-NMR (376 MHz, CD2Cl2, 25 °C): δ = -123.56, -140.17

136
(q, J = 13.1 Hz). 11B-NMR (128 MHz, CD2Cl2, 25 °C): δ = 1.17 (t, J = 14.4 Hz). FT-IR (ATR): (cm-1): ṽ = 1043 (s),
1066 (s) 1106 (s).

8.4.5 LCl_BF2
Yield 61.8 %. Elemental analysis (%) calcd. for C14H10BF2ClN2O: C 53.43, H 3.20, N 8.90. Found C 53.46, H 3.40,
N 9.07. 1H-NMR (400 MHz, CD2Cl2, 25 °C): δ = 8.68 – 8.32 (m, 1H), 7.78 (m, 1H), 7.66 – 7.54 (m, 1H), 7.34 (dd,
J = 8.9, 2.4 Hz, 1H), 7.18 (d, J = 8.9 Hz, 1H), 7.06 – 6.92 (m, 2H), 2.74 (s, 3H). 13C-NMR (100 MHz, CD2Cl2, 25
°C): δ = 153.66, 131.70, 129.73, 127.83, 124.61, 123.41, 122.14, 121.93, 121.37, 121.05, 119.29, 117.71,
111.72, 9.64. 19F-NMR (376 MHz, CD2Cl2, 25 °C): δ = -139.80 (q, J = 11.4 Hz). 11B-NMR (128 MHz, CD2Cl2, 25
°C): δ = 1.14 (t, J = 14.2 Hz). FT-IR (ATR): (cm-1): ṽ = 1033 (s), 1085 (s) 1103 (s).

8.4.6 LBr_BF2
Yield 72.5 %. Elemental analysis (%) calcd. for C14H10BF2N2OBr: C 47.91, H 2.87, N 7.98. Found: C 47.75, H 2.96,
N 7.63. 1H-NMR (400 MHz, CD2Cl2, 25 °C): δ = 8.54 – 8.37 (m, 1H), 7.92 (d, J = 2.3 Hz, 1H), 7.67 – 7.49 (m, 1H),
7.43 (dd, J = 8.8, 2.3 Hz, 1H), 7.03 (d, J = 8.8 Hz, 1H), 6.99 – 6.86 (m, 2H), 2.63 (s, 3H). 13C-NMR (100 MHz,
CD2Cl2, 25 °C): δ = 153.95, 134.37, 128.05, 124.49, 123.15, 122.22, 121.90, 121.14, 119.13, 117.76, 111.51,
9.36. 19F-NMR (376 MHz, CD2Cl2, 25 °C): δ = -139.89 (q, J = 12.7 Hz). 11B-NMR (128 MHz, CD2Cl2, 25 °C): δ =
1.03 (t, J = 14.4 Hz). FT-IR (ATR): (cm-1): ṽ = 1032 (s), 1069 (s) 1094 (s).

8.4.7 LI_BF2
Yield 55.4 %. Elemental analysis (%) cacld. for C14H10BF2IN2O: C 42.25, H 2.53, N 7.04. Found C 41.95, H 2.68,
N 7.00. 1H-NMR (400 MHz, CD2Cl2, 25 °C): δ = 8.49 – 8.41 (m, 1H), 8.07 (d, J = 2.1 Hz, 1H), 7.69 – 7.54 (m, 2H),
7.09 – 6.92 (m, 3H), 2.73 (s, 3H). 13C-NMR (100 MHz, CD2Cl2, 25 °C): δ = 154.71, 140.39, 130.12, 129.29, 127.82,
123.36, 122.78, 122.15, 120.96, 119.26, 117.74, 113.25, 80.89, 9.66. 19F-NMR (376 MHz, CD2Cl2, 25 °C): δ =
139.74 (q, J = 11.4 Hz). 11B-NMR (128 MHz, CD2Cl2, 25 °C): δ = 1.08 (t, J = 14.3 Hz). FT-IR (ATR): (cm-1): ṽ = 1036
(s), 1056 (s) 1094 (m).

8.4.8 LNO2_BF2
Yield 73.6 %. Elemental analysis (%) calcd. for C14H10BF2N3O3: C 53.04, H 3.18, N 13.25. Found: C 53.03, H 3.43,
N 12.91. 1H-NMR (400 MHz, CD2Cl2, 25 °C): δ = 8.91 (d, J = 2.4 Hz, 1H), 8.70 – 8.67 (m, 1H), 7.86 – 7.67 (m,
1H), 7.34 (dd, J = 9.1, 1.3 Hz, 1H), 7.26 – 7.08 (m, 2H), 2.79 (s, 3H). 13C-NMR (100 MHz, CD2Cl2, 25 °C): δ =
159.91, 140.25, 128.54, 122.16, 121.75, 121.15, 120.56, 119.26, 118.56, 118.40, 115.13, 114.98, 9.37. 19F-
NMR (376 MHz, CD2Cl2, 25 °C): δ = -138.77 (q, J = 12.4 Hz). 11B-NMR (128 MHz, CD2Cl2, 25 °C): δ = 1.05 (t, J =
13.4 Hz). FT-IR (ATR): (cm-1): ṽ = 1061 (s) 1093 (s).

137
Figure 8.8: 1H-NMR of LH_BF2 in CH2Cl2-d2.

Figure 8.9: 13C-NMR of LH_BF2 in CH2Cl2-d2.

138
Figure 8.10: 11B-NMR of LH_BF2 in CH2Cl2-d2.

Figure 8.11: 19F-NMR of LH_BF2 in CH2Cl2-d2.

139
Figure 8.12: 1H-NMR of LMe_BF2 in CH2Cl2-d2.

Figure 8.13: 13C-NMR of LMe_BF2 in CH2Cl2-d2.

140
Figure 8.14: 11B-NMR of LMe_BF2 in CH2Cl2-d2.

Figure 8.15: 19F-NMR of LMe_BF2 in CH2Cl2-d2.

141
Figure 8.16: 1H-NMR of LOMe_BF2 in CH2Cl2-d2.

Figure 8.17: 13C-NMR of LOMe_BF2 in CH2Cl2-d2.

142
Figure 8.18: 11B-NMR of LOMe_BF2 in CH2Cl2-d2.

Figure 8.19: 19F-NMR of LOMe_BF2 in CH2Cl2-d2.

143
Figure 8.20: 1H-NMR of LF_BF2 in CH2Cl2-d2.

Figure 8.21: 13C-NMR of LF_BF2 in CH2Cl2-d2.

144
Figure 8.22: 11B-NMR of LF_BF2 in CH2Cl2-d2.

Figure 8.23: 19F-NMR of LF_BF2 in CH2Cl2-d2.

145
Figure 8.24: 1H-NMR of LCl_BF2 in CH2Cl2-d2.

Figure 8.25: 13C-NMR of LCl_BF2 in CH2Cl2-d2.

146
Figure 8.26: 11B-NMR of LCl_BF2 in CH2Cl2-d2.

Figure 8.27: 19F-NMR of LCl_BF2 in CH2Cl2-d2.

147
Figure 8.28: 1H-NMR of LBr_BF2 in CH2Cl2-d2.

Figure 8.29: 13C-NMR of LBr_BF2 in CH2Cl2-d2.

148
Figure 8.30: 11B-NMR of LBr_BF2 in CH2Cl2-d2.

Figure 8.31: 19F-NMR of LBr_BF2 in CH2Cl2-d2.

149
Figure 8.32: 1H-NMR of LI_BF2 in CH2Cl2-d2.

Figure 8.33: 13C-NMR of LI_BF2 in CH2Cl2-d2.

150
Figure 8.34: 11B-NMR of LI_BF2 in CH2Cl2-d2.

Figure 8.35: 19F-NMR of LI_BF2 in CH2Cl2-d2.

151
Figure 8.36: 1H-NMR of LNO2_BF2 in CH2Cl2-d2.

Figure 8.37: 13C-NMR of LNO2_BF2 in CH2Cl2-d2.

152
Figure 8.38: 11B-NMR of LNO2_BF2 in CH2Cl2-d2.

Figure 8.39: 19F-NMR of LNO2_BF2 in CH2Cl2-d2.

153
8.5 SYNTHESIS OF hydrHLR LIGANDS

In a Parr steel autoclave previously purged with argon, methanol (30 mL) was thoroughly
deoxygenated with argon, then HLR (ca. 1 g) and 10% Pd/C (0.1 g) were added and the suspension
further deoxygenated with argon for other 30 min. Then the autoclave was carefully charged with
14 atm of hydrogen and the system was maintained under stirring at room temperature for 72 h.
The mixture was filtered over a pad of celite to remove Pd/C, the solution was rotary evaporated
to dryness and the crude product was crystallized with diethyl ether.

8.6 SYNTHESIS OF hydrLR_BF2 COMPOUNDS

hydrHLR (1 equivalent) is dissolved in 6 mL of deoxygenated dichloromethane, then, boron


trifluoride diethyletherate (2.5 equivalents) is added to the solution. Finally triethylamine (1.2
equivalents) is added to the reaction mixture. The reaction is stirred for 12 hours, then the solvent
is reduced to half the volume and the solid formed is filtered. The crude product is redissolved in a
minimum amount of dichloromethane and the solutions is filtered over a silica pad. The filtrate is
finally concentrated to dryness under vacuum, giving the final product.

8.6.1 hydrLH_BF2
Yield: 75.0 %. Elemental analysis (%) calcd. C14H15BF2N2O: C 60.90, H 5.48, N 10.14. Found: C 60.87, H 5.52, N
10.30. 1H-NMR (400 MHz, CD2Cl2, 25 °C): δ = 7.76 (dd, J = 8.1, 1.6 Hz, 1H), 7.45 – 7.40 (m, 1H), 7.16 (dd, J =
8.3, 1.2 Hz, 1H), 7.04 – 6.99 (m, 1H), 4.38 (t, J = 6.2 Hz, 2H), 2.87 – 2.78 (m, 2H), 2.34 (s, J = 1.1 Hz, 3H), 2.17 –
2.06 (m, 2H), 2.00 – 1.89 (m, 2H). 13C-NMR (100 MHz, CD2Cl2, 25 °C): δ = 131.85, 123.76, 119.82, 119.12, 46.70,
22.97, 20.51, 18.69, 8.90. 19F-NMR (376 MHz, CD2Cl2, 25°C): δ = -139.09 (q, J = 14.9 Hz). 11B NMR (128 MHz,
CD2Cl2, 25°C): δ = 0.78 (t, J = 15.4 Hz). IR (nujol mull) ṽ = 1032 – 1162 cm-1 (BF2).

8.6.2 hydrLMe_BF2
Yield 78.5 %. Elemental analysis (%) calcd. for C15H17BF2N2O: C 62.10, H 5.91 N 9.66. Found: C 62.11, H 6.03,
N, 9.72. 1H-NMR (400 MHz, CD2Cl2, 25 °C): δ = 7.54 (d, J = 2.1 Hz, 1H), 7.25 (dd, J = 8.3, 2.1 Hz, 1H), 7.05 (d, J
= 8.4 Hz, 1H), 4.39 (t, J = 6.2 Hz, 2H), 2.83 (m, 2H), 2.33 (s, 1H), 2.17 – 2.06 (m, 2H), 2.00 – 1.89 (m, 2H). 13C-
NMR (100 MHz CD2Cl2, 25 °C): δ = 153.62, 132.76, 128.43, 126.03, 125.15, 123.71, 119.50, 110.25, 46.74,
22.98, 20.57, 18.73, 8.91. 19F-NMR (376.5 MHz, CD2Cl2, 25 °C): δ = -139.36 (q, J = 15.1 Hz). 11B-NMR (128 MHz,
CD2Cl2, 25°C) δ 0.81 (t, J = 15.4 Hz). IR (nujol mull) ṽ = 1034 – 1168 cm-1 (BF2).

154
8.6.3 hydrLOMe_BF2
Yield 74.4 %. Elemental analysis (%) calcd. for C15H17BF2N2O2: C 58.85, H 5.60, N 9.15. Found: C 58.23, H 5.63,
N 9.05. 1H-NMR (400 MHz, CD2Cl2, 25°C): δ = 7.23 (d, J = 2.8 Hz, 1H), 7.06 (d, J = 9.0 Hz, 1H), 7.00 (dd, J = 9.0,
2.9 Hz, 1H), 4.35 (t, J = 6.2 Hz, 2H), 3.80 (s, 3H), 2.79 (t, J = 6.3 Hz, 2H), 2.29 (s, 3H), 2.16 – 1.96 (m, 2H), 1.96
– 1.82 (m, 2H). 13C-NMR (100 MHz, CD2Cl2, 25°C): δ = 152.24, 120.26, 117.68, 114.99, 109.11, 55.97, 46.57,
22.94, 20.52, 18.70, 8.92. 19F-NMR (376 MHz, CD2Cl2, 25°C): δ = -139.63 (q, J = 14.9 Hz). 11B-NMR (128 MHz,
CD2Cl2, 25°C): δ = 0.84 (t, J = 15.9 Hz). IR (nujol mull) ṽ = 1038 - 1103 cm-1 (BF2).

8.6.4 hydrLF_BF2
Yield 65.1 %. Elemental analysis (%) calcd. for C14H14BF3N2O: C 57.18, H 4.80, N 9.53. Found: C 57.17, H 4.94,
N 9.27. 1H-NMR (400 MHz, CD2Cl2, 25°C): δ = 7.42 (dd, J = 9.6, 2.7 Hz, 1H), 7.20 – 6.99 (m, 2H), 4.32 (t, J = 6.2
Hz, 2H), 2.79 (t, J = 6.5, 2H), 2.30 (s, 3H), 2.17 – 2.02 (m, 2H), 2.03 – 1.76 (m, 2H). 13C-NMR (100 MHz, CD2Cl2,
25°C): δ = 156.60, 154.25, 151.95, 126.75, 125.67, 120.93, 118.81, 109.55, 46.51, 22.83, 20.51, 18.60, 8.90.
19
F-NMR (376 MHz, CD2Cl2, 25°C): δ = -124.51 (q, J = 8.4 Hz), -139.14 (q, J = 14.5 Hz). 11B-NMR (128 MHz,
CD2Cl2, 25°C): δ = 0.79 (t, J = 15.3 Hz). IR (nujol mull) ṽ = 1037 - 1168 cm-1 (BF2).

8.6.5 hydrLOH_BF2
Yield 87 %. Elemental analysis (%) calcd. for C14H15BF2N2O2: C 57.57, H 5.18, N 9.59. Found: C 60.65, H 5.21, N
9.62. 1H-NMR (400 MHz, acetone-d6, 25 °C): δ = 8.03 (s), 7.35 (d, J = 2.7 Hz, 1H), 7.10 – 6.69 (m, 2H), 4.48 (t, J
= 6.2 Hz, 2H), 2.84 (t, J = 6.6 Hz, 2H), 2.79 (s, 1H), 2.26 (t, J = 1.1 Hz, 3H), 2.21 – 2.11 (m, 2H), 2.00 – 1.87 (m,
2H). 13C-NMR (100 MHz, acetone-d6, 25°C): δ = 149.37, 126.51, 124.48, 119.85, 119.24, 109.79, 46.43, 22.68,
20.21, 18.42, 8.32. 19F-NMR (376.5 MHz, acetone-d6, 25 °C): δ = -138.94 (q, J = 14.2 Hz). 11B-NMR (128 MHz,
acetone-d6, 25 °C): δ = 0.88 (t, J = 15.1 Hz). IR (nujol mull) ṽ = 1032 - 1173 cm-1 (BF2), 3482 cm-1 (OH).

155
Figure 8.40: 1H-NMR of hydrLH_BF2 in CH2Cl2-d2.

Figure 8.41: 13C-NMR of hydrLH_BF2 in CH2Cl2-d2.

156
Figure 8.42: 11B-NMR of hydrLH_BF2 in CH2Cl2-d2.

Figure 8.43: 19F-NMR of hydrLH_BF2 in CH2Cl2-d2.

157
Figure 8.44: 1H-NMR of hydrLMe_BF2 in CH2Cl2-d2.

Figure 8.45: 13C-NMR of hydrLMe_BF2 in CH2Cl2-d2.

158
Figure 8.46: 11B-NMR of hydrLMe_BF2 in CH2Cl2-d2.

Figure 8.47: 19F-NMR of hydrLMe_BF2 in CH2Cl2-d2.

159
Figure 8.48: 1H-NMR of hydrLOMe_BF2 in CH2Cl2-d2.

Figure 8.49: 13C-NMR of hydrLOMe_BF2 in CH2Cl2-d2.

160
Figure 8.50: 11B-NMR of hydrLOMe_BF2 in CH2Cl2-d2.

Figure 8.51: 19F-NMR of hydrLOMe_BF2 in CH2Cl2-d2.

161
Figure 8.52: 1H-NMR of hydrLF_BF2 in CH2Cl2-d2.

Figure 8.53: 13C-NMR of hydrLF_BF2 in CH2Cl2-d2.

162
Figure 8.54: 11B-NMR of hydrLF_BF2 in CH2Cl2-d2.

Figure 8.55: 19F-NMR of hydrLF_BF2 in CH2Cl2-d2.

163
Figure 8.56: 1H-NMR of hydrLOH_BF2 in acetone-d6.

Figure 8.57: 13C-NMR of hydrLOH_BF2 in acetone-d6.

164
Figure 8.58: 11B-NMR of hydrLOH_BF2 in acetone-d6.

Figure 8.59: 19F-NMR of hydrLOH_BF2 in acetone-d6.

165
8.7 CRYSTALLOGRAPHIC DATA

8.7.1 [Zn(LH)2]

[Zn(LH)2]

Formula C28H22N4O2Zn ּ
0.5 C2H6O

Space Group P -1

a = 8.760(2)
Cell Lengths b = 16.655(3)
c = 16.863(3)
α =80.29(3)°
Cell Angles β = 89.72(3)°
γ = 85.88(3)°

Cell Volume 2418.7

Z Z=4

R-Factor (%) 6.28


Figure 8.60: Cristal structure of [Zn(LH)2].
Table 8.1: Crystallographic parameters obtained by X-ray diffraction for [Zn(LH)2].

Length Length Length Length


Atom1 Atom2 Atom1 Atom2 Atom1 Atom2 Atom1 Atom2
(Å) (Å) (Å) (Å)
C1 C2 1.40(1) C17 H17 0.93 C31 C32 1.37(1) C48 C49 1.45(1)
C1 C6 1.42(1) C17 C18 1.38(2) C32 H32 0.93 C49 N6 1.36(1)
C1 O1 1.31(1) C18 H18 0.93 C32 C33 1.36(2) C49 N8 1.38(1)
C2 H2 0.931 C18 C19 1.37(1) C33 H33 0.93 C50 H50 0.93
C2 C3 1.38(1) C19 H19 0.93 C33 C34 1.43(1) C50 C51 1.32(2)
C3 H3 0.93 C19 C20 1.40(1) C34 C35 1.45(1) C50 N8 1.41(1)
C3 C4 1.39(1) C20 C21 1.44(1) C35 N5 1.34(1) C51 H51 0.93
C4 H4 0.93 C21 N1 1.35(1) C35 N7 1.37(1) C51 C52 1.43(2)
C4 C5 1.38(1) C21 N3 1.39(1) C36 H36 0.931 C52 H52 0.93
C5 H5 0.931 C22 H22 0.93 C36 C37 1.35(1) C52 C53 1.36(1)
C5 C6 1.41(1) C22 C23 1.33(1) C36 N7 1.41(1) C53 H53 0.93
C6 C7 1.47(1) C22 N3 1.40(1) C37 H37 0.93 C53 C54 1.39(1)
C7 N2 1.31(1) C23 H23 0.93 C37 C38 1.41(1) C54 C55 1.40(1)

166
Length Length Length Length
Atom1 Atom2 Atom1 Atom2 Atom1 Atom2 Atom1 Atom2
(Å) (Å) (Å) (Å)
C7 N4 1.40(1) C23 C24 1.42(2) C38 H38 0.93 C54 N8 1.42(1)
C8 H8 0.93 C24 H24 0.93 C38 C39 1.36(1) C55 C56 1.50(1)
C8 C9 1.37(1) C24 C25 1.34(1) C39 H39 0.93 C55 N6 1.36(1)
C8 N4 1.39(1) C25 H25 0.93 C39 C40 1.39(1) C56 H56A 0.96
C9 H9 0.93 C25 C26 1.44(1) C40 C41 1.41(1) C56 H56B 0.96
C9 C10 1.42(2) C26 C27 1.37(1) C40 N7 1.41(1) C56 H56C 0.96
C10 H10 0.93 C26 N3 1.39(1) C41 C42 1.49(1) N5 Zn2 1.991(7)
C10 C11 1.35(2) C27 C28 1.48(1) C41 N5 1.35(1) N6 Zn2 1.965(8)
C11 H11 0.93 C27 N1 1.37(1) C42 H42A 0.96 O3 Zn2 1.927(6)
C11 C12 1.43(1) C28 H28A 0.959 C42 H42B 0.96 O4 Zn2 1.916(6)
C12 C13 1.35(1) C28 H28B 0.96 C42 H42C 0.96 C57 H57A 0.97
C12 N4 1.42(1) C28 H28C 0.96 C43 C44 1.40(1) C57 H57B 0.97
C13 C14 1.49(1) N1 Zn1 1.982(7) C43 C48 1.42(1) C57 C58 1.44(2)
C13 N2 1.39(1) N2 Zn1 1.996(7) C43 O4 1.32(1) C57 O5 1.44(1)
C14 H14A 0.96 O1 Zn1 1.905(6) C44 H44 0.93 C58 H58A 0.96
C14 H14B 0.96 O2 Zn1 1.903(7) C44 C45 1.39(1) C58 H58B 0.96
C14 H14C 0.96 C29 C30 1.41(1) C45 H45 0.93 C58 H58C 0.96
C15 C16 1.40(1) C29 C34 1.40(1) C45 C46 1.39(2)
C15 C20 1.42(1) C29 O3 1.36(1) C46 H46 0.93
C15 O2 1.34(1) C30 H30 0.93 C46 C47 1.38(2)
C16 H16 0.93 C30 C31 1.36(1) C47 H47 0.93
C16 C17 1.36(2) C31 H31 0.93 C47 C48 1.38(1)
Table 8.2: Bond lengths obtained by X-ray diffraction for [Zn(LH)2].

Atom1 Atom2 Atom3 Angle (°) Atom1 Atom2 Atom3 Angle (°) Atom1 Atom2 Atom3 Angle (°)
C2 C1 C6 116.1(8) C26 C27 C28 130.1(9) C44 C45 C46 121(1)
C2 C1 O1 120.1(8) C26 C27 N1 107.8(8) H45 C45 C46 119
C6 C1 O1 123.8(8) C28 C27 N1 122.1(8) C45 C46 H46 122
C1 C2 H2 118.5 C27 C28 H28A 109.5 C45 C46 C47 116(1)
C1 C2 C3 122.9(9) C27 C28 H28B 109.4 H46 C46 C47 122
H2 C2 C3 118.5 C27 C28 H28C 109.5 C46 C47 H47 118
C2 C3 H3 120 H28A C28 H28B 110 C46 C47 C48 124.7(9)
C2 C3 C4 119.8(9) H28A C28 H28C 110 H47 C47 C48 117.7
H3 C3 C4 120 H28B C28 H28C 109 C43 C48 C47 118.9(8)
C3 C4 H4 120 C21 N1 C27 110.3(7) C43 C48 C49 120.0(8)
C3 C4 C5 120.0(9) C21 N1 Zn1 116.0(6) C47 C48 C49 121.1(8)
H4 C4 C5 120 C27 N1 Zn1 127.5(6) C48 C49 N6 125.7(8)
C4 C5 H5 119.8 C7 N2 C13 110.6(7) C48 C49 N8 126.8(8)
C4 C5 C6 120.2(9) C7 N2 Zn1 115.8(6) N6 C49 N8 107.5(7)
H5 C5 C6 119.9 C13 N2 Zn1 125.6(6) H50 C50 C51 121
C1 C6 C5 120.9(8) C21 N3 C22 128.4(8) H50 C50 N8 120.4
C1 C6 C7 119.3(8) C21 N3 C26 108.8(7) C51 C50 N8 119.1(9)

167
Atom1 Atom2 Atom3 Angle (°) Atom1 Atom2 Atom3 Angle (°) Atom1 Atom2 Atom3 Angle (°)
C5 C6 C7 119.8(8) C22 N3 C26 122.3(8) C50 C51 H51 119
C6 C7 N2 128.6(8) C7 N4 C8 132.4(7) C50 C51 C52 122(1)
C6 C7 N4 123.2(8) C7 N4 C12 106.2(7) H51 C51 C52 119
N2 C7 N4 108.2(8) C8 N4 C12 121.0(7) C51 C52 H52 120
H8 C8 C9 120.1 C1 O1 Zn1 124.9(6) C51 C52 C53 119(1)
H8 C8 N4 120.1 C15 O2 Zn1 122.8(6) H52 C52 C53 120
C9 C8 N4 119.9(8) N1 Zn1 N2 127.4(3) C52 C53 H53 120
C8 C9 H9 121 N1 Zn1 O1 113.4(3) C52 C53 C54 119.8(9)
C8 C9 C10 118.6(9) N1 Zn1 O2 97.4(3) H53 C53 C54 120
H9 C9 C10 121 N2 Zn1 O1 96.6(3) C53 C54 C55 135.4(8)
C9 C10 H10 118 N2 Zn1 O2 109.0(3) C53 C54 N8 119.9(8)
C9 C10 C11 124(1) O1 Zn1 O2 113.7(3) C55 C54 N8 104.7(7)
H10 C10 C11 118 C30 C29 C34 119.0(8) C54 C55 C56 127.5(8)
C10 C11 H11 121 C30 C29 O3 116.9(8) C54 C55 N6 109.4(8)
C10 C11 C12 117.8(9) C34 C29 O3 124.1(8) C56 C55 N6 123.0(8)
H11 C11 C12 121 C29 C30 H30 119.7 C55 C56 H56A 109.5
C11 C12 C13 133.7(9) C29 C30 C31 120.6(9) C55 C56 H56B 109.5
C11 C12 N4 118.8(8) H30 C30 C31 120 C55 C56 H56C 109.5
C13 C12 N4 107.5(7) C30 C31 H31 119 H56A C56 H56B 109.5
C12 C13 C14 130.9(8) C30 C31 C32 122(1) H56A C56 H56C 109.5
C12 C13 N2 107.5(7) H31 C31 C32 119 H56B C56 H56C 109.5
C14 C13 N2 121.5(8) C31 C32 H32 121 C35 N5 C41 109.9(7)
C13 C14 H14A 109.4 C31 C32 C33 119(1) C35 N5 Zn2 119.1(6)
C13 C14 H14B 109.5 H32 C32 C33 121 C41 N5 Zn2 127.4(6)
C13 C14 H14C 109.4 C32 C33 H33 119 C49 N6 C55 109.5(7)
H14A C14 H14B 109.5 C32 C33 C34 121.8(9) C49 N6 Zn2 119.3(6)
H14A C14 H14C 109.4 H33 C33 C34 119.1 C55 N6 Zn2 128.3(6)
H14B C14 H14C 109.6 C29 C34 C33 117.8(8) C35 N7 C36 130.1(8)
C16 C15 C20 116.4(8) C29 C34 C35 121.7(8) C35 N7 C40 109.5(7)
C16 C15 O2 120.8(8) C33 C34 C35 120.0(8) C36 N7 C40 120.3(7)
C20 C15 O2 122.8(8) C34 C35 N5 124.3(8) C49 N8 C50 131.5(7)
C15 C16 H16 118.5 C34 C35 N7 127.5(8) C49 N8 C54 108.8(7)
C15 C16 C17 122.8(9) N5 C35 N7 107.7(8) C50 N8 C54 119.5(7)
H16 C16 C17 119 H36 C36 C37 121.6 C29 O3 Zn2 118.3(5)
C16 C17 H17 119 H36 C36 N7 121.7 C43 O4 Zn2 119.8(6)
C16 C17 C18 121(1) C37 C36 N7 116.7(8) N5 Zn2 N6 114.7(3)
H17 C17 C18 119 C36 C37 H37 118 N5 Zn2 O3 95.7(3)
C17 C18 H18 121 C36 C37 C38 124.0(9) N5 Zn2 O4 118.1(3)
C17 C18 C19 118.3(9) H37 C37 C38 118 N6 Zn2 O3 118.6(3)
H18 C18 C19 121 C37 C38 H38 120.7 N6 Zn2 O4 95.4(3)
C18 C19 H19 118.9 C37 C38 C39 118.5(9) O3 Zn2 O4 116.1(3)
C18 C19 C20 122.2(9) H38 C38 C39 120.8 H57A C57 H57B 108
H19 C19 C20 118.9 C38 C39 H39 120 H57A C57 C58 109
C15 C20 C19 119.2(8) C38 C39 C40 119.9(9) H57A C57 O5 109
C15 C20 C21 121.1(8) H39 C39 C40 120 H57B C57 C58 109

168
Atom1 Atom2 Atom3 Angle (°) Atom1 Atom2 Atom3 Angle (°) Atom1 Atom2 Atom3 Angle (°)
C19 C20 C21 119.7(8) C39 C40 C41 135.6(9) H57B C57 O5 109
C20 C21 N1 127.1(8) C39 C40 N7 120.5(8) C58 C57 O5 114(1)
C20 C21 N3 126.5(8) C41 C40 N7 103.9(7) C57 C58 H58A 109
N1 C21 N3 106.4(7) C40 C41 C42 129.4(8) C57 C58 H58B 109
H22 C22 C23 121 C40 C41 N5 109.0(8) C57 C58 H58C 109
H22 C22 N3 121.2 C42 C41 N5 121.6(8) H58A C58 H58B 109
C23 C22 N3 117.6(9) C41 C42 H42A 109.5 H58A C58 H58C 110
C22 C23 H23 119 C41 C42 H42B 109.4 H58B C58 H58C 110
C22 C23 C24 122(1) C41 C42 H42C 109.4
H23 C23 C24 119 H42A C42 H42B 109
C23 C24 H24 119 H42A C42 H42C 110
C23 C24 C25 121(1) H42B C42 H42C 109
H24 C24 C25 119 C44 C43 C48 116.9(9)
C24 C25 H25 121 C44 C43 O4 117.1(8)
C24 C25 C26 118.1(9) C48 C43 O4 126.0(8)
H25 C25 C26 120.9 C43 C44 H44 119
C25 C26 C27 134.9(9) C43 C44 C45 121.8(9)
C25 C26 N3 118.4(8) H44 C44 C45 119
C27 C26 N3 106.7(8) C44 C45 H45 119
Table 8.3: Angles obtained by X-ray diffraction for [Zn(LH)2].

8.7.2 LH_BF2

LH_BF2

Formula C14H11BF2N2O

Space Group P mcn

a = 6.8839(5)
Cell Lengths b = 10.0250(7)
c = 17.8687(10)
α = 90°
Cell Angles β = 90°
γ = 90°

Cell Volume 1233.14

Figure 8.61: Cristal structure of LH_BF2.


Z Z=4

R-Factor (%) 6.64


Table 8.4: Crystallographic parameters obtained by X-ray diffraction for LH_BF2.

169
Length Length
Atom1 Atom2 Atom1 Atom2
(Å) (Å)
C1 C2 1.383 C9 H9 0.68
C1 C6 1.395 C9 C10 1.411
C1 O1 1.404 C10 H10 0.98
C1 O1 1.404 C10 C11 1.327
C2 H2 0.96 C11 H11 0.96
C2 C3 1.364 C11 C12 1.422
C3 H3 1.15 C12 C13 1.35
C3 C4 1.345 C12 N2 1.408
C4 H4 1 C13 C14 1.483
C4 C5 1.386 C13 N1 1.369
C5 H5 1.02 C14 H13 0.9
C5 C6 1.358 C14 H12 1.17
C6 C7 1.448 C14 H13 0.9
C7 N1 1.332 B1 F1 1.337
C7 N2 1.371 B1 N1 1.578
C8 H8 1.05 B1 O1 1.443
C8 C9 1.338 B1 F1 1.337
C8 N2 1.372 B1 O1 1.443
Table 8.5: Bond lengths obtained by X-ray diffraction for LH_BF2.

Angle Angle Angle


Atom1 Atom2 Atom3 Atom1 Atom2 Atom3 Atom1 Atom2 Atom3
(°) (°) (°)
C2 C1 C6 120 H8 C8 C9 122 F1 B1 N1 109.3
C2 C1 O1 117.1 H8 C8 N2 118 F1 B1 O1 96.2
C2 C1 O1 117.1 C9 C8 N2 119.5 F1 B1 F1 106.9
C6 C1 O1 119.9 C8 C9 H9 116 F1 B1 O1 125.2
C6 C1 O1 119.9 C8 C9 C10 122.4 N1 B1 O1 108.4
O1 C1 O1 34.9 H9 C9 C10 122 N1 B1 F1 109.3
C1 C2 H2 126 C9 C10 H10 117 N1 B1 O1 108.4
C1 C2 C3 120.9 C9 C10 C11 119.1 O1 B1 F1 125.2
H2 C2 C3 113 H10 C10 C11 124 F1 B1 O1 96.2
C2 C3 H3 116 C10 C11 H11 123 C7 N1 C13 110.6
C2 C3 C4 119.6 C10 C11 C12 120.4 C7 N1 B1 121.5
H3 C3 C4 125 H11 C11 C12 116 C13 N1 B1 127.9
C3 C4 H4 119 C11 C12 C13 133.8 C7 N2 C8 132.8
C3 C4 C5 120 C11 C12 N2 118.7 C7 N2 C12 107.3
H4 C4 C5 121 C13 C12 N2 107.4 C8 N2 C12 119.8
C4 C5 H5 113 C12 C13 C14 128 C1 O1 B1 119.2
C4 C5 C6 122.1 C12 C13 N1 107.3 C1 O1 O1 72.5
H5 C5 C6 125 C14 C13 N1 124.7 B1 O1 O1 73
C1 C6 C5 117.4 C13 C14 H13 112 C1 O1 B1 119.2
C1 C6 C7 116 C13 C14 H12 112 C1 O1 O1 72.5
C5 C6 C7 126.6 C13 C14 H13 112 B1 O1 O1 73

170
Angle Angle Angle
Atom1 Atom2 Atom3 Atom1 Atom2 Atom3 Atom1 Atom2 Atom3
(°) (°) (°)
C6 C7 N1 122.6 H13 C14 H12 107
C6 C7 N2 130 H13 C14 H13 106
N1 C7 N2 107.4 H12 C14 H13 107
Table 8.6: Angles obtained by X-ray diffraction for LH_BF2.

8.7.3 LOMe_BF2

LOMe_BF2

Formula C15H13BF2N2O2

Space Group P 21/c

a = 17.9831(13)
Cell Lengths b = 7.1160(4)
c = 22.4770(17)
α = 90°
Cell Angles β = 111.246(6)°
γ = 90°

Cell Volume 2680.84

Z Z=8

R-Factor (%) 11.73

Figure 8.62: Cristal structure of LOMe_BF2.


Table 8.7: Crystallographic parameters obtained by X-ray diffraction for LOMe_BF2.

Length Length Length


Atom1 Atom2 Atom1 Atom2 Atom1 Atom2
(Å) (Å) (Å)
C1 C2 1.39(1) C13 N1 1.394(9) C28 H28 0.931
C1 C6 1.411(9) C14 H14A 0.96 C28 C29 1.35(1)
C1 O1 1.369(6) C14 H14B 0.958 C28 N4 1.394(7)
C2 H2 0.93 C14 H14C 0.96 C29 H29 0.929
C2 C3 1.359(8) C15 H15A 0.96 C29 C30 1.429(9)
C3 H3 0.931 C15 H15B 0.96 C30 H30 0.932
C3 C4 1.41(1) C15 H15C 0.961 C30 C31 1.355(9)

171
Length Length Length
Atom1 Atom2 Atom1 Atom2 Atom1 Atom2
(Å) (Å) (Å)
C4 C5 1.38(1) C15 O2 1.441(9) C31 H31 0.931
C4 O2 1.384(7) B1 F1 1.393(9) C31 C32 1.41(1)
C5 H5 0.93 B1 F2 1.371(7) C32 C33 1.403(8)
C5 C6 1.422(8) B1 N1 1.58(1) C32 N4 1.395(8)
C6 C7 1.45(1) B1 O1 1.448(9) C33 C34 1.475(8)
C7 N1 1.359(6) C21 C22 1.38(1) C33 N3 1.38(1)
C7 N2 1.378(9) C21 C26 1.43(1) C34 H34A 0.96
C8 H8 0.929 C21 O3 1.367(7) C34 H34B 0.96
C8 C9 1.33(1) C22 H22 0.93 C34 H34C 0.96
C8 N2 1.407(7) C22 C23 1.372(8) C35 H35A 0.96
C9 H9 0.93 C23 H23 0.932 C35 H35B 0.96
C9 C10 1.425(9) C23 C24 1.40(1) C35 H35C 0.96
C10 H10 0.932 C24 C25 1.39(1) C35 O4 1.429(9)
C10 C11 1.356(9) C24 O4 1.384(7) B2 F3 1.396(7)
C11 H11 0.93 C25 H25 0.929 B2 F4 1.381(8)
C11 C12 1.40(1) C25 C26 1.401(8) B2 N3 1.580(9)
C12 C13 1.394(9) C26 C27 1.436(9) B2 O3 1.440(9)
C12 N2 1.396(8) C27 N3 1.363(6)
C13 C14 1.496(8) C27 N4 1.363(8)
Table 8.8: Bond lengths obtained by X-ray diffraction for LOMe_BF2.

Angle Angle Angle


Atom1 Atom2 Atom3 Atom1 Atom2 Atom3 Atom1 Atom2 Atom3
(°) (°) (°)
C2 C1 C6 121.0(6) H15A C15 H15B 109.5 C28 C29 C30 122.7(6)
C2 C1 O1 119.1(6) H15A C15 H15C 109.3 H29 C29 C30 118.7
C6 C1 O1 119.9(5) H15A C15 O2 109.5 C29 C30 H30 120.3
C1 C2 H2 119.8 H15B C15 H15C 109.5 C29 C30 C31 119.4(6)
C1 C2 C3 120.3(6) H15B C15 O2 109.6 H30 C30 C31 120.3
H2 C2 C3 119.9 H15C C15 O2 109.5 C30 C31 H31 120.7
C2 C3 H3 119.8 F1 B1 F2 110.3(5) C30 C31 C32 118.6(6)
C2 C3 C4 120.4(6) F1 B1 N1 108.3(5) H31 C31 C32 120.7
H3 C3 C4 119.8 F1 B1 O1 111.3(5) C31 C32 C33 132.5(6)
C3 C4 C5 120.2(6) F2 B1 N1 109.8(5) C31 C32 N4 120.8(6)
C3 C4 O2 115.6(6) F2 B1 O1 109.7(5) C33 C32 N4 106.8(5)
C5 C4 O2 124.2(6) N1 B1 O1 107.4(5) C32 C33 C34 129.1(6)
C4 C5 H5 120 C7 N1 C13 110.2(5) C32 C33 N3 106.4(6)
C4 C5 C6 120.1(6) C7 N1 B1 119.2(5) C34 C33 N3 124.5(6)
H5 C5 C6 119.9 C13 N1 B1 130.4(5) C33 C34 H34A 109.5
C1 C6 C5 117.8(6) C7 N2 C8 130.1(5) C33 C34 H34B 109.6
C1 C6 C7 116.7(5) C7 N2 C12 109.2(5) C33 C34 H34C 109.4
C5 C6 C7 125.4(6) C8 N2 C12 120.3(5) H34A C34 H34B 109.4
C6 C7 N1 121.3(5) C1 O1 B1 119.8(5) H34A C34 H34C 109.4
C6 C7 N2 131.7(6) C4 O2 C15 118.4(5) H34B C34 H34C 109.5
N1 C7 N2 107.0(5) C22 C21 C26 119.8(6) H35A C35 H35B 109.5

172
Angle Angle Angle
Atom1 Atom2 Atom3 Atom1 Atom2 Atom3 Atom1 Atom2 Atom3
(°) (°) (°)
H8 C8 C9 120.9 C22 C21 O3 119.0(6) H35A C35 H35C 109.4
H8 C8 N2 120.9 C26 C21 O3 121.2(6) H35A C35 O4 109.5
C9 C8 N2 118.2(6) C21 C22 H22 119.4 H35B C35 H35C 109.4
C8 C9 H9 118.8 C21 C22 C23 121.3(6) H35B C35 O4 109.5
C8 C9 C10 122.1(6) H22 C22 C23 119.4 H35C C35 O4 109.5
H9 C9 C10 119.1 C22 C23 H23 120.3 F3 B2 F4 109.6(5)
C9 C10 H10 119.9 C22 C23 C24 119.6(6) F3 B2 N3 109.4(5)
C9 C10 C11 120.3(6) H23 C23 C24 120.2 F3 B2 O3 108.7(5)
H10 C10 C11 119.8 C23 C24 C25 120.5(6) F4 B2 N3 109.2(5)
C10 C11 H11 120.7 C23 C24 O4 114.7(6) F4 B2 O3 112.6(5)
C10 C11 C12 118.5(6) C25 C24 O4 124.8(6) N3 B2 O3 107.3(5)
H11 C11 C12 120.8 C24 C25 H25 119.9 C27 N3 C33 110.4(5)
C11 C12 C13 132.5(6) C24 C25 C26 120.1(6) C27 N3 B2 120.3(5)
C11 C12 N2 120.4(6) H25 C25 C26 120 C33 N3 B2 129.2(5)
C13 C12 N2 107.1(5) C21 C26 C25 118.7(6) C27 N4 C28 130.0(5)
C12 C13 C14 129.0(6) C21 C26 C27 115.6(5) C27 N4 C32 109.2(5)
C12 C13 N1 106.5(6) C25 C26 C27 125.7(6) C28 N4 C32 120.5(5)
C14 C13 N1 124.5(6) C26 C27 N3 121.5(5) C21 O3 B2 120.6(5)
C13 C14 H14A 109.5 C26 C27 N4 131.3(5) C24 O4 C35 118.3(5)
C13 C14 H14B 109.5 N3 C27 N4 107.2(5)
C13 C14 H14C 109.5 H28 C28 C29 121.2
H14A C14 H14B 109.6 H28 C28 N4 121
H14A C14 H14C 109.4 C29 C28 N4 117.8(6)
H14B C14 H14C 109.3 C28 C29 H29 118.6
Table 8.9: Angles obtained by X-ray diffraction for LOMe_BF2.

173
8.7.4 hydrLOMe_BF2

hydrLOMe_BF2

Formula C15H17BF2N2O2

Space Group P 21/n

a = 7.5790(10)
Cell Lengths b = 10.0958(11)
c = 18.603(3)
α = 90°
Cell Angles β = 100.816(12)°
γ = 90°

Cell Volume 1398.14

Z Z=4
Figure 8.63: Cristal structure of hydrLOMe_BF2.

R-Factor (%) 6.55


Table 8.10: Crystallographic parameters obtained by X-ray diffraction for hydrLOMe_BF2.

Length Length
Atom1 Atom2 Atom1 Atom2
(Å) (Å)
C1 C2 1.382(6) C10 H10A 0.97
C1 C6 1.417(4) C10 H10B 0.97
C1 O1 1.348(5) C10 C11 1.518(7)
C2 H2 0.931 C11 H11A 0.97
C2 C3 1.370(7) C11 H11B 0.971
C3 H3 0.931 C11 C12 1.490(5)
C3 C4 1.384(6) C12 C13 1.344(5)
C4 C5 1.382(5) C12 N2 1.389(5)
C4 O2 1.376(5) C13 C14 1.480(5)
C5 H5 0.93 C13 N1 1.388(5)
C5 C6 1.390(5) C14 H14A 0.96
C6 C7 1.452(5) C14 H14B 0.96
C7 N1 1.336(5) C14 H14C 0.961
C7 N2 1.352(4) C15 H15A 0.96
C8 H8A 0.97 C15 H15B 0.961
C8 H8B 0.97 C15 H15C 0.96
C8 C9 1.518(6) C15 O2 1.421(5)
C8 N2 1.468(5) B1 F1 1.384(5)
C9 H9A 0.97 B1 F2 1.390(5)

174
Length Length
Atom1 Atom2 Atom1 Atom2
(Å) (Å)
C9 H9B 0.97 B1 N1 1.561(5)
C9 C10 1.479(7) B1 O1 1.429(5)
Table 8.11: Bond lengths obtained by X-ray diffraction for hydrLOMe_BF2.

Angle Angle
Atom1 Atom2 Atom3 Atom1 Atom2 Atom3 Atom1 Atom2 Atom3 Angle (°)
(°) (°)
C2 C1 C6 118.8(3) C9 C8 N2 110.4(3) C13 C14 H14B 109.5
C2 C1 O1 118.6(3) C8 C9 H9A 109.1 C13 C14 H14C 109.5
C6 C1 O1 122.6(3) C8 C9 H9B 109.1 H14A C14 H14B 109.4
C1 C2 H2 119 C8 C9 C10 112.3(4) H14A C14 H14C 109.4
C1 C2 C3 121.9(4) H9A C9 H9B 107.8 H14B C14 H14C 109.4
H2 C2 C3 119.1 H9A C9 C10 109.2 H15A C15 H15B 109.4
C2 C3 H3 120.2 H9B C9 C10 109.2 H15A C15 H15C 109.5
C2 C3 C4 119.7(4) C9 C10 H10A 109.4 H15A C15 O2 109.5
H3 C3 C4 120.1 C9 C10 H10B 109.5 H15B C15 H15C 109.4
C3 C4 C5 119.7(4) C9 C10 C11 111.0(4) H15B C15 O2 109.5
C3 C4 O2 125.1(4) H10A C10 H10B 108 H15C C15 O2 109.5
C5 C4 O2 115.2(3) H10A C10 C11 109.5 F1 B1 F2 109.2(3)
C4 C5 H5 119.5 H10B C10 C11 109.4 F1 B1 N1 110.3(3)
C4 C5 C6 121.3(3) C10 C11 H11A 109.2 F1 B1 O1 108.8(3)
H5 C5 C6 119.2 C10 C11 H11B 109.2 F2 B1 N1 107.8(3)
C1 C6 C5 118.5(3) C10 C11 C12 112.4(3) F2 B1 O1 111.8(3)
C1 C6 C7 115.9(3) H11A C11 H11B 107.9 N1 B1 O1 108.9(3)
C5 C6 C7 125.6(3) H11A C11 C12 109.1 C7 N1 C13 108.6(3)
C6 C7 N1 121.1(3) H11B C11 C12 109.1 C7 N1 B1 122.5(3)
C6 C7 N2 130.6(3) C11 C12 C13 130.7(3) C13 N1 B1 128.9(3)
N1 C7 N2 108.3(3) C11 C12 N2 121.9(3) C7 N2 C8 128.2(3)
H8A C8 H8B 108.1 C13 C12 N2 107.4(3) C7 N2 C12 108.1(3)
H8A C8 C9 109.6 C12 C13 C14 129.4(4) C8 N2 C12 123.6(3)
H8A C8 N2 109.6 C12 C13 N1 107.6(3) C1 O1 B1 123.3(3)
H8B C8 C9 109.6 C14 C13 N1 122.9(3) C4 O2 C15 116.9(3)
H8B C8 N2 109.6 C13 C14 H14A 109.6
Table 8.12: Angles obtained by X-ray diffraction for hydrLOMe_BF2.

175
8.8 COMPUTATIONAL DETAILS

All calculations were carried out at the density functional (DFT) level of theory with the
ADF2020.102 (LR_BF2), ADF2018.105 (hydrLR_BF2) and ADF2017.113 ([Zn(LR)2]), program
package.153 The PBE functional plus a D3 dispersion correction energy term (PBE-D3)154 was
employed for all calculations and later PBE0 was used. Frequency analyses were performed for all
optimized structures to establish the nature of the stationary points. TF-DFT implemented in the
ADF package was used to determine the excitation energies: the 40 lowest single-singlet excitations
were calculated by using the optimized geometries. For geometry optimization, the C, H, N, O, B
and F atoms were described through TZ2P basis sets (triplet-Slater-type orbitals (STOs) plus two
polarization functions); QZ4P basis set (quadrupole-STO plus four polarization functions) was used
for Zn, Br and I atoms and relativistic effects were included for I atoms using the scalar relativistic
zeroth-order regular approximation (ZORA) formalism.155 The corresponding augmented basis set
was employed in TD-DFT calculations.156 Restricted formalism, no-frozen-core approximation (all-
electron) and no-symmetry constraints were used in all calculations. Solvent effects (CH2Cl2) were
simulated employing the conductor-like continuum solvent model (COSMO)157 as implemented in
ADF suit.

176
Appendix: DFT Calculations

177
PREMISE: THE HARTREE-FOCK METHOD
The energy of a molecular system can be in principle calculated using the Schrödinger
equation:

𝐻𝛹 = 𝐸𝛹 eq.1

where E represents the energy of the system, H the Hamiltonian operator and  the
wavefunction describing the system. The Schrödinger equation is an example of differential
equation whose solution is a function (eigenfunction), which after the application of the
Hamiltonian operator gives the function itself multiplied for a scalar value (eigenvalue, in
this case the energy of the system). The energy E represents the formation energy of the
molecule starting from its fundamental components (electrons and nuclei), initially
considered at an infinite distance.
The Hamiltonian operator includes five terms:
1) electron-electron repulsive interaction;
2) electron-nucleus attractive interaction;
3) nucleus-nucleus attractive interaction;
4) kinetic energy of the electrons;
5) kinetic energy of the nuclei.

Without approximation, because of its high mathematical complexity, the equation cannot
be solved, even for the smallest molecule H2. In particular, all the attractive and repulsive
interactions among the particles cannot be exactly described, since the motion of the
particles is strictly correlated.

BORN-OPPENHEIMER APPROXIMATION
The first approximation (Born-Oppenheimer approximation) regards the nuclear motion.
The nuclei are heavier than the electrons, therefore it is possible to consider the electronic
motion in a system of fixed nuclei, separating the wavefunction in two terms, the electronic
and the nuclear terms, thus limiting the calculations to the electronic wavefunction using
a simplified Hamiltonian operator:

178
1) electron-electron repulsive interaction;
2) electron-nucleus attractive interaction: easier to calculate as the nuclei are considered
to be fixed;
3) nucleus-nucleus attractive interaction: can be considered as a constant potential VN that
can be separated from the operator;
4) kinetic energy of the electrons;
5) kinetic energy of the nuclei: considered 0.

As a consequence, the Schrödinger equation allows the calculation of the electronic energy
of the system:

𝐻 𝛹 =𝐸 𝛹 eq.2

where the total energy of the system is E = Er + VN.


Despite this simplification, the electronic Hamiltonian operator still presents a problem:
the electron-electron interaction, as the electrons are not independent, but they are
correlated one to another, is still very complex.

INDEPENDENT ELECTRONS APPROXIMATION


Not considering the electronic interactions, the Hamiltonian operator can be split into a
sum of n mono-electronic operator, where n is the number of electrons in the system:

𝐻 = ℎ + ℎ +⋅⋅⋅ +ℎ eq.3

hi considers only the kinetic energy and the interaction with the nuclei of a single electron.
The complete elimination of the inter-electronic interactions though, leads to unacceptable
mistakes in the energy evaluation.
A compromise can be used, splitting the Hamiltonian operator again in n monoelectronic
operators (called Fock operators) that include the mean repulsive interaction of the i-th
electron considered with the other n-1 electrons. This is the Hartree-Fock method (HF).

179
In this case the electronic Hamiltonian operator is not the sum of n Fock operators, but to
every monoelectronic operator fi, a monoelectronic wavefunction i, solution of the
monoelectronic Schrodinger equation, can be associated.

𝑓𝜓 = 𝜀 𝜓 eq.4

This wavefunction represents the i-th molecular orbital with energy i.
In the general approach, a term depending on the electronic spin must be considered as
well, in turn obtained from another eigenvalue equation. As a consequence, for every
electron, a spin-orbital (j) can be defined as the product between a  orbital and a spin
eigenfunction s:

𝜒 (𝑗) = 𝜓 (𝑗) ⋅ 𝑠(𝑗) eq.5

Where i represents the energetic state (orbital) where the j electron can be found.
The total wavefunction of the polyelectronic system can be derived from the resolution of
the Slater determinant:

𝜒 (1) 𝜒 (2) 𝜒 (𝑛)


𝜒 (1) 𝜒 (2) 𝜒 (𝑛)
𝛹 = eq.6
√ !
𝜒 (1) 𝜒 (2) 𝜒 (𝑛)

Moreover, the polyelectronic wavefunction must consider the Pauli’s exclusion principle,
thus being antisymmetric.
The spin eigenfunctions can only assume two different forms, indicated with α and β
depending on the two spin states that can be assumed by the electron. Thus, taking the H2
molecule as an example (composed by two paired electrons in one orbital), the
determinant has the following form:

𝜓𝛼(1) 𝜓𝛽(1)
𝛹 = = [𝜓𝛼(1) ⋅ 𝜓𝛽(2) − 𝜓𝛼(2) ⋅ 𝜓𝛽(1)] eq.7
√ 𝜓𝛼(2) 𝜓𝛽(2) √

180
In complex cases (beyond the monoelectronic H2+ molecule), the solution of the Slater
determinant cannot be obtained analytically, but only numerically using an iterative
method called Self-Consistent-Field (SCF).
Using SCF, the starting Fock operators are derived from a series of initial molecular orbitals,
and then, they are used to solve the Slater determinant, finding new function and their
associated energies. This procedure is repeated until the energy difference between two
cycles is below a certain threshold.

THE WAVEFUNCTION
As described, the Hartee-Fock method approximates the polyelectronic Hamiltonian
operator using the Fock operators, in turn used to solve the Slater determinant.
The initial wavefunctions (molecular orbitals) used in this process are usually represented
using a linear combination of atomic orbitals (LCAO method). The mathematical forms of
the atomic orbitals, indicated with , are exponential function with respect to the r distance
of the electrons from the atomic nucleus:

𝜑∝𝑒 eq.8

Therefore, a molecular orbital is represented as the following sum:

𝜓 =∑ 𝑎 𝜑 eq.9

The linear combination is weighted using the am. In the SCF method, the molecular orbitals
are varied modifying these coefficients, leading to the best representation in terms of
energy.
The exponential functions used to describe the atomic orbitals are called Slater Type
Orbitals (STO). From a physical perspective, this kind of orbitals are the best to describe the
dependance on the distance r, but, merely mathematically speaking, STOs can lead to
problems, especially due to some difficulties in the calculation of the exchange integrals. A
possible solution is to describe the STOs as linear combinations of gaussian functions
depending on r, in the form of 𝑒 . This gaussian functions are called “primitive functions”

181
and their combination leads to the STOs. Many commercially available software, like
Gaussian, use this approach.
ADF program (Amsterdam Density Functional) instead, uses the direct approach, using non-
approximated STOs.
In any case, either using STOs or GTOs (Gaussian Type Orbitals), only one atomic orbital can
be descripted by one STO or by a combination of more STOs, which are called “basis
functions”.
For example, if the 1s orbital is described by just one basis function, a SZ (single-ζ, single
zeta) type basis set is obtained. Using GTO notation, the same basis set would be indicated
with STO-3G, being an STO approximated with the combination of three gaussian functions.
Frequently though, the atomic orbital is described by more than one basis functions,
generating the basis set DZ or TZ.
Notably, even if it can be thought that, for example, for the hydrogen a single atomic orbital
(1s) is enough, or five orbitals (1s, 2s, 2px, 2py, 2pz) for the carbon, in order to have a better
description of the energy of the system, it is necessary to use more “movement elasticity”
for the electrons populating the molecular orbitals. In these regards, polarization extra-
functions are added in the calculation (usually p orbitals for the H, or d or f orbitals for
heavier atoms).
A further addition to the basis set are the “diffused functions”, representing in a better way
the final parts of the atomic orbitals in the radial distribution probability diagrams, which
are the more distant area from the nucleus. This diffused bases are extremely important in
the calculation of the electronic transitions.

To sum up:
1) with the appropriate basis set, the atomic orbitals of the elements in the molecules are
approximated;
2) with the atomic orbitals, a first approximation of the molecular orbitals and their
energies is obtained;
3) the process continues with iteration cycles until a constant energy value is obtained;
4) in order to calculate an optimized geometry, the procedure must be repeated, changing
the nuclear coordinates until a minimum in the total energy is achieved.

182
THE DENSITY FUNCTIONAL THEORY (DFT)
As described above, the Hartee-Fock calculation method has the goal the approximate
resolution of the Schrödinger equation. The wavefunction Ψ gives all the information about a
particular state in a system. However, it depends on 3N spatial coordinates, with N = number of
electrons for which a spin variable must be added, therefore being extremely complex. For this
reason, treating chemical and biological systems, often definitely bigger than the simplest
molecules, needs a lot of calculation time.
Using the DFT method, in principle, the complex electronic wavefunction can be substituted by a
simpler function, depending only on three spatial variables and one spin variable: the electronic
density. This is defined as the integral with respect to all the spin and on all the spatial coordinates
of all the electrons, except one (r: [x,y,z,s]):

𝜌 = 𝜌 (𝑟) = 𝑁 ∫ 𝜓 ∗ (𝑟 , . . . , 𝑟 ) 𝜓(𝑟 , . . . , 𝑟 ) d𝑠 𝑑𝑟 , . . . , 𝑑𝑟 eq.10

where ρel(r) represents the electronic density that, if integrated spatially, gives the number
of electrons in the system, while r is the vector (x,y,z):

∫ 𝜌 (𝑟) 𝑑𝑉 = 𝑁 eq.11

Hohenberg and Kohn demonstrated158 in 1964 that the energy of a non-degenerate ground
state of a molecule is determined univocally by its electron density; the energy of the
ground state E0 is therefore a function of ρ(r):

𝐸 = 𝐸[𝜌(𝑟)] eq.12

It must be noted that any test electron density function ρtest has an associated energy
greater than or equal to the true energy of the system E0:

𝐸[𝜌 ]≥ 𝐸 eq.13

183
If the exact density function ρ0 is considered, the true energy can be found. This principle
includes the variation principle of the DFT theory, allowing to calculate the properties of
the ground state without knowing the wavefunction.

DEPENDANCE OF THE ENERGY ON THE ELECTRONIC DENSITY: THE FUNCTIONAL


In order to express the energy as a function of ρ, the same procedure applied for the
Hamiltonian operator can be used: the energy is split into its main terms, the kinetic energy
of the electrons and nuclei, the nuclei-electrons attractions and the inter-electronic and
inter-nuclear repulsions. The Born-Oppenheimer approximation can be applied as well,
considering the nuclear kinetic energy to be zero and the inter-nuclear repulsion to be a
constant.
The electronic energy of the system, Eel, now depends on the electronic kinetic energy, Tel,
on the nucleus-electron attraction, Vnu-el, and on the inter-electronic repulsion, Vel-el:

𝐸 [𝜌] = −𝑇𝑒𝑙 [𝜌] − 𝑉𝑛𝑢−𝑒𝑙 [𝜌] + 𝑉𝑒𝑙−𝑒𝑙 [𝜌] eq.14

Eel[ρ], Tel[ρ], Vnu-el[ρ] and Vel-el[ρ] are density functionals, functions whose argument is the
function ρ(r).
The nuclear-electron interaction can be written using the Coulomb law. Vel-el contains a
coulombian term that can be easily calculated, but also other terms due to the correlation
energy and exchange energy, to which a further corrective term must be added due to the
self-interaction (SID, Self-Interaction Correction) because of the contribution of each
electron to the electronic density with which it interacts.

𝑉 =𝑉 + 𝑉 + 𝑉 + 𝑉 eq.15

Analogously, it can be difficult to express the electronic kinetic energy: if we consider a gas
made of non-interacting electrons, the total kinetic energy can be express as the sum of
the kinetic energies of all the electrons:

𝑇 = ∑ ∫ 𝛻 𝜌 𝑑𝑟 eq.16

184
where ρi represents the electronic density due to the single electron and 2 represents the
Laplace operator. To this term, another term due to the electronic interaction must be added:

𝑇 = 𝑇 +𝑇 eq.17

Concluding, it can be said that the density functional can be express as the sum of four
terms:

𝐸 [𝜌] = −𝑇𝑒𝑙 [𝜌] − 𝑉𝑛𝑢−𝑒𝑙 [𝜌] + 𝑉𝐶𝑜𝑢𝑙𝑜𝑚𝑏 [𝜌] + 𝐸𝑒𝑥𝑐 [𝜌] eq.18

The first three can be accurately known and can be considered as a constant of every
functional, while the fourth term includes all the complex functionals that must be
evaluated:
1) the kinetic correlation energy, Tcorr;
2) the exchange energy, Vexc;
3) the electrostatic correlation energy, Vcorr;
4) the SIC, VSIC.
The exact formulation of the functional Eex is not known, therefore these approximated
terms lead to the formulation of the various functional typologies reported in the literature.
Generally speaking, three different class of functionals are reported:
1) LDA (Local Density Approximation);
2) GGA (Generalized Gradient Approximation) and Meta-GGA;
3) hybrid functionals.
The choice of the approximated functional is a critical step in the DFT calculation, as this
factor is the determining one in terms of how the values obtained depart from the exact
value.
Using LDA functionals, only the electronic density value in a certain point determines the
contribute to the exchange energy, Eex, in that point, while every other non-local effect is
not considered. This model treats the molecules as they have a homogeneous electron gas,
uniformly distributed over a positive charge field. Obviously, even if this representation is
unrealistic, LDA functionals gives a relatively accurate structure, overestimating the
correlation energy and underestimating the exchange energy.

185
GGA approximation adds a term considering the electronic density gradient in a given
point, thus partially considering the non-homogeneous electron density. The model has
been refined considering also the second derivative of the electronic density (meta-GGA).
Finally, the hybrid functionals combine the correlation and exchange DFT functionals
(usually GGA) with the exact HF exchange, evaluated calculating the exchange integrals.
These functionals are the most used. Among these, B3LYP159 and PBE0160 are worth to be
mentioned. The B3LYP functional considers three parameters which describe the
contribute of the exact HF exchange, the local exchange and the exchange and correlation
contribute corrected for the electron density gradient. The PBE0 functional instead,
consider the HF exchange and the correlation and exchange contribute GGA:

𝐸 =𝐸 + 𝑎(𝐸 −𝐸 ) eq.19

These two functionals are equally valid depending on the system considered.

THE DENSITY FUNCTION


In order to describe the density function el, a method similar to the HF method can be applied:
the electronic density is described as a sum of the squared wavefunctions describing the molecular
orbitals of the system:

𝜌 = ∑ |𝜓 | eq.20

where each orbital is described as linear correlation of atomic orbitals:

𝜓 =∑ 𝑎 𝜑 eq.21
𝜌 = 𝑓(𝜑 , 𝜑 , . . . , 𝜑 ) eq.22

The basis set used to model the molecular orbitals is usually the same used in the HF
method.
At this point, starting from the functional used to represent the electronic energy of the
system, mono-electronic operators (Kohn-Sham operators, ki)161 are generated, which are

186
dependent on the electronic density. Using ki, eigenvalue monoelectronic equations are
formulated:

𝑘 (𝜌 )𝜓 = 𝜀 𝜓 eq.23

From equation 23, the molecular orbital and its energy can be calculated using an iterative
method until a convergence is reached.

TIME DEPENDENT DFT (TD-DFT): CALCULATIONS OF THE EXCITED STATES


DFT method can be applied only on the electronic ground state of a system. TD-DFT162
method instead, is based on a more general formulation of the Hohenberg-Kohn theorem,
thus being applicable to all the excited states, regardless of the spin or symmetry.
Computational TD-DFT calculations are therefore used to obtain the electronic transitions
and they are based on the variation of the electronic density of the ground state in response
to an electromagnetic field (thus to a time-dependent external perturbation). They use a
linear approximation: the changes in the electronic density are directly proportional to the
variation of the external field.

Figure A.1: Schematic representation of the SCF in the density functional theory.

To obtain the energies and the oscillator strengths of the electronic transitions, TD-DFT
method considers as the molecules were subject to time-dependent perturbations due to
the electric field of an incident electromagnetic wave. Therefore, using this method, the
electronic density depends not only on the spatial coordinates, but also on the time.

187
An adiabatic approximation is used in this kind of calculations, considering the correlation
and exchange potential Eexc, identical to the one used in DFT calculations of the ground
state, considered to be non-time-dependent. From the electronic density, corresponding
to the minimum of the energy functional, the dynamic polarizability, α(ν), can be
calculated, which depends on the frequency of the incident wave. The dynamic
polarizability can be expressed in terms of electronic transition energies and the
corresponding oscillator strengths (f):


𝛼(𝜈) = ∑
( )
= ∑
( )
eq.24

The sum is done for all the excited states; E0 and En are the energies of the ground state
and of the excited states; ν0 and νn are the corresponding excitation frequencies. The
transition energies are those values for which this function is discontinuous. This occurs
whenever the frequency of the incident light is equal to the frequency of the transition.
The oscillator strength is defined by the following equation:

𝑓 = 3e ℏ
𝜈 |𝜇 | = 3e ℏ
𝐸 |𝜇 | eq.25

where μn is the dipole moment associated to the transition. The intensity of the electronic
transition is therefore proportional to μn and to the transition energy; the oscillator
strength determines the intensity of the absorption band; this latter relation allows the
simulation of absorption spectra.
The prevision of the electronic transitions can be more problematic, because of issues due
to the charge transfer excited states, for which the electron is excited from a donor orbital
to an acceptor orbital localized on a different moiety of the molecule. Because of this local
nature of the transition, the pure functional approximations (LDA and GGA) cannot
describe properly the long-distance hole-electron separation. Using these functionals, the
excited electron is considered as originated in the acceptor region and not in the donor
region. For this reason, the incorrect energies of donor and acceptors orbitals lead to an
underestimation of the charge transfer transition energy. For high energy states, hybrid
functionals usually are more accurate.

188
Furthermore, it is possible that the deviation of the calculated spectrum from the
experimental one is due to neglecting the solvent effect. For many molecules, the
absorption bands are influenced by the solvatochromic effect, which causes a shift
depending on the solvent polarity, as the interaction solvent-molecule can be different
from the ground state to the excited state. In particular, charge transfer transitions are
greatly influenced by this effect, since the band shifts are linked to the extension of the
charge separation and to the variation of the molecular dipole moment. The analysis of the
solvatochromic effect is usually experimentally used to find this kind of transitions.
TD-DFT calculations of the charge transfer transitions can be refined including the solvent
in the calculations themselves. In these regards, in this Thesis the COSMO (Conductor-Like
Continuum Solvent Model) model has been applied, for which the solvent media is
considered as a dielectric continuous: the molecules are positioned in solvent cavities with
the same shape of the solute. Then, the interactions between the solute electrons and the
electrostatic field of the solvent are implemented in the calculation.
In order to have the best accordance between experimental and calculated spectra, it is
fundamentally important to have a calculated geometry the closest possible to the real
one. The energy of a molecule is described by the surface of potential energy (PES). The
geometry optimization is the process in which the equilibrium and excited states
geometries are calculated, in absolute or relative minima of this surface. Methods based
on wavefunctions provide good results for small systems, while describing systems
containing transition metals can be less accurate, as the metal-ligand interaction is not
properly described. Only DFT methods provide accurate results for this kind of systems.

A special thanks to Prof. G.A. Ardizzoia for this very easy-understandable “DFT in a nutshell”
breviary.

189
References

190
1
UN General Assembly, Transforming our world: the 2030 Agenda for Sustainable Development, 21
October 2015.
2
a) R.T. Williams, J.W. Bridges, J. Clin. Path., 1964, 17, 371.
b) J.R. Lakowicz, Principle of Fluorescence Spectroscopy, second edition, 1999, Kluwer
Academic/Plenum Publishers.
3
W. McCall, T.M. Christy, D.A. Pipp, B. Jaster, J. White, J. Goodrich, J. Fontana, S. Doxtader,
Environmental Earth Science, 2018, 77, 374.
4
B. Valeur, M.N. Berberan-Santos, Molecular Fluorescence, second edition, 2012, Wiley-VCH Verlag &
Co.
5
R. Livingston, W.F. Watson, J. McArdle, J. Amer. Chem. Soc., 1949, 71, 1542.
6
A. Rosen, R.T. and Williams, Bull. photoelect. Spectrom. Grp,, 1961, 13, 339.
7
F. Feigl, H.E. Feigl, D. Goldstein, J. Amer. chem. Soc., 1955, 77, 4162.
8
C. Förster, K. Heinze, Chem. Soc. Rev., 2020, 49, 1057.
9
H. Xu, R. Chen, Q. Sun, W. Lai, Q. Su, W. Huang, X. Liu, Chem. Soc. Rev., 2014, 43, 3259-3302.
10
S. Zhang, X. Yang, Y. Numata, L. Han, Energy Environ. Sci., 2013, 6, 1443.
11
M.W. Cooke, M.P. Santoni, B. Hasenknopf, G.S. Hanan, Dalton Trans., 2016, 45, 17850-17858.
12
Y. You, W. Nam, Chem. Soc. Rev., 2012, 41, 7061-7084.
13
J.A.G. Williams, S. Develay, D.L. Rochester, L. Murphy, Coord. Chem. Rev., 2008, 252, 2596-2611.
14
C.B. Larsen, H. van der Salm, G.E. Shillito, N.T. Lukas, K.C. Gordon, Inorg. Chem., 2016, 55, 8446-
8458.
15
C. Wegeberg, O.S. Wenger, JACS Au, 2021, 1, 1860-1876.
16
O.S. Wenger, Chem. Eur. J., 2019, 25, 6043−6052.
17
L.A. Büldt, X. Guo, R. Vogel, A. Prescimone, O.S. Wenger, J. Am. Chem. Soc., 2017, 139, 985−992.
18
C. Wegeberg, D. Häussinger, O.S. Wenger, J. Am. Chem. Soc., 2021, 143, 15800-15811.
19
P. Herr, C. Kerzig, C.B. Larsen, D. Häussinger, O.S. Wenger, Nat. Chem., 2021, 13, 956-962.
20
B.W. Pfennig, M.E. Thompson, A.B. Bocarsly, Organometallics, 1993, 12, 649−655.
21
J.W. Kenney, D.R. Boone, D.R. Striplin, Y.H. Chen, K.B. Hamar, Organometallics, 1993, 12, 3671−3676.
22
C.E. Housecroft, E.C. Constable, Chem. Soc. Rev., 2015, 44, 8386−8398.
23
M. Gernert, U. Müller, M. Haehnel, J. Pflaum, A.A. Steffen, Chem. Eur. J., 2017, 23, 2206−2216.
24
G.A. Ardizzoia, S. Brenna, F. Civati, V. Colombo, A. Sironi, CrystEngComm., 2017, 19, 6020.
25
R. Hamze, J.L. Peltier, D. Sylvinson, M. Jung, J. Cardenas, R. Haiges, M. Soleilhavoup, R. Jazzar, P.I.
Djurovich, G. Bertrand, M.E. Thompson, Science, 2019, 363, 601-606.
26
For representative examples: a) R. Diana, B. Panuzzi, Molecules, 2020, 25, 4984. b) P. Matozzo, A.
Colombo, C. Dragonetti, S. Righetto, D. Roberto, P. Biagini, S. Fantacci, D. Marinotto, Inorganics, 2020,
8, 25. c) A. Gusev, V. Shul’gin, E. Braga, E. Zamnius, M. Kryukova, W. Linert, Dyes and Pigments, 2020,
183, 108626.
27
B.W. D’Andrade, S.R. Forrest, Adv. Mater., 2004, 16, 1585-1598.
28
V.C. Bender, T.B. Marchesan, J.M. Alonso, IEEE Industrial Electronic Magazine, 2015, 9, 6-16.
29
E. Huitema, Information Display, 2012, 28, 6-10.
30
J. George, C.S. Menon, Surface and Coatings Technology, 2000, 132, 45-48
31
For representative examples see: a) C. Adachi, T. Tsutsui, S. Saito, Appl. Phys. Lett., 1989, 55, 1489.
1491. b) S. A. VanSlyke, C. H. Chen, C. W. Tang, Appl. Phys. Lett., 1996, 69, 2160–2162.
32
J. Kalinowski, V. Fattori, M. Cocchi, J.A. Gareth Williams, Coordination Chemistry Reviews, 2011, 255,
2401-2425.
33
C.W. Tang, S. A. VanSlyke, Appl. Phys. Lett., 1987, 51, 913–915.
34
L. Duan, L. Hou, T.W. Lee, J. Qiao, D. Zhang, G. Dong, L. Wang, Y. Qiu, J. Mater. Chem., 2010, 20,
6392-6407.
35
S.R. Tseng, H.F. Meng, K.C. Lee, S.F. Horng, Appl. Phys. Lett., 2008, 93, 153308.
36
J. Kido, C. Ohtaki, K. Hongawa, K. Okuyama, K. Nagai, Jpn. J. Appl. Phys., 1993, 32, 917.
37
S.J. Su, Y. Takahashi, T. Chiba, T. Takeda, J. Kido, Adv. Funct. Mater., 2009, 19, 1260-1267.
38
T.D. Anthopoulos, J.P.J. Markham, E.B. Namdas, I.D.W. Samuel, S.C. Lo, P.L. Burn, Appl. Phys. Lett., 2003,
82, 4824–4826.
39
N. Rehmann, D. Hertel, K. Meerholz, H. Becker, S. Heun, Appl. Phys. Lett., 2007, 91, 103507.
40
J.D. Bower, G.R. Ramage, J. Chem. Soc., 1955, 2834-2837.
41
R. Grigg, P. Kennewell, V. Savic, V, Sridharan, Tetrahedron, 1992, 48, 10423-10430.
42
Z. Hu, J. Hou, J. Liu, W. Yu, J. Chang, Org. Biomol. Chem., 2018, 16, 5653-5660.

191
43
F. Shibahara, R. Sugiura, E. Yamaguchi, A. Kitagawa, T. Murai, J. Org. Chem., 2009, 74, 3566–3568.
44
J. Wang, L. Dyers, R. Mason, P. Amoyaw, X. R. Bu, J. Org.Chem. 2005, 70, 2353–2356.
45
A. Kamal, G. Ramakrishna, M. Janaki Ramaiah, A. Viswanath, A.V. Subba Rao, C. Bagul, D.
Mukhopadyay, S.N.C.V.L. Pushpavalli, M. Pal-Bhadra, Med. Chem. Commun., 2013, 4, 697-703.
46
A. Kamal, A.V. Subba Rao, V.L. Nayak, N.V. Subba Reddy, K. Swapna, G. Ramakrishna, M. Alvala, Org.
Biomol. Chem., 2014, 12, 9864.
47
M.S. Khan, M.H. Baig, S. Ahmad, S.A. Siddiqui, A.K. Srivastava, K.V. Srinavasan, I.A. Ansari, PLoS ONE,
2013, 8, 69982.
48
Y.Q. Ge, F.R. Li, Y.J. Zhang, Y.S. Bi, X.Q. Cao, G.Y. Duan, J.W. Wang, Z.L. Liu, Luminescence, 2014, 29,
293–300.
49
R. Nirogi, A.R. Mohammed, A.K. Shinde, N. Bogaraju, S.R. Gagginapalli, S.R. Ravella, L. Kota, G.
Bhyrapuneni, N.R. Muddana, V. Benade, R.C. Palacharla, P. Jayarajan, R. Subramanina, V.K. Goyal, Eur.
J. Med. Chem., 2015, 103, 289-301.
50
Y. Ge, X. Xing, A. Liu, R. Ji, S. Sehn, X. Cao, Dyes and Pigments, 2017, 146, 136-142.
51
S. Chen, H. Li, P. Hou, Sensors and Actuators B, 2018, 256, 1086–1092.
52
S. Chen, P. Hou, J. Sun, H. Wang, L. Liu, Spectrochimica Acta Part A: Molecular and Biomolecular
Spectroscopy, 2020, 225, 117508.
53
Y. Li, G. Zhang, C. Ma, F. Chen, J. Dong, Y. Ge, Dyes and Pigments, 2021, 191, 109381.
54
X. Zhang, Y. Huang, Q. Zhang, D. Li, Y. Li, Eur. J. Inorg. Chem., 2021, 14, 1349 – 1357.
55
J.T. Hutt, J. Jo, A. Olasz, C.H. Chen, D. Lee, Z.D. Aron, Org. Lett., 2012, 14, 3162-3165.
56
X. Zhang, G.J. Song, X.J. Cao, J.T. Liu, M.Y. Chen, X.Q. Cao, B.X. Zhao, RSC Adv., 2015, 5, 89827-89832.
57
G.J. Song, S.Y. Bai, X. Dai, X.Q. Cao, B.X. Zhao, RSC Adv., 2016, 6, 41317-41322.
58
P. Hou, J. Sun, H. Wang, L. Liu, L. Zou, S. Chen, Sens. Actuators, B, 2020, 304, 127244.
59
A.D. Elliott, Curr. Protoc. Cytom., 2020, 92, 69.
60
W. Hoheisel, W. Jacobsen, B. Lüttge, W. Weiner, Macromol. Mater. Eng., 2001, 286, 663–668.
61
F. Yagishita, C. Nii, Y. Tezuka, A. Tabata, H. Nagamune, N. Uemura, Y. Yoshida, T. Mino, M. Sakamoto,
Y. Kawamura, Asian J. Org. Chem., 2018, 7, 1614 –1619.
62
S. Priyanga, T. Khamrang, M. Velusamy, S. Karthi, B. Ashokkumar, R. Mayilmurugan, Dalton Trans.,
2019, 48,1489-1503.
63
K. Hoshi, M. Itaya, K. Tahara, A. Matsumoto, A. Tabata, H. Nagamune, Y. Yoshida, T. Minamikawa, T.
Yasui, T. Katayama, A. Furube, K. Minagawa, Y. Imada, F. Yagishita, RSC Adv., 2021, 11, 26403-26407.
64
G. Volpi, C. Garino, E. Priola, E. Diana, R. Gobetto, R. Buscaino, G. Viscardi, C. Barolo, Dyes and
Pigments, 2017, 143, 284-290.
65
C.M. Álvarez, L. Álvarez-Miguel, R. García-Rodríguez, D. Miguel, Dalton Trans., 2012, 41, 7041.
66
E. Fresta, G. Volpi, C. Garino, C. Barolo, R.D. Costa, Polyhedron, 2018, 140, 129-137.
67
A.L. Guckian, M. Doering, M. Ciesielski, O. Walter, J. Hjelm, N.M. O’Boyle, W. Henry, W.R. Browne,
J.J. McGarvey, J.G. Vos, Dalton Trans., 2004, 3943-3949.
68
T. Murai, E. Nagaya, K. Miyahara, F. Shibahara, T. Maruyama, Chem. Lett., 2013, 42, 828830.
69
F. Yagishita, T. Kinouchi, K. Hoshi, Y. Tezuka, Y. Jibu, T. Karatsu, N. Uemura, Y. Yoshida, T. Mino, M.
Sakamoto, Y. Kawamura, Tetrahedron, 2018, 74, 3728-3733.
70
G. Volpi, E. Priola, C. Garino, A. Daolio, R. Rabezzana, P.Benzi, A. Giordana, E. Diana, R. Gobetto,
Inorg. Chim. Acta, 2020, 509, 119662.
71
M.D. Weber, C. Garino, G. Volpi, E. Casamassa, M. Milanesio, C. Barolo, R.D. Costa, Dalton Trans.,
2016, 45, 8984.
72
M.E. Bluhm, C. Folli, D. Pufky, M. Kröger, O. Walter, M. Döring, Organometallics, 2005, 24, 4139-
4152.
73
K. Öfele, J. Organomet. Chem., 1968, 12, 42.
74
A.J. Arduengo, R.L. Harlow, M. Kline, J. Am. Chem. Soc., 1991, 113, 361-363.
75
W.A. Herrmann, Angew. Chem. Int. Ed., 2002, 41, 1290-1309 and references cited therein.
76
Y.W. Zhang, R. Das, Y. Li, Y.Y. Wang, Y.F. Han, Chem. Eur.J., 2019, 25, 5472 –5479.
77
Y. Koto, F. Shibahara, T. Murai, Org. Biomol. Chem., 2017, 17, 1810..
78
M. Alcarazo, S.J. Roseblade, A.R. Cowley, R. Fernández, J.M. Brown, J.M. Lassaletta, J. Am. Chem.
Soc., 2005, 127, 3290-3291
79
H.R. Pan, Y.J. Li, C.X. Yan, J. Xing, Y. Cheng, J. Org. Chem., 2010, 75, 6644–6652.
80
H.R. Pan, X.R. Wang, C.X. Yan, Z.X. Sun, Y. Cheng, Org. Biomol. Chem., 2011, 9, 2166.
81
Y. Cheng, J.H. Peng, J.Q. Li, J. Org. Chem., 2010, 75, 2382–2388.
82
C. Burstein, C.W. Lehmann, F. Glorius, Tetrahedron, 2005, 61, 6207-6217.
192
83
K. Chen, W. Chen, X. Yi, W. Chen, M. Liu, H. Wu, Chem. Commun., 2019, 55, 9287.
84
M. Nonnenmacher, D. Kunz, F. Rominger, T. Oeser, Journal of Organometallic Chemistry, 2007, 692,
2554-2563.
85
F. Yagishita, K. Nomura, S. Shiono, C. Nii, T. Mino, M. Sakamoto, Y. Kawamura, ChemistrySelect, 2016,
1, 4560–4563.
86
J. Kim, H. Hahm, J.Y. Ryu, S. Byun, D.A. Park, S.H. Lee, H. Lim, J. Lee, S. Hong, Catalysts, 2020, 10,
758.
87
M. Mihorianu, M.H. Franz, P.G. Jones, M. Freytag, G. Kelter, H.H. Fiebig, M. Tamm, I. Neda, Appl.
Organometal. Chem., 2016, 30, 581–589.
88
L. Jhulki, R. Purkait, H.K. Kisan, V. Bertolasi, A.A. Isab, C. Sinha, J. Dinda, Appl. Organomet. Chem.,
2020, 34, 5673.
89
F. Yagishita, T. Nagamori, S. Shimokawa, K. Hoshi, Y. Yoshida, Y. Imada, Y. Kawamura, Tetrahedron
Letters, 2020, 61, 151782.
90
G. Volpi, G. Magnano, I. Benesperi, D. Saccone, E. Priola, V. Gianotti, M. Milanesio, E. Centerosito,
C. Barolo, G. Viscardi, Dyes and Pigments, 2017, 137, 152-164.
91
a) M. Nakatsuka, T. Shimamura, Chem. Abstr. 2001, 134, 170632. Jpn. Kokai Tokkyo JP 2001035664,
2001. b) T. Tominaga, T. Kohama, A. Takano, Chem. Abstr. 2001, 134, 93136. Jpn. Kokai Tokkyo JP
2001006877, 2001. c) D. Kitasawa, T. Tominaga, A. Takano, Chem. Abstr. 2001, 134, 200276. Jpn. Kokai
Tokkyo JP 2001057292, 2001.
92
E. Fresta, G. Volpi, C. Garino, C. Barolo, R.D. Costa, Polyhedron, 2018, 140, 129-137.
93
F. Shibahara, E. Yamaguchi, A. Kitagawa, A. Imai, T. Murai, Tetrahedron, 2009, 65, 5062-5073.
94
E. Yamaguchi, F. Shibahara, T. Murai, J. Org. Chem., 2011, 76, 6146-6158.
95
D.R. Mohbiya, N. Sekar, ChemistrySelect, 2018, 3, 1635-1644.
96
L. Salassa, C. Garino. A. Albertino. G. Volpi, C. Nervi, R. Gobetto, K. Hardcastle, Organometallics,
2008, 27, 1427-1435.
97
G. Albrecht, C. Rössiger, J.M. Herr, H. Locke, H. Yanagi, R. Göttlich, D. Schlettwein, Phys. Status Solidi
B, 2020, 257, 1900677.
98
G. Volpi, C. Garino, E. Fresta. E. Casamassa, M. Giordano, C. Barolo, G. Viscardi, Dyes and Pigments,
2021, 192, 109455.
99
G. Ulrich, R. Ziessel, A. Harriman, Angew. Chem. Int. Ed., 2007, 47, 1184-1201.
100
A. Treibs, F.H. Kreuzer, Justus Liebigs Ann. Chem., 1968, 718, 208–223.
101
L.H. Thoresen, J. Kim, M.B. Welch, A. Burghart, K. Burgess, Synlett, 1998, 1276-1278.
102
J. Tao, D. Sun, L. Sun, Z. Li, B. Fu, J. Liu, L. Zhang, S. Wang, H. Xu, Dyes and Pigments, 2019, 168, 166-
174.
103
G. Ulrich, S. Goeb, A. De Nicola, P. Retailleau, R. Ziessel, Synlett, 2007, 10, 1517-1520.
104
N. Boens, V. Leen, W. Dehane, Chem. Soc. Rev., 2012, 41, 1130-1172.
105
T. Werner, C. Huber, S. Heinl, M. Kollmannsberfer, J. Daub, O.S. Wolfbeis, Fresenius’ J. Anal. Chem.
1997, 359, 150-154.
106
Y. Urano, D. Asanuma, Y. Hama, Y. Koyama, T. Barret, M. Kamiya, T. Nagano, T. Watanabe, A.
Hasegawa, P.L. Choyke, H. Kobayashi, Nat. Med. (N. Y.), 2009, 15, 104-409.
107
G. Gareis, C. Huber, O.S. Wolbeis, J. Daub, Chem Commun., 1997, 18, 1717-1718.
108
J.P. Malval, I. Leary, B. Valeur, New J. Chem., 2005, 29, 1089-1094.
109
K. Yamada, Y. Nomura, D. Citterio, N. Iwasawa, K. Suzuki, J. Am. Chem. Soc., 2005, 127, 6956-6957.
110
N.R. Cha, S.Y. Moon, S.K. Chang, Tetrahedron Lett., 2003, 44, 8265-8269.
111
I. Móczár, O. Huszthy, Z. Maidics, M. Kádár, K. Tóth, Tetrahedron, 2009, 65, 8250-8258.
112
K. Rurack, M. Kollmannsberger, J. Daub, Angew. Chem. Int. Ed., 2001, 40, 385-387.
113
Z. Ekmekci, M.D. Yilmaz, E.U. Akkaya, Org. Lett., 2008, 10, 461-464.
114
R.E. Gawley, H. Mao, M.M. Haque, J.B. Thorne, J.S. Pharr, J. Org. Chem., 2007, 72, 2187-2191.
115
A.K. Parhi, M.P. Kung, K. Plœssl, H.F. Kung, Tetrahedron Lett., 2008, 49, 3395-3399.
116
N. Di Cesare, J.R. Lakowicz, Tetrahedron Lett., 2001, 42, 9105-9108.
117
T. Kálai, É. Hideg, J. Jekő, K. Hideg, Tetrahedron Letters, 2003, 44, 8497-8499.
118
H.L. Kee, C. Kirmaier, L. Yu, P. Thamyongkit, W.J. Youngblood, M.E. Calder, L. Ramos, B.C. Noll, D.F.
Bocian, W.R. Scheidt, R.R. Birge, J.S. Lindsey, D. Holten, J. Phys. Chem. B, 2005, 109, 20433-20443.
119
B.M. Squeo, M. Pasini, Supramolecular Chemistry, 2020, 32, 56-70.
120
J.M. Brom, J.L. Langer, J. Alloys Compd., 2002, 1, 112-115
121
D. Zhang, Y. Wen, Y. Xiao, G. Yu, Y. Liu, X. Qian, Chem. Commun., 2008, 39, 4777-4779.

193
122
M. Chapran, E. Angioni, N.J. Findlay, B. Breig, V. Cherpak, P. Stakhira, T. Tuttle, D. Volyniuk, J.V.
Grazulevicius, Y.A. Nastishin, O.D. Lavrentovich, P.J. Skabara, ACS Appl. Mater. Interfaces, 2017, 9,
4750-4757.
123
A. Loudet, R. Bandichhor, L. Wu, K. Burgess, Tetrahedron, 2008, 64, 3642-3654..
124
I.S. Tamgho, A. Hasheminasab, J.T. Engle, V.N. Nemykin, C.J. Ziegler, J. Am. Chem. Soc., 2014, 136,
5623-5626
125
Y. Yang, X. Su, C.N. Carrol, I. Aprahamian, Chem. Sci., 2012, 3, 610-613.
126
K. Ono, K. Yoshikawa, Y. Tsuji, H. Yamaguchu, R. Uozumi, M. Tomura, K. Taga, K. Saito, Tetrahedron,
2007, 63, 9354-9358.
127
a) Y. Im, S.Y. Byun, J.H. Kim, D.R. Lee, C.S. Oh, K.S. Yook, J.Y. Lee, Adv. Funct. Mater. 2017, 27,
1603007. b) J.H. Lee, C.H. Chen, P.H. Lee, H.Y. Lin, M.K. Leung, T.L. Chiu, C.F. Lin, J. Mater. Chem. C.,
2019, 7, 5874-5888.
128
G. A. Ardizzoia, S. Brenna, S. Durini, B. Therrien, M. Veronelli, Eur. J. Inorg. Chem., 2014, 26, 4310-4319.
129
The index, τ4, is defined by the equation: τ4 =360 - (β+α)/141, where α and β are the two major
angles of a four-coordinate system. Then, for ideally square planar geometry τ = 0 (α = β = 180°),
whereas for perfectly tetrahedral geometry τ = 1 (α = β = 109.5). For further explanation see: L. Yang,
D.R. Powell, R.P Houser, Dalton Trans., 2007, 9, 955-964.
130
All calculations were carried out at the density functional (DFT) level of theory with the ADF
program package (see: G. te Velde, F. M. Bickelhaupt, E. J. Baerends, C. Fonseca Guerra, S. J. A. van
Gisbergen, J. Z. Snijders, T. Ziegler, Chemistry with ADF. J. Comput. Chem., 2001, 22, 931-967).
131
G. A. Ardizzoia, S. Brenna, S. Durini, B. Therrien, Polyhedron, 2015, 90, 214-220.
132
The trigonality index, τ, is defined by the equation: τ = (β-α)/60, where α and β are the two major
angles of a five-coordinate system (with β > α). Then, for ideally square pyramidal geometry τ = 0 (α =
β = 180°), whereas for perfectly trigonal-bipyramidal geometry τ = 1 (α = 120°, β = 180°). For further
explanation see: A.W. Addison, T.N. Rao, J. Reedijk, J. van Rijn, G.C. Verschoor, J. Chem. Soc., Dalton
Trans., 1984, 1349-1356.
133
S. Durini, G. A. Ardizzoia, S. Brenna, B. Therrien, New J. Chem., 2017, 41, 3006-3014.
134
T.S. Andy Hor, Inorg. Chim. Acta, 1988, 149, 157-158.
135
A. Kaeser, B. Delavaux-Nicot, C. Duhyon, Y. Coppel, J.F. Nierengarten, Inorg. Chem., 2013, 52, 14343-
14354.
136
G.A. Ardizzoia, G. La Monica, A. Maspero, M. Moret, N. Masciocchi, Inorg. Chem., 1997, 36, 2321-
2328.
137
W.J. Geary, Coord. Chem. Rev., 1971, 7, 81-122.
138
G. A. Ardizzoia, D. Ghiotti, B. Therrien, S. Brenna, Inorganica Chimica Acta, 2018, 471, 384-390.
139
P. Furet, C. Batzl, A. Bhatnagar, E. Francotte, G. Rihs, M. Lang, J. Med. Chem., 2004, 36, 1393-1400.
140
a) B. Jiang, J. Wang, Z.G. Huang, Org. Lett., 2012, 14, 2070-2073. b) W. Li, X. Li, W. Wu, X. Liang, J.
Ye, Chem. Commun., 2011, 47, 8325-8327.
141
A. Marchesi, S. Brenna, G.A. Ardizzoia, Dyes and Pigments, 2019, 161, 457-463.
142
G.A. Ardizzoia, G. Colombo, B. Therrien, S. Brenna, Eur. J. Inorg. Chem., 2019, 13, 1825-1831.
143
H. Geneste, B. Schäfer, Synthesis, 2001, 15, 2259–2262.
144
G. Yuan, Y. Huo, X. Nie, H. Jiang, B. Liu, X. Fang, F. Zhao, Dalton Trans., 2013, 42, 2921–2929.
145
C. Hansch, A. Leo, R. W. Taft, Chem. Rev., 1991, 91, 165–195.
146
G. Colombo, G.A Ardizzoia, J. Furrer, B. Therrien, S. Brenna, Chem. Eur. J., 2021, 27, 12380-12387.
147
P. Bovonsombat, J. Leykajarakul, C. Khan, K. Pla-on, M. Krause, P. Khanthapura, R. Ali, N. Doowa,
Tetrahedron Lett., 2009, 50, 2664-2667.
148
J.-H.Kim,S.-Y.Kim,S. Choi,H.-J.Son,S. O. Kang, Inorg.Chem., 2021, 60, 246–262.
149
L.R. Snyder, Journal of Chromatography, 1974, 92, 223-230.
150
G. Colombo, A. Romeo, G.A. Ardizzoia, J. Furrer, B. Therrien, S. Brenna, Dyes and Pigments, 2020,
821 108636-108644.
151
S. Feser, K. Meerholz, Chem. Mater., 2011, 23, 5001-5005.
152
L.O. Pålsson, A.P. Monkman, Adv. Mater., 2002, 14, 757-758.
153
a) G. te Velde, F.M. Bickelhaupt, E.J. Baerends, C. Fonseca Guerra, S.J.A. van Gisbergen, J.G. Snijders,
T. Ziegler, J. Comput. Chem., 2001, 22, 931– 967. b) C. Fonseca Guerra, J.G. Snijders, G. te Velde, E.J.
Baerends, Theor. Chem. Acc., 1998, 99, 391–403. c) E.J. Baerends, T. Ziegler, J. Autschbach, D.
Bashford, A. Bérces, F.M. Bickelhaupt, C. Bo, P.M. Boerrigter, L. Cavallo, D.P. Chong, L. Deng, R.M.
Dickson, D.E. Ellis, M. van Faassen, L. Fan, T.H. Fischer, C. Fonseca Guerra, M. Franchini, A. Ghysels, A.
194
Giammona, S.J.A. van Gisbergen, A.W. Götz, J.A. Groeneveld, O.V. Gritsenko, M. Grüning, S. Gusarov,
F.E. Harris, P. van den Hoek, C.R. Jacob, H. Jacobsen, L. Jensen, J.W. Kaminski, G. van Kessel, F. Kootstra,
A. Kovalenko, M.V. Krykunov, E. van Lenthe, D.A. McCormack, A. Michalak, M. Mitoraj, S.M. Morton,
J. Neugebauer, V.P. Nicu, L. Noodleman, V.P. Osinga, S. Patchkovskii, M. Pavanello, P.H.T. Philipsen, D.
Post, C.C. Pye, W. Ravenek, J.I. Rodríguez, P. Ros, P.R.T. Schipper, H. van Schoot, G. Schreckenbach,
J.S. Seldenthuis, M. Seth, J.G. Snijders, M. Solà, M. Swart, D. Swerhone, G. te Velde, P. Vernooijs, L.
Versluis, L. Visscher, O. Visser, F. Wang, T.A. Wesolowski, E.M. van Wezenbeek, G. Wiesenekker, S.K.
Wolff, T.K. Woo, A.L. Yakovlev, ADF2017, SCM, Theoretical Chemistry, Vrije Universiteit, Amsterdam,
The Netherlands, http://www.scm.com.
154
S. Grimme, J. Antony, S. Ehrlich, S. Krieg, J. Chem. Phys., 2010, 132, 154104.
155
a) E. van Lenthe, E.J. Baerends, J.G. Snijders, J. Chem. Phys., 1993, 99, 4597-5600. b) E. van Lenthe, E.J.
Baerends, J.G. Snijders, J. Chem. Phys., 1994, 101, 9783-9792. c) E. van Lenthe, A.E. Ehlers, E.J. Baerends, J.
Chem. Phys. 1999, 110, 8943- 8953. d) E. van Lenthe, J.G. Snijders, E.J. Baerends, J. Chem. Phys., 1996, 105,
6505-6516. e) E. van Lenthe, R. van Leeuwen, E.J. Baerends, J.G. Snijders, Int. J. Quantum Chem. 1996, 57,
281-293.
156
D. P. Chong, Mol. Phys., 2005, 103, 749-761.
157
a) A. Klamt, G.J. Schürmann, J. Chem. Soc. Perkin Trans. 2., 1993, 799-805. b) A. Klamt, V. Jonas, J. Chem.
Phys., 1996, 105, 9972-9981. c) C.C. Pye, T. Ziegler, Theor. Chem. Acc., 1999, 101, 396-408.
158
P. Hohenberg, W. Kohn, Phys. Rev., 1964, 136, B864.
159
C. Lee, W. Yang, R G. Parr, Phys. Rev. B, 1988, 37, 785.
160
J.P.Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 386512.
161
W. Kohn, L.J. Sham, Phys. Rev.,1965, 140, A1133.
162
C.A. Ullrich, E.K.U. Gross, Phys. Rev. Lett., 1995, 74, 872.

195
196

You might also like