Dalton Trans., 2019, 48, 8642-8663 - Spiro Boro - Durka

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

Dalton

Transactions
View Article Online
PAPER View Journal | View Issue
Published on 09 May 2019. Downloaded by Università degli Studi dell'Insubria on 3/3/2021 11:33:48 AM.

The effect of locking π-conjugation in


Cite this: Dalton Trans., 2019, 48,
organoboron moieties in the structures of
8642 luminescent tetracoordinate boron complexes†
Mateusz Urban, *a Krzysztof Durka, *a Patrycja Górka,a
Gabriela Wiosna-Sałyga,b Krzysztof Nawara,c Piotr Jankowski a
and
Sergiusz Luliński a

A series of 8 luminescent borafluorene complexes were extensively studied both experimentally and
theoretically in order to elucidate the effect of organoboron moiety rigidification on the physicochemical
properties of these compounds. Due to the spiro geometry of the boron atom, borafluorene and ligand
units are perpendicularly aligned, which considerably affects the flexibility of the molecule as well as its
solid-state structure. Through comparative analysis with close diphenyl analogues, we show how these
structural features influence the thermal, photoluminescent and charge mobility behaviour of the studied
compounds. Crystal structural analysis revealed that the molecules are connected mostly through
C–H⋯O and C–H⋯ π interactions formed between perpendicularly aligned borafluorene and ligand moi-
eties from neighboured molecules, serving as a complementary donor and acceptor of electron density,
respectively. This also efficiently prevents molecules from engaging in unfavoured π-stacking contact.
Furthermore, structural analysis suggests that borafluorene complexes possess a considerable degree of
flexibility due to OBN heterocycle distortions and mutual borafluorene-ligand plane movements. The
magnitude of these effects strictly depends on the ligand structure and may lead either to enhancing or
lowering the quantum yield value with respect to BPh2 analogues, while the absorption and emission
wavelength are slightly affected. The measured photophysical parameters for solid-state samples showed
that the studied complexes are much better emitters in their crystalline states that in amorphous films.
The TD-DFT and NTO calculations revealed a significant change in frontier molecular distribution, with
the HOMO localized on the borafluorene moiety. However, as the HOMO–LUMO transition is geometri-
cally not favoured, excitation occurred from HOMO−1 localized on the ligand. Finally, aggregation effects
Received 28th March 2019, were discussed based on supramolecular arrangements in crystal structures and charge transfer rates
Accepted 9th May 2019
obtained from theoretical calculations in the framework of the Marcus–Hush approximation. They
DOI: 10.1039/c9dt01332f suggest that borafluorene complexes are much better electron carriers with respect to non-annulated
rsc.li/dalton BPh2 complexes.

Introduction for applications in optoelectronic devices including organic


light-emitting diodes (OLEDs), organic field-effect transistors,
One of the most important current issues in material science photovoltaic systems, photoresponsive materials, sensors, and
is the development of highly-efficient light-emitting materials imaging materials for biological systems. Organoboron com-
pounds show analogous Lewis acidic behavior to heavier group
a
13 congeners and form luminescent complexes with various
Department of Physical Chemistry, Faculty of Chemistry,
Warsaw University of Technology, Noakowskiego 3, 00-664 Warsaw, Poland.
organic ligands. Furthermore, they turned out to be more
E-mail: kdurka@ch.pw.edu.pl, murban@ch.pw.edu.pl stable due to the more covalent character of B–C bonds when
b
Department of Molecular Physics, Faculty of Chemistry, compared with bonds formed by other main group metal
Lodz University of Technology, Żeromskiego 116, 90-924 Łódź, Poland elements. Hence, a variety of organoboron chelate complexes
c
Faculty of Mathematics and Science, Cardinal Stefan Wyszyński University,
have been extensively studied during the last few decades.1–3 It
Dewajtis 5, 01-815 Warsaw, Poland
† Electronic supplementary information (ESI) available. CCDC 1905165–1905171
has been proven that their optical properties such as their
and 1529228. For ESI and crystallographic –data in CIF or other electronic absorption and emission wavelengths can be easily adjusted
format see DOI: 10.1039/c9dt01332f by ligand (chromophore) selection and its further modification

8642 | Dalton Trans., 2019, 48, 8642–8663 This journal is © The Royal Society of Chemistry 2019
View Article Online

Dalton Transactions Paper

by extending π-electron systems or attaching electron-with- prompted some scientific groups to return to the idea of the
drawing and/or electron-donating functional groups.4–9 synthesis of B-functionalized luminescent complexes. It has
Therefore, most attention was focused on chromophore ligand been realized that proper organoboron moiety modification
diversification, while the categories of substituents on boron may strongly improve macroscopic properties on different
atoms are highly underdeveloped. This grows from the general levels. Such modifications go beyond simple color tuning, but
conviction that the type of organic group attached to the boron usually affect the quantum yield of emission (qy), thermal
center only slightly affects the physicochemical properties of stability and charge carrier mobilities. This is especially notice-
Published on 09 May 2019. Downloaded by Università degli Studi dell'Insubria on 3/3/2021 11:33:48 AM.

the complex. This thesis has been strongly supported by early able for rigid spiro complexes, i.e., where the boron atom is a
works in this area.3,10 In addition, the synthesis of functiona- part of two perpendicularly arranged heterocyclic systems
lized organoboron systems is considered rather demanding. derived from a ligand and two conjugated B-functional groups,
Thus, the simple phenyl and fluoro substituents are the most respectively. For instance, we found that 9,10-dihydro-9,10-
popular groups employed in four-coordinate boron com- diboraanthracene bis(8-oxyquinolinates) are far better emitters
pounds reported so far (Scheme 1). than any other known boron-based 8-oxyquinolinato complex
The proper selection of organic materials for the construc- owing to their high rigidity and intramolecular interplay
tion of light-emitting systems is not simple, as there are many between two chromophore units.12 This was confirmed by the
parameters that may determine its final performance in the Z. Zhang group, who synthesized strongly luminescent blue-
device. First of all, emitting materials should be characterized light emitters through the chelation of benzimidazolephenol
by good optical parameters, good electron and hole transport or pyridylphenol ligands to the diboraanthracene scaffold.13 In
capabilities and sufficient thermal, photochemical and hydro- addition, such complexes are characterized by considerably
lytic stability, with the last feature also playing an important higher thermal stability than their corresponding diphenyl-
role in the fabrication process. Although the selected groups of borin (BPh2) analogues. The 9-borafluorene (Bfl) scaffold rep-
compounds fulfill the aforementioned criteria, they may be resents another interesting alternative. Due to the enhanced
not attractive from a practical point of view due to problems Lewis acidity of the boron atom in the central 5-membered
associated with their synthesis (expensive substrates, multi- (formally anti-aromatic) borole ring, its complexation with
step reactions, low reaction yields and high cost of purifi- various donors should be generally stronger compared to dia-
cation). Another important aspect is related to the fact that rylboron centres.14–17 Moreover, one could generally expect
quantum efficiency strongly depends on the structure of the that planarization (and rigidification) of the organoboron
luminescence material and its interactions with its surround- moiety should result in the improved luminescence properties
ings. Many fluorescent dyes quench or reduce their emission of respective complexes.18,19 For example, the borafluorene
upon increasing their concentration. This is mainly due to the moiety was used for the construction of highly efficient deep-
aggregation-induced quenching effect (ACQ).11 To avoid this blue emitters.20 As established by the authors, the perpendicu-
problem, non-conjugated structures are commonly adopted. lar arrangement of organoboron and ligand moieties hampers
Unfortunately, most compounds are not efficiently emissive molecules to form close intermolecular π-stacking aggregates.
because their flexible molecular structures may likely cause As a result, the fluorescence process in the solid state was
non-radiative relaxation. In turn, the solid-state non-conju- barely affected by the environment, while the fluorescence
gated structures are usually less conductive than π-conjugated quantum yield was maintained on a similar level as in solu-
systems, which may reduce device performance. tion. Moreover, such systems are characterized by improved
One of the methods used to circumvent these problems is thermal and charge carrier properties, which strongly benefit
to stabilize the molecular conformation through an internal from the perspective of their application in OLED devices. In
covalent linker, which locks the planar conformation of a subsequent studies by Wang, rigid spiro-BODIPy complexes
molecule intentionally leading to an enhanced emission inten- derived from benzoxaborin, benzothioxaborin and borafluor-
sity and increased charge carrier mobility. Besides others, this ene scaffolds were found to form J-type aggregates rather than
typically face-to-face π-stacking dimers, facilitating bright emis-
sion in the solid state.21 Similar molecular design was used for
the synthesis of four-coordinate organoboron complexes con-
taining electron-rich azaborin moieties and electron-poor
ligand enabling emission through the thermally activated
delayed fluorescence (TADF) mechanism.22 Along this line, Ho,
Chi, and Chou showed that the emission in isoquinolinyl pyra-
zolate complexes of diarylborinic acid exhibited dual emission
from locally excited intraligands or organoboron-to-ligand
charge transfer states, while analogous systems based on bora-
fluorene or oxaborin cores revealed solely CT emissions.23
However, this was accompanied by a significant drop in fluo-
Scheme 1 Modifications of boron appendages in tetracoordinate rescence quantum yields (from almost 100% for the BPh2
boron complexes. system to less than 0.5% for the boracyclic complex).

This journal is © The Royal Society of Chemistry 2019 Dalton Trans., 2019, 48, 8642–8663 | 8643
View Article Online

Paper Dalton Transactions

The above-mentioned examples demonstrate the high thus eliminating the necessity for base addition and simplify-
potential of boracyclic complexes for their application in opto- ing product isolation. In this manner complexes 3a–8a were
electronic devices, related to their promising physicochemical synthesized with good yields (Scheme 3, I). The respective
features in solid-state systems. This has encouraged us to BPh2-substituted compounds were synthesized in reactions of
conduct comprehensive studies on borafluorene complexes the previously obtained borinic ester Ph2BOEt with corres-
(Bfl) bearing different types of (N,O) chelate ligands. ponding ligands according to the literature procedure.24
Consequently, we designed and synthesized a series of 7 new Naturally, this method allows for the facile synthesis of
Published on 09 May 2019. Downloaded by Università degli Studi dell'Insubria on 3/3/2021 11:33:48 AM.

luminescent compounds (Scheme 2). This was supplemented unsubstituted borafluorene complexes; however, the synthesis
with one borafluorene complex previously studied by Zhang of various borafluorene-functionalized systems is often proble-
and a series of corresponding BPh2 systems used for compara- matic. This comes from the fact that boron halides are incom-
tive reasons.20 First, we revised the synthetic methodology, patible with ethereal solvents like Et2O and THF, commonly
allowing us to optimize the reaction conditions. Subsequently, used for generating dilithiated intermediates, while the double
all obtained complexes were fully characterized and compared Br/Li exchange in the non-coordinating solvent is usually not
to their BPh2 analogues addressing stability issues and dis- effective. The alternative method relies on the transmetallation
cussing optical properties in solution, solid phases and thin reaction between toxic stannylfluorene and hazardous boron
films. Finally, the aggregation effects were evaluated in terms trihalide.25 Therefore, we decided to investigate other possible
of crystal structures and charge transfer rates derived from synthetic pathways employing milder reagents, like B(OMe)3,
theoretical calculations in the framework of the Marcus–Hush to obtain 9-methoxy substituted borafluorene derivatives.
approximation. However, upon reacting B(OMe)3 with 2,2′-dilithiobiphenyl,
the borole heterocycle does not form (Scheme 3, II). This was
most likely caused by hampered rotation along the aryl-aryl
Results and discussion bond of the biphenyl core, resulting from the presence of the
bulky anionic B(OMe)3− group strongly interacting with a sol-
Synthesis and characterization vated lithium cation. The problem of steric hindrance was also
The synthesis of complexes 3a–8a proceeded according to the previously observed in synthesis utilizing BCl3.26 Surprisingly,
improved literature protocol.17,21 In the common procedure, we isolated a side product formed most likely by the reaction
9-chloroborafluorene is mixed with the corresponding ligand between 2,2′-dilithiobiphenyl and two molecules of an inter-
in the presence of amines producing the target complex. mediate “ate” complex, which after HCl/Et2O quenching
However, in our initial syntheses, we observed that the amine and the addition of 8-hydroxyquinoline gave compound 13
competes with the ligand to the boron atom, and the obtained (Scheme 3, III). This compound was thoroughly studied in our
reaction yields were often not satisfactory. Thus, in our pro- recent paper.24
cedure, prior to ligand addition, 9-chloroborafluorene was con- We managed to overcome these difficulties inspired by a
verted to 9-methoxyborafluorene by the addition of anhydrous protocol developed by Yamaguchi for the synthesis of polycyc-
methanol. The byproduct (HCl) was removed under vacuum, lic thiophene-fused boroles,15,27 later improved by us28 and by

Scheme 2 Studied borafluorene (Bfl) complexes 1a–8a and their diphenylboron (BPh2) analogues 1b–8b.

8644 | Dalton Trans., 2019, 48, 8642–8663 This journal is © The Royal Society of Chemistry 2019
View Article Online

Dalton Transactions Paper


Published on 09 May 2019. Downloaded by Università degli Studi dell'Insubria on 3/3/2021 11:33:48 AM.

Scheme 3 Utilization of BCl3 and B(OMe)3 for the synthesis of the borafluorene scaffold. Complexes 3a–8a were obtained from
9-methoxyborafluorene.

Wagner29 in the synthesis of 9,10-dihydro-9,10-diboraanthra- multinuclear NMR spectroscopy (1H, 13C, 11B) and HRMS ana-
cenes (Scheme 4). First, 2-bromo-2′-(dimethoxyboryl)biphenyl lysis. They are soluble in common organic solvents but are,
(11) was obtained and then was reacted with t-BuLi/THF at however, somewhat weaker than their BPh2 analogues. They
−105 °C. In the resultant boron-lithium intermediate, the showed long-time stability under ambient conditions both in
steric hindrance of the B(OMe)2 group was low enough to the solid state and in solution. In our previous work, we
facilitate spontaneous borole ring cyclization, affording demonstrated that boranil complexes are prone to hydrolytic
9-methoxyborafluorene after quenching with HCl/Et2O. It was deboronation in dilute solutions, which may limit their poten-
isolated as 1a after complexation with 8-hydroxyquinoline with tial application.30,31 Thus, the solution-stability of 7a and 8a
a reasonable yield (70%). The established protocol was then complexes were compared to their BPh2 analogues using time-
applied for the synthesis of a more problematic perfluoro depending UV-vis absorption experiments. As depicted in
derivative. Consequently, 2-bromo-2′-(dimethoxyboryl)octa- Fig. S36–39 (ESI†), boranil 7a did not undergo hydrolysis even
fluorobiphenyl (12) was obtained as the key intermediate in after 7 days, whereas 7b decomposed within a few hours.
the first step with a high yield (91%). After halogen/lithium Similar improvement was as also achieved for bis(boranil) 8a
exchange followed by cyclization, HCl quenching and 8-hydro- (Fig. S38–39, ESI†).
xyquinoline addition, the perfluorinated borafluorene 8-oxy- According to the DSC and TGA measurements, all bora-
quinolinato complex 2a was isolated with a very good yield of fluorene complexes showed improved thermal stabilities com-
75%. pared with their non-annulated analogues 1b–8b (Table 1,
The obtained borafluorene complexes were isolated as col- Fig. S2–20, ESI†). The melting and decomposition tempera-
orful crystalline solids. All complexes were characterized by tures are systematically higher by 50–100 °C, owing to the

Scheme 4 Two-step protocol for the synthesis of 9-metoxyborafluorene leading to 1a and 2a after complexation with 8-HQ.

This journal is © The Royal Society of Chemistry 2019 Dalton Trans., 2019, 48, 8642–8663 | 8645
View Article Online

Paper Dalton Transactions

Table 1 Summary of DSC and TGA analysis for Bfl complexes (1a–8a) and their BPh2 analogues (1b–8b)

1 2 3 4 5 6 7 8

Bfl Tm/°C a
218.1 278.0 319.1 301 20
256.8 302.8 206.1 —c
Tdec/°Cb 261.0 286.8 319.4 362 20 262.3 303.7 268.6 449.4
BPh2 Tm/°Ca 206.6 161.1 200 32 257 20 160 4 211 4 156.7 305.2
Tdec/°Cb 270.3 227.6 247 33 383 20 238 4 264 4 288.0 363.0
a
From onset. b Temp. at 5% loss of mass. c No melting point was observed.
Published on 09 May 2019. Downloaded by Università degli Studi dell'Insubria on 3/3/2021 11:33:48 AM.

higher rigidity of the coupled systems. From the studied average (Table S2, ESI†). This is quite surprising taking into
group, bis(boranil) complex 8a showed superior thermal stabi- account the high boron Lewis acidity of the borafluorene tri-co-
lity, as it began to decompose at above ca. 450 °C. We also ordinated precursor.14
observed that the fluorination of the borafluorene moiety sig- Another intriguing conclusion derived from the structural
nificantly enhanced thermal stability. Furthermore, the analysis is the fact that Bfl complexes can display substantial
selected systems (1a, 4a, 7a) did not decompose after melting, conformational mobility. Specifically, the chelate boracyclic
which is rather rare for tetracoordinate boron chelate com- OBN ring can adopt 3 extreme conformations including one
plexes, as melting is typically followed by rapid decomposition. flat and two half-chair conformations, B-distorted and O-dis-
torted. In the first form, the entire OBN boracycle is essentially
Crystal structure analysis flat, being also coplanar with the rest of the ligand. Such a geo-
Borafluorene complexes were obtained as crystalline materials metry is naturally expected for 5-membered complexes 1a and
and studied using single-crystal X-ray diffraction techniques. 2a, but it is also observed in 5a and 8a containing a 6-mem-
This was supplemented by the already known structure of 4a.20 bered OBN ring. The ring distortions can be described in
The molecular structures are depicted in Fig. 1. They revealed terms of the distance from boron or oxygen atoms to the
similar molecular features, which resulted from the perpen- ligand plane (Fig. 2 and Table 1). Accordingly, complexes 3a,
dicular spiro arrangement of the ligand and borafluorene moi- 4a and 6a are classified in the O-distorted group, as the oxygen
eties linked by a central boron atom. Despite ligand structure atoms significantly deviate out of the ligand plane, whereas 7a
diversification, the B–N, B–O and B–C distances remained on a has been classified to B-distorted group, as such a deviation
very similar level, oscillating near 1.61 Å, 1.50 Å and 1.62 Å, occurs for the boron atom. It seems that the corresponding
respectively. When compared to the corresponding non-annu- non-annulated PBh2 complexes stabilize in distorted confor-
lated BPh2 analogues 1b–8b, the B–N and B–O bonds are only mations more often (5a vs. 5b and 8a vs. 8b), and the magni-
slightly shorter (by 0.010 Å and 0.004 Å on average, respect- tude of the OBN ring distortions is more significant (Table 1).
ively), while the B–C distances are longer by only 0.005 Å on This is also confirmed by the theoretical calculations (B3LYP/

Fig. 1 The molecular structures of 1a–8a. Thermal motions are given as ADPs at the 50% probability level. Hydrogens atoms are omitted for clarity.

8646 | Dalton Trans., 2019, 48, 8642–8663 This journal is © The Royal Society of Chemistry 2019
View Article Online

Dalton Transactions Paper


Published on 09 May 2019. Downloaded by Università degli Studi dell'Insubria on 3/3/2021 11:33:48 AM.

Fig. 2 OBN ring deviations, ligand-borafluorene twisting, in-plane movement and out-of-plane movement observed in the studied complexes. The
numerical values for each system are provided in Table 2.

6-311+G(d,p)) showing that most BPh2 complexes are stable be partially attributed to the higher π-electron stabilization in
solely in their distorted forms, while in the case of the Bfl com- the naphthalene-based ditopic ligand, which stiffens the
plexes, the OBN ring is substantially flattened. Furthermore, ligand frame.
the interconversion barrier between distorted and flat forms is Finally, the above-discussed molecular distortions seemed
relatively low (5–10 kJ mol−1), indicating some expected degree to affect the geometry of the ligand in 3a–8a. As indicated by
of flexibility in solution. Apart from the OBN ring distortions, the τP1–P4 angle (Fig. 2), the aza/pyridine and phenolate parts
Bfl complexes may also be affected by the orientational distor- of the ligand may be twisted along the C–C bond. This was
tions of ligand and borafluorene mean planes (P1 and P2, especially noticeable in boranils 7–8, where the dihedral angle
respectively, Fig. 2). For instance, P1 and P2 may deviate from between the phenolate and phenyl groups ranged from 35° to
their mutual perpendicular arrangement (τP1–P2). Furthermore, 50°. Secondly, the comparison with non-annulated complexes
the rudder-like out-of-plane displacement is defined by the di- 3b–8b revealed that this angle was partially reduced in Bfl
hedral angle τP1–P3 between the P1 and P3 planes, the latter is leading to the flattening of ligand structure and the more
perpendicular to P2 and bisects the borafluorene middle C–C efficient delocalization of electrons. The conformational
bond and the boron atom. Another important effect is related adjustment can be attributed to the release of steric hindrance
to the in-plane ligand shift, which can be described by the between organoboron and ligand moieties, as the flat
angle αCent1–Cent2 formed by the OBN centroid, boron atom π-conjugated borafluorene skeleton is more compact. As a result
and inner borafluorene ring centroid. of more efficient π-electron delocalization, the absorption and
From the comparative analysis of all studied structures, it emission peaks in 3a–8a are bathochromically shifted with
seems that 5-membered 8-oxyquinolinato complexes 1a and 2a respect to 3b–8b. This was also confirmed by the τP1–P4-con-
are the most symmetrical and rigid. Some internal strains were strained optimisation energy scan performed for compound 7a,
released by the Bfl-ligand plane twisting (τP1–P2). In contrast, in showing that the ligand flattened upon excitation, and both the
2-pyridylphenol- and azole-derived complexes 3a–6a, the Bfl absorption and emission wavelengths decreased with the twist-
and ligand planes are nearly perpendicular, while they are ing of ligand structure (Fig. S51†). The 8-oxyquinolinato com-
highly susceptible to ligand rudder-like deviations (τP1–P3 plexes are naturally not affected by such effects, which stays in
ranging from 14° to 24°) and in-plane movements. Despite the agreement with our spectroscopic observations (see Fig. 5).
fact that αCent1–Cent2 is rather small in 5a and 6a, the elonga- The obtained borafluorene complexes pack in a different
tion of carbon ellipsoids in the borafluorene ring suggests in- manner (with exception to 5a and 6a, which are isostructural);
plane conformational lability, which will definitely be more however, a closer inspection of intermolecular interaction
pronounced in solution. Regarding molecular geometry, struc- pattern revealed that several structural motifs are often
ture 7a deserves special attention. It features a highly distorted encountered. The most characteristic one involves a C–H⋯O
OBN heterocycle accompanied by a significant twist (τP1–P2 = hydrogen bond between the Lewis basic oxygen atom from the
71.4(3)°) as well as in-plane (αCent1–Cent2 = 19.3(3)°) and out-of- OBN boracycle and the C–H group from the ligand moiety of
plane (τP1–P3 = 42.2(3)°) shifts of the ligand moiety with respect the neighboured molecule. It is usually accompanied by
to the borafluorene ring (P2). In contrast, the bisboranil ana- C–H⋯π interactions between the ligand and borafluorene
logue 8a was almost unaffected by such deviations. This can π-electron cloud (Fig. 3a). Alternatively, in the case of 6a, the

This journal is © The Royal Society of Chemistry 2019 Dalton Trans., 2019, 48, 8642–8663 | 8647
View Article Online

Paper Dalton Transactions

Table 2 Comparison of the selected geometrical parameters in borafluorene complexes (1a–8a) and their non-annulated analogues (1b–8b). dB–P1
and dO–P1 add to the distance of the boron and oxygen atoms from the P1 plane, as shown in Fig. 2. αCent1−cent2 represents the angle formed
between corresponding centroids defined by two spiro-derived heterocyclic rings peaked at the boron atom; τ are dihedral angles formed between
P1–4 best fit mean-square planes, defined in Fig. 2

dB–P1b/Å dO–P1/Å αCent1–Cent2/° αC–B–C/° τP1–P2/° τP1–P3/° τP1–P4/°

1a a 0.019(4) 0.041(4) 3.8(2) 100.6(2) 85.9(3) 0.4(3) —


0.110(4) 0.070(4) 4.6(2) 100.8(2) 83.3(3) 4.9(3)
Published on 09 May 2019. Downloaded by Università degli Studi dell'Insubria on 3/3/2021 11:33:48 AM.

1b 34 0.043 0.016 — 117.1 — — —


2a 0.014(1) 0.100(1) 2.8(1) 99.1(1) 81.9(1) 7.5(1) —
2b 35 0.133 0.035 — 116.9 — — —
3a 0.066(4) 0.474(4) 13.4(2) 109.3(2) 87.8(4) 20.1(4) 11.8(4)
3b a 32 −0.114 0.699 — 109.3 — — 20.6
−0.102 0.587 116.5 13.8
4a 20 −0.164 0.341 16.3 100.3 87.67 23.24 8.0
4b 20 −0.039 0.596 — 116 — — 13.5
5a −0.070(2) 0.171(2) 6.9(1) 100.1(1) 89.6(1) 14.7(1) 4.6(1)
5b 4 −0.039 0.596 — 116.1 — — 11.0
6a −0.077(2) 0.171(2) 4.6(1) 100.1(1) 86.1(1) 16.9(1) 8.5(1)
6b 4 −0.112 0.540 — 114.9 — — 12.5
7a 0.642(5) 0.034(5) 19.3(3) 99.8(3) 74.1(3) 42.2(3) 35.8(3)
7b 31 0.707 0.035 — 113.7 — — 42.3
8a a −0.006(3) 0.041(3) 7.3(2) 100.2(2) 89.1(2) 3.3(2) 42.4(2)
−0.169(3) 0.004(3) 2.4(2) 100.3(2) 84.2(2) 4.8(2) 53.6(2)
8b a 0.739(1) 0.037(1) — 116.3(1) — — 47.66(1)
0.831(1) 0.090(1) 116.3(1) 50.34(1)
a
Two non-equivalent boron centres in the asymmetric unit. b Negative values are used to distinguish between distortions below and above the
plane.

Fig. 3 1D chains formed in (a) 1a, (b) 6a and (c) 3a. (d) Packing diagram showing the dispositions of molecules near the 8-oxyquinolinato moiety in
1a. (e) π–π stacking dimeric motif in 5a.

8648 | Dalton Trans., 2019, 48, 8642–8663 This journal is © The Royal Society of Chemistry 2019
View Article Online

Dalton Transactions Paper

sulfur atom acts as an electron acceptor from the π-density of


the borafluorene unit (d1S⋯C(π) = 3.270(1) Å and d2S⋯C(π) =
3.323(1) Å) in accordance with the σ-hole concept originating
from the anisotropic distribution of electron density around
the sulfur atom (Fig. 3b).36–39 Further propagation of these
dimeric motifs in structures 1a, 3a–6a leads to the formation
of an infinite one-dimensional molecular chain, which may
Published on 09 May 2019. Downloaded by Università degli Studi dell'Insubria on 3/3/2021 11:33:48 AM.

also adopt a ribbon topology, as observed in 3a (Fig. 3c).


Besides, the studied complexes tend to form numerous
C–H⋯π interactions within other structural motifs. In most
cases, the ligand acts as a C–H donor, imposing the perpen-
dicular arrangement of ligand and electron-rich borafluorene
moieties from neighboured molecules (Fig. 3d). The ligand
can also act as an acceptor of C–H⋯π interactions from per-
pendicularly aligned borafluorene or ligand moieties, however,
these contacts are less common. Fig. 4 Molecular layers in structure 2a.
Comprehensive analysis of all structural motifs in the
studied systems (excluding 2a) shows that π-stacking aggre-
gates are rather avoided, unless the ligand surface area is at were grouped together in the crystal structure forming π–π
least 1.5 bigger than the borafluorene surface area, thereby stacking molecular layers propagating in a direction perpen-
giving enough space for parallel arrangements. For instance, dicular to the borafluorene planes. They were separated by the
π–π head-to-tail oriented dimers have been found in the struc- layers formed by 8-oxyquinolinato units also arranged by
tures bearing flat extended π-conjugated oxazole (5a) and thia- π-stacking interactions. Thus, the supramolecular structure of
zole (6a) ligands (Fig. 3e). The distance between the planes of 2a possessed a double-layer architecture. It is noticeable that
the stacking rings is equal to 3.409(2) Å and 3.415(2) Å, C–H⋯ π interactions with the borafluorene ring were not
respectively, but phenoxa or phenothiazole moieties only par- observed due to its much weaker electron-donating properties
tially overlap. Centroids of stacked molecules are separated by reduced by the electron-withdrawing fluorine atoms.
9.634(5) Å (5a) and 9.538(5) Å (6a), with the slip angle defined
by the centroid-to-centroid vector and ligand plane normal
vector equalling 20.7° and 21.2°. Such dimers likely exhibit Photophysical properties
J-aggregation type behaviour.40,41 It was previously postulated UV-vis absorption and emission spectra of complexes 1a–8a
that spiro arrangement around the boron centre promotes were measured in dichloromethane (DCM) solution. For com-
J-aggregation.21 However, we have noticed that its occurrence parison, the spectra and emission quantum yields for the
strictly depends on some specific structural and molecular referential BPh2 series 1b–8b were also measured. All results
requirements. For instance, in structure 4a, the perpendicu- are summarized in Table 3 and the corresponding absorption
larly aligned N-Ph group blocks the formation of J-aggregates. and emission spectra are presented in Fig. 5. For all studied
Furthermore, in structures comprising ligand and borafluor- systems, switching from BPh2 to Bfl resulted in the red shift of
ene units of similar sizes (such as 1a and 3a), their spiro absorption bands. The magnitude of this effect depends on
arrangement efficiently prevents the π–π stacking interactions the ligand type. In 8–oxyquinolinato (1a and 2a) and 2-( pyri-
of adjacent molecules. Similarly, steric requirements preclude dine-2-yl)phenolato (3a) complexes, the change is almost negli-
the formation of π-stacking dimers in boranils 7a and 8a. gible (1–2 nm), but in the remaining compounds (4a–7a), the
Specifically in 7a, the boracycle ring bisects the ligand struc- bathochromic shift of the absorption band is ca. 10 nm, reach-
ture in two approximately similar parts, impeding the parallel ing 17 nm for bis(boranil) 8a. With respect to BPh2 complexes,
approach. In the case of 8a, the naphthalene core is located in the rise of molar absorption coefficients by 460–1300 M−1
the inner part of the molecule; thus, such π–π stacking cm−1 was observed for 1a–3a; meanwhile, it significantly
arrangement is hampered by steric hindrance. Instead, in the decreased for 4a–6a and 8a (by 2000–3000 M−1 cm−1). A
crystal structures of bornils, numerous C–H⋯ π contacts similar trend was observed for the corresponding emission
appear between the ligand and borafluorene units. We noted spectra. For 1a, the bathochromic shift was very subtle (2 nm),
that such contacts affect the conformation of the phenyl but for 3a–6a and 8a, the fluorescence bands were red shifted
group, eventually leading to its rotation along the C–C(N) bond by 10–14 nm and even more (by 25 nm) for 7a. The observed
(τP1–P4) to maximize the efficiency of C–H⋯ π interaction with red shift can be connected to the flattening of the ligand
a neighbouring molecule. frame, as already described in the crystal structure section. In
A completely different picture of the supramolecular assem- the case of 7a, it can be also associated with higher ligand
bly was observed in the perfluorinated borafluorene complex flexibility and its tendency to flattening upon excitation (more
2a (Fig. 4). As commonly encountered for systems comprising information in ESI,† Fig. S51 and Table S4). Exceptionally,
fluorine-rich molecular fragments, the F8-borafluorene rings complex 2a displayed a slight hypsochromic shift of the emis-

This journal is © The Royal Society of Chemistry 2019 Dalton Trans., 2019, 48, 8642–8663 | 8649
View Article Online

Paper Dalton Transactions

Table 3 UV-vis absorption and emission data for 1–8 in DCM and in the solid state

λabs/nm (ε/M−1 cm−1) λem/nmb φfb

Solution Filma Solutionc Powder Film Solutionc Powder Film Δṽ/cm−1 τ/ns

1a 397 (4450) 403/397 513 494 520 0.49 0.45 0.16 5696 25.3
1b 395 (3700) — 511 492 — 0.40 f 24 0.60 0.50 5750 27.7
2a 391 (3790) 415/407 510 516 521 0.38 0.40 0.20 5968 24.0
2b 390 (2480) — 514 479 — 0.32 0.50 — 6186 22.7
Published on 09 May 2019. Downloaded by Università degli Studi dell'Insubria on 3/3/2021 11:33:48 AM.

3a 362 (6160) 373/369 482 462 478 0.28 0.20 0.12 6877 8.3
3b 360 (5700) — 471 457 448 33 0.44 (0.40d)32 0.38 0.39 33 6546 9.8
4a 365 (9325) 369/366 443 445 454 (440)20 0.36 (0.75)20 0.32 0.10 (0.57)20 4975 3.5; 7.1
4b 357 (12 400) — 435 434 440 0.72 (0.73)20 0.13 0.4720 5023 5.7
5a 383 (9380) 393/386 461 456 471 0.16 0.20 0.07 4418 2.0; 5.8
5b 373 (12 200) — 449 (452)47 455 (462)47 — 0.68 (0.55)47 0.53 47 — 4538 6.2; 10.7
6a 400 (7550) 411/404 496 492 508 0.22 0.15 0.09 4839 4.3
6b 389 (10 800) — 482 (485)47 484 (494)47 — 0.70 (0.65)47 0.60 47 — 5088 8.6
7a 409 (4670) 417/415 559 542 556 0.003 0.05 —e 6561 —e
7b 400 (4600) — 534 518 — 0.10 f 31 0.09 — 6273 2.3
8a 546 (11 800) 552/552 604 662 622 0.65 0.37 0.04 1759 8.6
8b 529 (13 900) — 594 622 — 0.76 30 0.18 — 2069 10.5
a
Maxima of the absorption and excitation spectra, respectively. b In brackets the previously published values are presented. c Unless stated other-
wise, solvent = DCM. d CHCl3. e Fluorescence was too weak for the precise calculation of the quantum yield and evaluation of lifetime. f The value
was recalculated using the quantum yield of standard equal to 0.544.

sion band (by 4 nm) respective to the non-annulated analogue photophysical processes. More significant changes were
2b. We supposed that it may be a result of the stronger elec- observed for the complexes of the azole family 4–6. The intro-
tron withdrawing character of F8-borafluorene as compared to duction of borafluorene in 4a resulted in biexponential decay
B(C6F5)2. The shape of the emission bands of all Bfl complexes with lifetimes of 3.5 ns and 7.1 ns. In the case of 5, both 5a
remained analogous to their BPh2 references. The only differ- and 5b exhibited biexponential decay; however, the fluorescent
ence was observed in 8a, where the fine structure was slightly lifetime was significantly (2–3 times) shorter in the borafluor-
more pronounced as a result of enhanced structural rigidity. ene complex. On the other hand, in the case of 6, monoexpo-
Even though the absorption and emission bands were not nential decay was observed, with a two-fold shortening of life-
strongly affected by the type organoboron moiety (with some time for 6a with respect to 6b. Although the lifetimes of 3.5 ns
exceptions), important differences were observed for the in 4a and 2.0 ns in 5a resemble those for the corresponding
quantum yields of emission. Regarding 8-oxyquinolinato com- ligands,42,43 their origins seem to be different here, as both
plexes, a boost of qy was observed upon introduction of the compounds are stable (the ligand was not detected in solu-
borafluorene scaffold. Specifically, in 1a, qy increased from tion), and the ligands did not absorb light at the excitation
0.40 to 0.49 and for 2a from 0.32 to 0.38. For the remaining wavelength (overlays are presented in ESI, Fig. S35†).
complexes, a decrease in qy values was observed. This is As boron-based complexes are often considered as emitters
especially noticeable for the azole ligand family (4a–6a), where in OLEDs, we also performed measurements of their optical
qy changes from 0.72 to 0.36 (4b vs. 4b), from 0.68 to 0.16 (5b properties in amorphous thin films deposited by spin-coating
vs 5a) and from 0.70 to 0.22 (6b vs 6a). It can be noted that the on a quartz substrate (Fig. 6). With the exception of 3a and 7a,
qy measured by us for 4a was inconsistent with the literature we observed a red shift of the emission spectra by 7–18 nm
value (0.36 vs. 0.75), while the qy values for the non-annulated with respect to solution measurements. Complexes 3a and 7a
4b system were very similar (0.72 vs 0.73).20 We supposed that exhibited a very slight blue shift of 4 and 3 nm, respectively.
fluorescence quenching results from substantial confor- The most important difference, however, was the strong
mational mobility related to the OBN ring and ligand-bora- quenching of fluorescence in thin films. The quantum yields
fluorene plane deviations. In an extreme case (7a), the fluo- dropped at least two-fold for all Bfl complexes. This stands in
rescence was almost completely quenched (0.003/0.10 for 7a/ opposition to most previously studied BPh2 complexes, where
7b, respectively), reflecting the very high conformational flexi- the quantum yields usually remain on a similar level or are
bility of this molecule. On the other hand, in comparison with even higher upon layer formation.13,20,33,44,45 Accordingly,
8b (0.76), the higher π-electron stabilization of naphthalene- measured by us, the qy for the thin film of the BPh2 complex
based ditopic ligand in bis(boranil) 8a confirms its intense 1b is higher than in solution (0.50 and 0.40, respectively).
fluorescence (0.67). Especially puzzling was, again, the much lower qy of the 4a
The fluorescence lifetimes were evaluated from DCM solu- layer with regard to the already published value (0.10 vs 0.57).
tions, and exemplary decays are depicted in ESI (Fig. S33 and This may be, however, attributed to a different method applied
S34†). For 1–3 and 8, the decays are similar when comparing for the layer deposition, i.e., spin-coating or sublimation,
a/b pairs, in accordance with their similar mechanism of leading to a different local molecular environment.46

8650 | Dalton Trans., 2019, 48, 8642–8663 This journal is © The Royal Society of Chemistry 2019
View Article Online

Dalton Transactions Paper


Published on 09 May 2019. Downloaded by Università degli Studi dell'Insubria on 3/3/2021 11:33:48 AM.

Fig. 5 Overlay of absorption (left column) and emission (right column) spectra in DCM for borafluorene and BPh2 complexes.

The absorption and excitation spectra of thin layers were were deposited by the rapid evaporation of chloroform solutions
also measured (Fig. 6). The red shift of the absorption band by through spin-counting. The blue shift of the excitation spectra
5–14 nm with respect to the solution was observed in all cases. suggests that emission proceeds from other types of aggregates
For 7a and 8a, normalized absorption and excitation spectra comprised of weakly interacting molecules. To verify this
aligned fairly well; however in the remaining cases, the exci- hypothesis, the fluorescence was measured for powder samples.
tation spectra were blue-shifted, and for 1a, 4a and 6a, they In contrast to amorphous layers, the fluorescence quantum
resembled the absorption spectra from solution. On the basis yields of powdered samples were retained from solution (1a–
of these observations, we supposed that there was more than 4a, 6a) or even slightly increased (5a, 7a). Interestingly, in most
one UV-vis absorbing entity in the layer, some of which were cases (1a, 3a–7a), the emission was blue shifted with regard to
not luminescent. Both the bathochromic shift in absorption the film samples. Compared to solution measurements, the
and the quenching of fluorescence would presumably indicate fluorescence was either blue shifted (1a, 3a, 7a) or remained
the formation of agglomerates interacting through π-stacking. unchanged (4a–6a). Exceptionally, the red shift of the emission
Although, with the exception of 2a, such interactions are maxima was observed for perfluorinated 2a. This effect was
avoided in crystal structures, we supposed that such agglomer- likely associated with its distinct crystal structure dominated
ates may be kinetically formed in amorphous films, as layers by π-stacking forces.

This journal is © The Royal Society of Chemistry 2019 Dalton Trans., 2019, 48, 8642–8663 | 8651
View Article Online

Paper Dalton Transactions


Published on 09 May 2019. Downloaded by Università degli Studi dell'Insubria on 3/3/2021 11:33:48 AM.

Fig. 6 Overlay of absorption, emission and excitation spectra in the solid state ( powder, thin film) and in solution (DCM) for complexes 1a–8a.

It could also be noted that compound 8a displays a signifi- efficiency, is in accordance with our previous studies showing
cant red shift of the emission band going from the solution that bis(boranil) complexes are susceptible to aggregation
(604 nm), through the film (622 nm) to the powder sample effects.30 On the other hand, qy is still much higher when
(662 nm). This, along with the decrease of fluorescence compared to the value measured for the powder sample of 8b

8652 | Dalton Trans., 2019, 48, 8642–8663 This journal is © The Royal Society of Chemistry 2019
View Article Online

Dalton Transactions Paper

(0.37 vs. 0.18). We supposed that qy enhancement, systems are stored in the ESI (Fig. S40–S48†). Our study was
accompanied by the red shift of the emission band, most complemented by the evaluation of the absorption and emis-
likely resulted from the planarization of phenyl moieties on sion energies using time-dependent density functional theory
two ends of the ligand frame, leading to more effective (TD-DFT) at the CAM-B3LYP/6-311++G(d,p) level of theory
π-electron coupling. (Table S3).
The calculations for BPh2 systems 1b–8b show that the
Electrochemical properties HOMO and LUMO orbitals are located mainly on the ligand
Published on 09 May 2019. Downloaded by Università degli Studi dell'Insubria on 3/3/2021 11:33:48 AM.

The electrochemical properties of all complexes were analyzed moiety and that the first excitation results from the electron
based on corresponding peak potentials (oxidation – Eox1/2 and transition between these orbitals. The second excitation is
reduction – Ered1/2) obtained using cyclic voltammetry (CV) associated with the electron transfer from deeper HOMO levels
measurements in CH2Cl2 and were reported with respect to (usually HOMO−1) located on the BPh2 moiety. The TD-DFT
the FeCp2/FeCp2+ redox couple. For irreversible redox pro- calculations show that the latter transition is not effective as
cesses, the potentials were estimated based on the onset indicated by the low value of oscillator strength ( f ). In the
values. The voltammograms are shown in ESI (Fig. S21 and case of borafluorene complexes, the HOMO orbital located on
22†), and the data are summarized in Table 4. Most observed the borafluorene ring is significantly elevated with respect to
electrochemical processes were irreversible. Complexes 3a and non-annulated BPh2 complexes, thereby exceeding the energy
8a displayed a quasi-reversible one-electron oxidation wave, of the occupied orbital located on the ligand. Thus, the lowest
while the reduction process was partially reversible for laying transition possesses borafluorene-to-ligand charge
complex 6a. The oxidation potentials for complexes 1a and transfer character and can be characterized by a significantly
3–8a were below 1 V, usually remaining between 0.8 and 0.9 red-shifted absorption wavelength. However, this transition
V. They are systematically lower by ca. 0.2 V with respect to the was not observed in the corresponding UV-vis spectra, which
values for BPh2 complexes, indicating that the oxidation was also in agreement with the TD-DFT calculations giving the
process presumably occurs at the borafluorene moiety. The low value of the oscillator strength for this process. Instead,
second anodic peak in 1a and 6a may be ascribed to ligand the second excitation is symmetry-allowed with a larger value
oxidation. In an extreme example (3a), the oxidation potential of the oscillator strength and is therefore much more efficient.
was lowered by 0.5 V compared to 3b. Regarding the cathodic This occurs from the lower-lying HOMO level (usually
process, with some exceptions, the reduction potentials of the HOMO−1) delocalized over the ligand frame and partially bor-
Bfl systems were on a similar level or slightly lower with afluorene moiety. The obtained absorption and emission wave-
respect to the BPh2 complexes. The absolute values strictly lengths are consistent with the experimental values. The dis-
depend on the ligand type. This indicated that an extra elec- tinct electronic behavior observed for the BPh2 complexes was
tron locates onto the ligand during the reduction process. The retained in 2a. The HOMO level is mainly located on the
above statements remain in agreement with our theoretical cal- ligand, while the borafluorene moiety contributes to lower-
culations, showing that the HOMO orbital is located on the Bfl laying HOMOs, which result from the strong inductive effect of
moiety, while the LUMO resides on a ligand. This was also 8 fluorine atoms.
confirmed by the electrochemical behavior of the pair 2a–2b, In the next step, we utilized Marcus theory48–50 in order to
where both oxidation and reduction potentials are almost establish a connection between supramolecular structure and
identical. For both complexes, the HOMO and LUMO levels electron/hole mobilities in the solid state. The calculated elec-
are located on the 8-oxyquinolinato ligand, while the orbital tron (Λe) and hole (Λh) reorganization energies (B3LYP/6-311+G
associated with the borafluorene core was substantially (d,p)) (Table 5) suggest that borafluorene complexes are much
lowered upon fluorination. Finally, it can be noted that the better hole-transporting materials than BPh2 ones, while the
electrochemical stability window of borafluorene complexes electron-transporting abilities remain on a similar level for
was narrower, while still sufficient for their potential appli- those two groups of complexes. This was especially noticeable
cation in optoelectronic devices. for systems 2a–4a and 6a, where Λh decreased by more than
0.1 eV. Furthermore, it is noticeable that the hole reorganiz-
Theoretical calculations ation energies were much smaller than the electron reorganiz-
To gain a deeper insight into the photophysical features of all ation energies. The lowest values of the reorganisation ener-
studied complexes, a series of theoretical calculations were gies were obtained for complexes 1a and 2a, bearing the 8-oxy-
performed using the Gaussian09 package. The molecular geo- quinolinato ligand, and we considered them as the most prom-
metries, fully optimized at the B3LYP/6-311++G(d,p) level of ising. This is in accordance with our previous observations of
theory, were generally preserved from the crystal structures them being the most rigid among the studied systems; thus,
with the exception of 7b, where significant deviations from they are less likely to undergo structural distortions upon char-
crystal geometry were observed. In addition to molecular ging. Complex 6a also seemed promising. Beneficially, it also
orbital calculations (summarized in Fig. 7), natural transition exhibited the most balanced hole and electron reorganization
orbitals (NTO) were obtained from calculations at the same energies, which are desirable for emitters in OLED appli-
level of theory. NTOs for the representative pair of compounds cations. In contrast, complexes 4a and 7a showed the largest
(1a, 1b) are depicted in Fig. 8; the results for the remaining values of reorganization energies due to their high molecular

This journal is © The Royal Society of Chemistry 2019 Dalton Trans., 2019, 48, 8642–8663 | 8653
View Article Online

Paper Dalton Transactions

Table 4 Redox potentials and energy band gap for 1–8 based on CV states of 7a with respect to its neutral form. According to
measurements. All potentials are given with respect to the FeCp2/ expectations, the rigidification of the boranil structure in
FeCp2+ redox couple
compound 8a, decreased both electron and hole reorganiz-
ation energies, even though the molecule is two times larger
ΔE/eV Optical
Ered1/2/V Eox1/2/V (ΔE/nm) bandgap than 7a.
We expanded our research by calculating the transfer inte-
1 a −2.04 0.88a/1.06b 2.92 (425) 3.12 (397)
3.10 (400) grals (V), following by the evaluation of charge hopping rates
Published on 09 May 2019. Downloaded by Università degli Studi dell'Insubria on 3/3/2021 11:33:48 AM.

b −2.00 1.03a/1.40b 3.03 (410) 3.14 (395) (W) and drift mobilities of charge hopping (μ) for various dimer
2 a −1.78 1.27 3.05 (407) 3.17 (391) pairs with the geometries extracted from the crystal structures.
b −1.77 1.22 2.99 (415) 3.18 (390)
3 a −2.50/−2.71c 0.72a <3.22 (385) 3.42 (362)d We selected complexes 1a, 2a and 6a for these studies as those
b 33 −2.24 1.20 3.44 (360) 3.44 (360) systems are characterized by the lowest electron and hole re-
4 a 20 −2.40 1.00 3.25 (381) 3.40 (365) organization energies derived from their high molecular rigidity.
b 20 −2.46 1.14 3.60 (344) 3.47 (357)
5 a −2.09 0.93 3.02 (411) 3.24 (383) The obtained values were compared to the ones calculated for
b4 −2.28 1.17 3.45 (359) 3.32 (373) the corresponding BPh2 complexes 1b, 2b and 6b. In addition,
6 a −2.08a 0.89/1.31b 2.97 (418) 3.10 (400) the borafluorene fluorination effect was elucidated from the
b4 −2.21 1.10 3.33 (372) 3.19 (389)
7 a −1.80 0.89 2.69 (461) 3.03 (409) comparison between 1a and 2a. Our study was complemented
b 31 −1.94 1.10 3.04 (408) 3.10 (400) by the analysis performed for 6a as a representative compound
8 a −1.34 0.78 2.12 (585) 2.27 (546) selected from the azole ligand family. Following the latest
b 30 −1.36 1.06 2.42 (512) 2.34 (529)
procedure developed by Mamada and Tanaka,51 we also
a
Mean of the anodic and cathodic peak potentials. b Two oxidation evaluated the contribution of each dimer to the total drift
peaks were observed. c The reduction potential of 3a was out of the
mobility. The calculated drift mobilities for the most contri-
electrochemical stability window of DCM and thus could not be
measured. The reduction potential was estimated based on the optical buting dimers are shown in Table 6. The dimers are depicted
bandgap from UV-vis spectrum. d Optical bandgap calculated in Fig. 9 (1a, 2a and 6a) and Fig. S52–54† (1b, 2b and 6b). The
from λabs.
complete data for all dimers can be found in the ESI
(Table S5†).
Even though the electron reorganization energies are
flexibility. Specifically, the calculations showed that 4a was higher than the hole reorganization energies, it is evident that
highly susceptible to N-Ph ring rotation upon charging. the electron-transporting properties are much better with
Similarly, significant conformational changes within the respect to the hole-transporting properties for all materials.
ligand frame (τP1–P4) were observed in the negative and positive This is in accordance with the experimental observations for

Fig. 7 Diagram showing calculated frontier orbital levels in all studied systems. Black, blue, red and brown horizontal lines depict the LUMO,
HOMO, HOMO−1 and HOMO−2 levels, respectively. Vertical lines depict the observed transition from the occupied orbital respective to the used
colour.

8654 | Dalton Trans., 2019, 48, 8642–8663 This journal is © The Royal Society of Chemistry 2019
View Article Online

Dalton Transactions Paper


Published on 09 May 2019. Downloaded by Università degli Studi dell'Insubria on 3/3/2021 11:33:48 AM.

Fig. 8 Natural transition orbital pairs for the first and second singlet excited-state geometry of 1a and 1b. The NTOs for other complexes are pre-
sented in the ESI (Fig. S40–S48†).

Table 5 Calculated reorganization energies for electrons (Λe) and holes hole-transporting properties can be improved by incorporation
(Λh) in the studied complexes of amine substituents in ligand structure.4 Secondly, fusing
the B-phenyl appendages to the borafluorene π-electron system
Λe/eV Λh/eV
significantly increases the electron mobility of the compound.
Bfl BPh2 Bfl BPh2 In turn, the hole mobility decreased in 1a and 2a, while both
hole and electron mobilities increased for 6a with respect to
1 0.370 0.379 0.196 0.213
2 0.414 0.473 0.234 0.343 6b. Similar behaviour was observed for the imidazole deriva-
3 0.361 0.384 0.305 0.548 tives 4a and 4b, where both electron and hole mobilities were
4 0.583 0.586 0.314 0.416 enhanced after fusing the biphenyl π-electron skeleton.20
5 0.429 0.438 0.282 0.362
6 0.367 0.390 0.298 0.401 A comprehensive analysis of all dimers showed that charge
7 0.577 0.572 0.308 0.361 mobility is highly anisotropic in the crystal structure. For com-
8 0.430 0.474 0.286 0.324 plexes 1a and 6a, the highest electron hopping rates were esti-
mated for dimer D1, responsible for the formation of the pre-
viously discussed infinite one-dimensional molecular chain
1b 10 and other studies of luminescent four-coordinate boron (Fig. 3a and b). Within this motif, molecules are relatively
complexes,47,52,53 suggesting that they can rather serve as elec- close to each other and are stabilized by several close C–H⋯O
tron transporting materials, although it was reported that the (1a, 6a), C–H⋯π (1a, 6a) and S⋯π (6a) interactions. The ligand

Table 6 Calculated hopping rates (We/Wh) and drift mobility contributions (%μe/%μh) given relative to the total value (μe/μh) for selected dimers in
complexes 1a, 1b, 2a, 2b, 6a and 6b. r corresponds to the distance between the centres of mass; N is the number of the same neighbouring pairs
found in the corresponding crystal structures

N r/Å We × 10−12/s−1 %μe μe/cm2 V−1 s−1 We × 10−12/s−1 %μh μh/cm2 V−1 s−1

1a D1 1 8.411 473.0 69.8% 0.167 9.0 7.9% 0.009


D2 1 9.049 15.2 0.1% 25.9 75.9%
D3 1 8.506 302.1 29.1% 12.3 15.1%
1b D1 1 6.138 207.5 58.4% 0.089 32.5 61.3% 0.014
D2 2 7.499 136.8 37.9% 8.3 6.0%
D3 2 8.590 57.5 2.2% 33.9 32.7%
2a D1 2 11.882 285.9 91.6% 0.403 4.5 69.1% 0.005
D2 2 8.889 115.8 8.4% 1.0 1.9%
D3 1 8.613 1.4 0.0% 8.0 29.0%
2b D1 2 7.805 103.8 95.8% 0.069 21.9 85.4% 0.014
D2 2 9.440 17.4 4.0% 7.4 14.4%
6a D1 2 8.036 243.1 99.4% 0.185 25.6 86.7% 0.017
6b D1 2 8.553 6.4 5.8% 0.017 10.8 32.8% 0.011
D2 2 9.767 17.4 56.3% 8.8 28.0%
D3 2 10.133 13.7 37.7% 9.9 38.6%

This journal is © The Royal Society of Chemistry 2019 Dalton Trans., 2019, 48, 8642–8663 | 8655
View Article Online

Paper Dalton Transactions


Published on 09 May 2019. Downloaded by Università degli Studi dell'Insubria on 3/3/2021 11:33:48 AM.

Fig. 9 Dimeric motifs contributing the most to the charge transfer processes in 1a, 2a and 6a.

and borafluorene/OBN heterocycle serve as a complementary their luminescent, structural, thermal and charge transport
pair of acceptors and donors of electron density, respectively. properties to their close BPh2 analogues in order to elucidate
Thus, borafluorene-to-ligand and oxygen-to-ligand electron the structure–property relationship governing their behaviour
hopping pathways were expected for these systems. It was also both in solution and in the solid state. In the first step, we pro-
noticeable that dimer D1 contributes the most to the total hole posed an alternative methodology for the synthesis of the bora-
mobility of 6a, which probably results from the ambiphilic fluorene scaffold. It utilized two-step protocol elaborating the
character of the sulfur atom. In turn, the effective hole trans- introduction of one B(OMe)2 group to 2′-bromobiphenyl, fol-
port in 1a apparently occurs through the borafluorene-to-bora- lowed by selective Br/Li exchange and spontaneous borafluo-
fluorene pathway realized within dimer D2. Different charge rene ring closure. Although the already developed protocol for
transport behaviour was observed for 2a. This naturally comes the synthesis of 9-chloroborafluorene is relatively simple, the
from the distinctive supramolecular packing of this com- proposed procedure enables the facile synthesis of functiona-
pound. It exhibits excellent electron and hole transport pro- lized borafluorene complexes such as perfluorinated
perties within dimer D1 through the C–H⋯F interactions borafluorene.
formed between the perpendicularly aligned ligand and the F8- From the analysis of all studied systems several important
borafluorene moieties. In addition, the contribution of dimer conclusions on the structure–property relation can be derived:
D3 to the hole transport is not negligible. However, we sup- (1) Formal coupling of B-phenyl groups resulting in the bora-
posed that the charge would flow through the C–H⋯F inter- fluorene scaffold does not necessarily lead to the rigidification
actions between F8-borafluorene and the ligand rather than of molecular geometry. In the case of complexes forming a
the π-stacked quinoline moieties. This thesis is supported by 5-membered boraheterocycle with ligand (1a, 2a), the structure
the poor charge transport behaviour of other π-stacking dimers indeed becomes more rigid. However, in 3a–8a, which are
analysed in this structure. Similarly, in 6a, the partially π–π based on the 6-membered OBN ring, the molecules regain
overlapping dimers, identified as J-aggregates (Fig. 3e), did not additional degrees of flexibility related to OBN ring distortions,
contribute to the charge transport process. ligand in-plane and out of plane movements, ligand twisting,
and Bfl-ligand rotation, all resulting from the more compact
nature of the borafluorene moiety with respect to two phenyl
Conclusions substituents.
(2) These molecular features can be related to observed
We performed a comprehensive study on a series of borafluo- photophysical behaviour. In the case of 1a and 2a, the
rene complexes with varying ligand structure and compared quantum yield in solution increased, which can be attributed

8656 | Dalton Trans., 2019, 48, 8642–8663 This journal is © The Royal Society of Chemistry 2019
View Article Online

Dalton Transactions Paper

to the higher rigidity of the molecule. This confirmed our reversible. On the other hand, Bfl complexes are more suscep-
initial assumption of the stabilization effect of molecular con- tible to irreversible oxidation.
formation, which reduces non-radiative energy loss through It is anticipated that further optimization of the structure of
oscillations and rotations. On the other hand, complexes luminescent boron complexes will afford new luminescent
3a–7a showed some substantial conformational mobility materials characterized by improved thermal, hydrolytic,
resulting from OBN ring deviations and ligand-borafluorene electrochemical stability, intense fluorescence with finely-
plane movements. This led to a decrease of the qy value; in an tuned emission colour and enhanced charge mobility.
Published on 09 May 2019. Downloaded by Università degli Studi dell'Insubria on 3/3/2021 11:33:48 AM.

extreme case 7a, the fluorescence was almost completely Consequently, a detailed understanding of the fundamental
quenched. In turn, the fluorescence in bis(boranil) remained factors influencing the photophysical behaviour of lumines-
intensive, which probably resulted from higher π-electron cent materials is essential. We suppose that this can be
stabilization in the naphthalene-based ditopic ligand. achieved through careful structural design taking into con-
(3) Another important structural effect is related to the sideration not only the molecular composition (ligand struc-
twisting of the ligand frame. The magnitude of this rotation is ture, etc.) but also its conformation. We hope that our study
governed by the steric hindrance of the boron substituents. will provide an important contribution in this area and will be
Ligand flattening in less-strained borafluorene structure helpful for designing of novel luminescent materials such as
results in the bathochromic shift of the absorption and emis- TADF emitters. Further exploration of spiro-tetracoordinate
sion bands. organoboron complexes is currently in progress in our
(4) The photophysical parameters for the solid-state laboratory.
samples showed that the Bfl complexes were much better emit-
ters in their crystalline states than in amorphous films. This is
contrary to our initial expectations, as it seems that bulk aggre-
gation does not necessarily quench qy. We suppose that this Experimental section
comes from the specific structural features of these complexes.
Crystal structural analysis revealed that molecules are con- General comment
nected through C–H⋯O and C–H⋯ π interactions leading to Starting materials: 2,2′-dibromobiphenyl, B(OMe)3, BCl3 (1 M
perpendicular arrangements of the molecular units. This also in hexane), n-BuLi (1.6 M and 10 M in hexane), 8-hydroxy-
efficiently prevents the molecules from engaging in π-stacking quinoline, N-phenyl-o-phenylenediamine, 2-aminophenol,
contacts. In line with these observations, the absorption and 2-aminotiophenol, aniline and salicylaldehyde were bought
emission bands are only slightly affected by aggregation. In from Sigma-Aldrich. t-BuLi (1.9 M in pentane) was bought
contrast, the aggregation in thin films is governed by the from Alfa-Aesar and 2-( pyridine-2-yl)phenol from ChemScene.
kinetic factors, which likely favour the occurrence of π–π stack- 2-Aminotiophenol was purified by distillation under reduced
ing aggregates. pressure and N-phenyl-o-phenylenediamine by recrystallization
(5) Although the calculated reorganization energies are from Et2O with the addition of activated carbon. The remain-
higher than the hole reorganization energies, the evaluated ing substrates were used as received without additional purifi-
charge mobilities suggest that borafluorene complexes could cation. THF and Et2O were dried by heating to reflux with
serve as electron transporting materials. In that respect, bora- sodium or potassium and benzophenone, and distilled under
fluorenes are more efficient electron carriers than the corres- argon. Hexane and toluene were purified using MBraun SPS
ponding BPh2 complexes. According to our calculations, the and stored over 3 Å molecular sieves. Reactions and manipula-
C–H⋯π, C–H⋯O and S⋯ π interactions facilitate electron tions involving air- and moisture-sensitive reagents were
transport along one-dimensional molecular chains with the carried out under argon atmosphere.
ligand acting as an acceptor, while the perpendicularly aligned 2,2′-Dibromooctafluorobiphenyl,54 Ph2BOEt,30 C6F5B(OEt)2,31
borafluorene ring/OBN oxygen atom acts as a donor of electron 2-(2′-hydroxyphenyl)-1,3-benzoxazole,55 and 1,5-dihydroxy-
density. Alternatively, the hole and electron transport are ben- naphthalene-2,6-dicarboxaldehyde30 were prepared according
eficial through C–H⋯F contacts formed between F8-borafluo- to the literature procedures. 9-Chloroborafluorene was syn-
rene and ligand moieties in 2a. In contrast, the π-stacking thesized from 2,2′-dibromobiphenyl using a known procedure
dimers and J-aggregates do not favour charge transport. with 75% yield after distillation (78–83 °C/1 × 10−3 Torr).26
(6) Theoretical calculations show that the HOMO orbital in Syntheses of 2-(2′-hydroxyphenyl)-N-phenyl-benzimidazole and
Bfl complexes (with expect 2a) was located on the borafluorene 2-(2′-hydroxyphenyl)-1,3-benzothiazole were conducted based
moiety; however, the charge transition is not effective from on procedures known for their analogues.56 Complexes 1b, 7b
this orbital. Instead, electron excitations occur mainly due to and 8b were synthesized according to the published
π–π* transition from a deeper HOMO orbital located on the procedures.24,30,31 Synthetic procedures for ligands and com-
ligand, which resembles the excitation mechanism in non- plexes 2b and 3b–6b were reported in the ESI.† Prior to per-
annulated BPh2 analogues. forming optical and CV measurements, all complexes were
(7) All obtained systems showed improved hydrolytic recrystallized from acetone (by dissolving in excess of solvent
and thermal stabilities with respect to BPh2 complexes. and slow evaporation on a rotary evaporator) and dried under
Furthermore, for the selected compounds the melting is vacuum.

This journal is © The Royal Society of Chemistry 2019 Dalton Trans., 2019, 48, 8642–8663 | 8657
View Article Online

Paper Dalton Transactions

Characterization data Hexane (5 mL) was added, and the obtained slurry was filtered
1 11 13 19
H, B, C, F NMR spectra were recorded on Bruker under an argon atmosphere. The remaining solid was washed
Advance III 300 MHz and Agilent NMR 400 MHz DDR2 spec- three times with hexane (3 × 5 mL) and the combined filtrate
trometers. 1H and 13C chemical shifts were referenced to TMS was concentrated in vacuo. A crude product was purified by dis-
using known chemical shifts of solvent residual peaks. 11B and tillation under reduced pressure to give pure 12 as a colourless
19
F NMR chemical shifts were given relative to BF3·Et2O and liquid, b.p. 67–68 °C (3 × 10−3 Torr). Yield 1.64 g (91%). 1H
CFCl3, respectively. In the 13C NMR spectra, the resonances of NMR (400 MHz, CDCl3) δ = 3.52 (d, J = 0.7 Hz, 1H) ppm. 19F
NMR (376 MHz, CDCl3) δ = −127.84 (ddd, J = 22.1, 10.1,
Published on 09 May 2019. Downloaded by Università degli Studi dell'Insubria on 3/3/2021 11:33:48 AM.

boron-bound carbon atoms were not observed in most cases as


a result of their broadening by the quadrupolar boron nucleus. 3.3 Hz, 1F), −130.57 (ddd, J = 22.0, 13.2, 3.3 Hz, 1F), −134.58
In addition, the signals corresponding to carbon atoms bound (ddt, J = 22.2, 9.8, 2.8 Hz, 1F), −136.04 to −136.58 (m, 1F),
to fluorine atoms were sparingly visible due to the multiple −151.10 (ddd, J = 21.8, 20.1, 3.7 Hz, 1F), −152.28 to −152.82
coupling with fluorine atoms (2a, 12). Used deuterated sol- (m, 2F), −154.82 (ddd, J = 22.0, 20.1, 3.2 Hz, 1F) ppm. 13C
vents were dried with 3 Å molecular sieves and moisture-sensi- NMR (101 MHz, CDCl3) δ = 52.90 (d, J = 1.2 Hz) ppm.
tive samples were prepared under argon atmosphere. HR-MS Complexes
analyses were performed on a GCT Premier mass spectrometer
equipped with an EI ion source and a MALDI SYNAPT G2-S Complex 1a – (8′-oxyquinolinato)-9-borafluorene. t-BuLi
HDMS spectrometer equipped with an ESI ion source. The (0.53 mL, 1.0 mmol, 1.9 M) was added to a solution of
HRMS of the new alkyl boronates were not measured due to dimethyl 2′-bromobiphenyl-2-boronate 11 (307 mg, 1.0 mmol)
their fast hydrolysis to the corresponding borinic acids under in THF at −105 °C over 20 min. The mixture was then warmed
standard analytical conditions. to −45 °C over 1 h. After stirring for 30 min HCl/Et2O
(0.50 mL, 1.0 mmol, 2 M) was added, and the mixture was
brought to rt. The obtained colourless mixture showed blue
Synthesis luminescent under UV, which suggests the formation of
11 – Dimethyl 2′-bromobiphenyl-2-boronate. n-BuLi (3.0 mL, 9-metoxyborafluorene. After stirring for 30 min, 8-hydroxy-
30 mmol, 10 M) diluted in hexane (10 mL) was added dropwise quinoline (0.15 g, 1 mmol) was added. 2 hours later, mixture
to a solution of 2,2′-dibromobiphenyl (9.36 g, 30 mmol) in a volatilities were removed under vacuum and the remaining
mixture of anhydrous Et2O (30 mL) and THF (30 mL) at solid was dissolved in CHCl3. The solution was extracted with
−78 °C, and stirred at this temperature for 1 h. To a resulting water; the organic phase was dried with Na2SO4 and concen-
suspension, a solution of B(OMe)3 (3.35 mL, 30 mmol) in Et2O trated on a rotary evaporator. To the obtained residue, EtOH
(10 mL) was added dropwise at −78 °C and stirred for 1 h. The (3 mL) was added and the mixture was vigorously stirred for
obtained clear solution was treated with Me3SiCl (3.8 mL, 10 min. The formed precipitate was filtered, washed with cold
30 mmol) and warmed to rt. The mixture was then stirred for EtOH and hexane, and dried under vacuum to give 1a as a
1 h, and the volatiles were removed under reduced pressure. bright yellow crystalline solid. Yield 216 mg (70%). 1H NMR
Et2O (20 mL) was added, and the obtained slurry was filtered (300 MHz, CDCl3) δ = 8.40 (dd, J = 8.3, 1.1 Hz, 1H), 8.02 (dd, J =
under an argon atmosphere. The filter cake was washed three 5.1, 1.0 Hz, 1H), 7.72 (dd, J = 8.4, 7.7 Hz, 1H), 7.68 (dt, J = 7.6,
times with Et2O (3 × 5 mL) whereupon the combined filtrate 0.9 Hz, 2H), 7.47 (dd, J = 8.3, 5.0 Hz, 1H), 7.32 (dd, J = 8.4,
was concentrated in vacuo. The crude product was purified by 0.7 Hz, 1H), 7.30 (ddd, J = 7.6, 7.1, 1.6 Hz, 2H), 7.20 (dd, J =
fractional distillation under reduced pressure to give the 7.7, 0.7 Hz, 1H), 7.11 (ddd, J = 7.1, 1.6, 0.8 Hz, 2H), 7.06 (td, J =
product as a colourless liquid, b.p. 85–88 °C (1 × 10−3 Torr). 7.1, 1.0 Hz, 2H) ppm. 11B NMR (96 MHz, CDCl3) δ = 12.4 (s)
Yield 6.89 g (75%). 1H NMR (400 MHz, CDCl3) δ = 7.61 (ddd, ppm. 13C NMR (101 MHz, CDCl3) δ 159.7, 149.7, 139.2, 138.7,
J = 8.0, 1.2, 0.5 Hz, 1H), 7.48–7.44 (m, 1H), 7.38–7.30 (m, 3H), 138.6, 133.0, 129.8, 128.9, 128.6, 127.1, 123.0, 119.6, 112.4,
7.28 (dd, J = 6.8, 1.2 Hz, 1H), 7.26 (ddd, J = 7.6, 2.3, 0.5 Hz, 109.8 ppm. HRMS (ESI): C21H14BNO [M + H]+ Calcd: 307.1168,
1H), 7.14 (ddd, J = 8.0, 6.8, 2.3 Hz, 1H), 3.36 (s, 6H) ppm. 13C found: 307.1162.
NMR (101 MHz, CDCl3) δ = 144.1, 143.6, 133.1, 132.2, 131.6, 2a – (8′-oxyquinolinato)octafluoro-9-borafluorene. Compound
130.1, 128.9, 128.4, 127.4, 127.1, 122.9, 52.5. 11B NMR 2a was synthesized as described for 1a using dimethyl
(96 MHz, CDCl3) δ = 29.3 (s) ppm. 2′-bromo-octafluorobiphenyl-2-boronate (476 mg, 1.06 mmol),
12 – Dimethyl 2′-bromo-octafluorobiphenyl-2-boronate. t-BuLi (0.56 mL, 1.06 mmol, 1.9 M), HCl/Et2O (0.53 mL,
n-BuLi (2.50 mL, 4.0 mmol, 1.6 M) was added dropwise to a 1.06 mmol, 2 M), and 8-hydroxyquinoline (0.16 g, 1.1 mmol). A
solution of 2,2′-dibromooctafluorobiphenyl (1.82 g, 4.0 mmol) bright yellow crystalline solid was obtained. Yield 357 mg
in anhydrous Et2O (40 mL) at −100 °C and stirred at this temp- (75%). 1H NMR (300 MHz, CDCl3) δ = 8.60 (dd, J = 8.3, 1.0 Hz,
erature for 1 h. The mixture was clear. B(OMe)3 (0.45 mL, 1H), 8.15 (dd, J = 5.1, 1.0 Hz, 1H), 7.78 (dd, J = 8.4, 7.7 Hz, 1H),
4.0 mmol) was added dropwise at −100 °C and allowed to 7.67 (dd, J = 8.3, 5.1 Hz, 1H), 7.44 (dd, J = 8.4, 0.6 Hz, 1H), 7.25
warm to −70 °C over 1 h. A white precipitate formed. After the (dd, J = 7.7, 0.6 Hz, 1H) ppm. 11B NMR (96 MHz, CDCl3) δ =
addition of HCl/Et2O (2.0 mL, 4.0 mmol, 2 M), the mixture 10.4 (s) ppm. 13C NMR (101 MHz, CDCl3) δ = 157.9, 140.7,
clarified. After warming to rt, the mixture was stirred for 139.1, 138.3, 133.5, 128.5, 123.2, 114.0, 110.9 ppm. 19F NMR
30 min and volatiles were removed under reduced pressure. (376 MHz, CDCl3) δ = −133.45 to −133.85 (m, 1F), −135.87 to

8658 | Dalton Trans., 2019, 48, 8642–8663 This journal is © The Royal Society of Chemistry 2019
View Article Online

Dalton Transactions Paper

−136.13 (m, 1F), −153.17 to −153.49 (m, 1F), −155.08 to 8.7, 7.2, 1.7 Hz, 1H), 7.34–7.27 (m, 5H), 7.12 (ddd, J = 8.5, 7.2,
−155.41 (m, 1F) ppm. HRMS (ESI): C21H6BF8NO [M]+ Calcd: 1.2 Hz, 1H), 7.09–7.03 (m, 3H), 7.01 (dt, J = 8.6, 0.9 Hz, 1H),
451.0415, found: 451.0398. 6.97 (ddd, J = 8.1, 7.2, 1.1 Hz, 1H) ppm.13C NMR (101 MHz,
3a – (2′-(2″-pyridyl)phenolato)-9-borafluorene. To a solution CDCl3) δ = 168.6, 160.2, 148.6, 144.6, 136.9, 130.1, 130.0, 128.7,
of 9-chloroborafluorene (0.20 g, 1.0 mmol) in anhydrous 128.3, 127.4, 127.1, 126.6, 121.9, 121.1, 120.7, 119.8, 119.6,
hexane (5 mL), anhydrous MeOH (ca. 0.05 mL, 1 mmol) was 115.1 ppm. 11B NMR (96 MHz, CDCl3) δ = 6.8 (s) ppm. HRMS
added at 0 °C, and the mixture discoloured. Then, volatilities (ESI): C25H16BNOS [M]+ Calcd: 389.1046, found: 389.1043.
7a – salicylideneanilinato-9-borafluorene. Compound 7a was
Published on 09 May 2019. Downloaded by Università degli Studi dell'Insubria on 3/3/2021 11:33:48 AM.

were removed under vacuum and the obtained residue was


redissolved in anhydrous DCM (5 mL). 2-(Pyridine-2-yl)phenol synthesized as described for 3a using 9-chloroborafluorene
(1 mmol) was added, and the mixture was left stirring over- (0.22 g, 1.1 mmol), MeOH (ca. 0.05 mL, 1 mmol), salicylalde-
night. The solvent was evaporated under vacuum and to the hyde (0.11, 1.0 mmol) and aniline (0.10 mL, 1.1 mmol). Yellow
obtained solid, EtOH (ca. 5 mL) was added, and the mixture crystalline solid. Yield 215 mg (60%). 1H NMR (300 MHz,
was left vigorously stirring for 10 min. The formed precipitate CDCl3) δ = 8.40 (s, 1H), 7.57–7.49 (m, 3H), 7.43 (dd, J = 7.8,
was filtered off, washed with EtOH and hexane, and dried 1.7 Hz, 1H), 7.36 (d, J = 7.0 Hz, 2H), 7.19 (td, J = 7.5, 1.2 Hz,
under vacuum to give 3a as a pale yellowish crystalline solid. 2H), 7.16–7.05 (m, 3H), 7.04–6.91 (m, 4H), 6.91–6.85 (m, 2H)
Yield 133 mg (62%). 1H NMR (400 MHz, CDCl3) δ = 8.05 (dt, J = ppm. 13C NMR (101 MHz, CDCl3) δ = 163.9, 163.6, 148.5,
8.2, 1.3 Hz, 1H), 8.03–7.94 (m, 2H), 7.86 (dd, J = 7.9, 1.6 Hz, 144.7, 138.8, 132.3, 130.0, 128.9, 128.7, 128.4, 126.8, 123.2,
1H), 7.64 (dt, J = 7.6, 0.9 Hz, 2H), 7.45 (ddd, J = 8.2, 7.2, 1.6 Hz, 120.5, 119.5, 119.4, 117.9 ppm. 11B NMR (96 MHz, CDCl3) δ =
1H), 7.26 (td, J = 7.4, 1.3 Hz, 2H), 7.23–7.14 (m, 3H), 7.11 (ddd, 6.6 (s) ppm. HRMS (ESI): C25H18BNO [M]+ Calcd: 359.1486,
J = 8.2, 1.2, 0.4 Hz, 1H), 7.06 (ddd, J = 7.9, 7.2, 1.3 Hz, 1H), 7.03 found: 359.1481.
(td, J = 7.2, 1.1 Hz, 2H) ppm. 13C NMR (101 MHz, CDCl3) δ = 8a – (2′,6′-bis(N-phenyliminomethyl)-1′,5′-dioxynaphthale-
160.6, 151.2, 149.0, 142.3, 141.2, 134.5, 129.9, 128.6, 127.1, nato)-bis(9-borafluorene). To a solution of 9-chloroborafluor-
125.6, 122.9, 121.6, 121.0, 120.1, 119.6, 119.1 ppm. 11B NMR ene (0.40 g, 2.0 mmol) in anhydrous hexane (5 mL), anhydrous
(96 MHz, CDCl3) δ = 6.6 (s) ppm. HRMS (ESI): C23H16BNO MeOH (ca. 0.08 mL, 2 mmol) was added dropwise at 0 °C until
[M + H]+ Calcd: 334.1398, found: 334.1392. the mixture discoloured. Then, volatilities were removed under
4a – (2′-(2″-oxyphenyl)-N-phenyl-benzimidazolato)-9-bora- vacuum and the remaining solid was redissolved in anhydrous
fluorene. Compound 4a was synthesized as described for 3a 1,2-dichloroethane (30 mL). Next, aniline (0.18 mL, 2.0 mmol)
using 9-chloroborafluorene (0.20 g, 1.0 mmol), MeOH (ca. and 1,5-dihydroxynaphthalene-2,6-dicarboxaldehyde (216 mg,
0.05 mL, 1 mmol) and 2-(2′-hydroxyphenyl)-1-phenyl-benzo- 1.0 mmol) were added sequentially, the mixture was warmed
imidazole (288 mg, 1.0 mmol). Whitish powder. Yield 334 mg to 50 °C and was left stirring for 3 days. The mixture was then
(75%). 1H NMR (400 MHz, CDCl3) δ = 7.76–7.69 (m, 5H), cooled to rt, and the solvent was evaporated under vacuum. To
7.59–7.53 (m, 2H), 7.38–7.27 (m, 5H), 7.18 (ddd, J = 8.5, 7.4, the obtained solid, EtOH (ca. 10 mL) was added, and the
1.1 Hz, 1H), 7.14 (dd, J = 8.4, 1.2 Hz, 1H), 7.09–6.98 (m, 4H), mixture was left vigorously stirring for 30 min. The formed pre-
6.81 (dd, J = 8.1, 1.6 Hz, 1H), 6.66–6.56 (m, 2H) ppm. cipitate was filtered off, washed with EtOH, DCM and hexane,
5a – (2′-(2″-oxyphenyl)-benzoxazolato)-9-borafluorene. and dried under vacuum to give 3a as a purple powder. Yield
Compound 5a was synthesized as described for 3a using 328 mg (47%). Before performing CV and UV-vis measure-
9-chloroborafluorene (0.20 g, 1.0 mmol), MeOH (ca. 0.05 mL, ments the product was recrystallised from DCM (8a is spar-
1 mmol) and 2-(2′-hydroxyphenyl)-1,3-benzoxazole (210 mg, ingly soluble in acetone) with the addition of activated carbon.
1.0 mmol). Pale yellow crystalline solid. Yield 251 mg (67%). 1
H NMR (300 MHz, CDCl3) δ = 8.44 (s, 2H), 7.79 (d, J = 8.4 Hz,
1
H NMR (400 MHz, CDCl3) δ = 7.95 (dd, J = 7.9, 1.7 Hz, 1H), 2H), 7.58 (dt, J = 7.5, 0.9 Hz, 4H), 7.41 (ddd, J = 7.1, 1.3, 0.7 Hz,
7.70 (dt, J = 7.6, 0.8 Hz, 2H), 7.60 (dt, J = 8.4, 0.9 Hz, 1H), 7.56 4H), 7.27 (d, J = 8.5 Hz, 2H), 7.23 (td, J = 7.5, 1.3 Hz, 4H),
(ddd, J = 8.5, 7.2, 1.8 Hz, 1H), 7.38 (dt, J = 7.0, 1.0 Hz, 2H), 7.16–7.05 (m, 6H), 7.01 (td, J = 7.3, 1.0 Hz, 4H), 6.92–6.86 (m,
7.36–7.25 (m, 3H), 7.15–7.02 (m, 4H), 7.01 (ddd, J = 8.1, 7.3, 4H) ppm.1H NMR (400 MHz, Pyridine-d5) δ = 8.96 (s, 2H), 7.84
1.0 Hz, 1H), 6.52 (ddd, J = 8.2, 1.2, 0.7 Hz, 1H) ppm. 13C NMR (d, J = 7.5 Hz, 4H), 7.76 (d, J = 7.0 Hz, 4H), 7.70 (d, J = 8.6 Hz,
(101 MHz, CDCl3) δ = 163.8, 161.9, 149.3, 148.6, 137.7, 131.8, 2H), 7.45 (d, J = 8.7 Hz, 2H), 7.36 (td, J = 7.5, 1.3 Hz, 4H),
130.1, 128.8, 127.0, 127.0, 126.7, 125.9, 120.9, 119.6, 119.2, 7.31–7.23 (m, 4H), 7.18 (td, J = 7.2, 1.0 Hz, 4H), 7.10–7.01 (m,
117.0, 111.3, 107.5 ppm. 11B NMR (96 MHz, CDCl3) δ = 6.4 (s) 6H) ppm. 13C NMR (101 MHz, Pyridine-d5) δ = 166.0, 162.9,
ppm. HRMS (ESI): C25H16BNO2 [M]+ Calcd: 373.1274, found: 149.4, 145.6, 132.0, 131.2, 129.5, 129.3, 129.1, 127.7, 127.5,
373.1257. 120.3, 115.9, 115.7 ppm. HRMS (ESI): C48H32B2N2O2 [M]+
6a – (2′-(2″-oxyphenyl)-benzothiazolato)-9-borafluorene. Calcd: 690.2656, found: 690.2650.
Compound 6a was synthesized as described for 3a using
9-chloroborafluorene (0.22 g, 1.1 mmol), MeOH (ca. 0.05 mL, Optical properties
1 mmol) and 2-(2′-hydroxyphenyl)-1,3-benzothiazole (248 mg, The UV-vis absorption spectra were recorded using a Hitachi
1.1 mmol). Yellow crystalline solid. Yield 334 mg (78%). 1H UV-2300II spectrometer. The emission spectra of the solutions
NMR (400 MHz, CDCl3) δ = 7.79 (d, J = 8.1, 1H), 7.71 (dt, J = were recorded using a Hitachi F-7000 spectrofluorometer,
7.5, 0.9 Hz, 2H), 7.69 (dd, J = 8.0, 1.7 Hz, 1H), 7.48 (ddd, J = equipped with a photomultiplier detector, calibrated using a

This journal is © The Royal Society of Chemistry 2019 Dalton Trans., 2019, 48, 8642–8663 | 8659
View Article Online

Paper Dalton Transactions

Spectral Fluorescence Standard Kit certified by the BAM (LDH-D-C-375 diode with PDL 800-D controller) were used.
Federal Institute for Materials Research and Testing.57 The The excitation wavelengths of 375 nm for complexes 1a,b–7a,b
measurements were performed at room temperature, accord- and 500 nm for complexes 8a and 8b were selected. For exci-
ing to published procedures.58,59 Suprasil quartz cuvettes tation using a Fianium laser, a continuously variable bandpass
(10.00 mm) were used. 1.5 nm slits were used for absorption, filter from the Delta Optical Thin Film was used. The IRF
and 2.5 nm slits were used for the emission spectra. To elimin- signal measured for this setup has 0.130 ns FWHM. The data
ate any background emissions, the spectrum of pure solvent were collected using SPCM software. The analysis of fluo-
Published on 09 May 2019. Downloaded by Università degli Studi dell'Insubria on 3/3/2021 11:33:48 AM.

was subtracted from the samples’ spectra. qy were determined rescence decays was performed using SPCImage 5.0 Software.
in diluted solutions (A < 0.1 for the longest wavelength band)
by comparison with known standards: coumarin 153 in Cyclic voltammetry
ethanol (c = 4 × 10−6 mol dm−3, qy = 0.544),60 9,10-diphenyl- Cyclic voltammetry measurements were carried out in a dry
anthracene (c = 4 × 10−6 mol dm−3, qy = 0.97) and rhodamine argon atmosphere using a VMP3 potentiostat (Bio-Logic) with
101 in methanol (c = 3 × 10−7 mol dm−3, qy = 1.00).61 The con- a scan rate of 0.1 Vs−1. The solutions of the studied com-
centrations of the solution complexes were in the range of pounds (1 mM) in 0.1 M Bu4NPF6/CH2Cl2 were prepared
0.5–2 × 10−5 mol dm−3 (the concentration was adjusted to inside the argon-filled glovebox and then measured in a
reach a similar absorbance to the absorbance of reference three-electrode electrochemical cell: a gold working electrode
solution at the excitation wavelength). The compounds were (ALS Co. Ltd; 2 mm2), a counter electrode ( platinum wire)
excited at their longest wavelength absorption band and a Ag/Ag+ reference electrode (silver wire in 0.1 M AgNO3/
maximum, with the exception of 8a, which was excited at AN). The potential of the reference electrode versus the ferro-
532 nm: cene redox couple was checked before and after the
experiments.
Fx 1  10Ast nx 2
qyx ¼ qyst   
Fst 1  10Ax nst 2 Thermal analyses
where F is the relative integrated photon flux of sample (x) and Differential scanning calorimetric analyses were conducted by
standard (st), A is the absorbance at the excitation wavelength, differential scanning calorimetry (DSC) (Model DSC 1, Mettler-
and n is the refractive index of used solvents. Toledo) under the flow of nitrogen atmosphere. The cali-
ð ð bration of the instrument was performed using the phase-tran-
I ðλem Þ
F ¼ Ic dλem ¼ dλem sition temperature and the phase-transition enthalpy of
sðλem Þ
indium as reference material. The thermogravimetric analyses
Photon fluxes (F) were calculated by the integration of the (TGA) were performed with a TGA/DSC 1 thermogravimetric
corrected spectra (Ic), obtained by (I) division of intensity of analyzer (Mettler-Toledo). The measured samples were heated
emission spectra by the spectral responsivity (s) in corres- at the heating rate of 5 K min−1 from 25 °C to 300/500 °C. The
ponding wavelengths (λem). All measurements were carried out melting points (Tm) were determined from the DSC thermo-
at room temperature. grams based on the onset temperatures. The decomposition
The fluorescence spectra (corrected) and excitation spectra temperatures were determined based on TGA analyses where
of solid samples ( powders and thin films) were measured >5 wg% mass loss was observed.
using an Edinburgh Instruments FLS980 fluorescence spectro-
meter using front face geometry. The quantum yield of the X-ray structural measurements and refinement details
fluorescence measurements was measured using an The single crystals of samples 1a–8a and 8b were obtained by
Edinburgh Instruments integrating sphere with BENFLEC the slow evaporation of CHCl3 (1a–8a) or acetone (8b) solu-
inside coating, according to known procedures with correction tions. They were measured at 100 K on a SuperNova diffract-
for indirect excitation.62 Solid samples were excited at the ometer equipped with an Atlas detector (Cu-Kα radiation, λ =
wavelength corresponding to the absorption maximum 1.54184 Å). Data reduction and analysis were carried out with
observed in solution. Thin films were deposited on quartz the CrysAlisPro program.63 All structures were solved by
plates by spin coating (1500 rpm, 30 s) of the solutions formed direct methods using SHELXS-9764 and refined using
by the samples in CHCl3 (5 mg mL−1). UV-vis absorption SHELXL-2014.65 All important crystallographic data including
spectra of thin films were measured using a Varian Cary 5000 measurement, reduction, structure solution and refinement
spectrophotometer. details were placed in the ESI (Table S1†). Crystallographic
The fluorescence lifetime measurements were acquired Information Files (CIFs) were deposited at the Cambridge
using a Becker-Hickl TSCPC module Simple-Tau 150. The Crystallographic Data Centre as supplementary publications
emitted photons were collected using a multiwavelength detec- no. 1905165 (1a), 1905166 (2a), 1905170 (3a), 1905171 (5a),
tion system: a polychromator and a 16-channel TCSPC detector 1905169 (6a), 1905167 (7a), 1905168 (8a), 1529228 (8b).
(PML-16, Becker-Hickl model), which registers 16 individual
fluorescence decays from the 200 nm spectral range (12.5 nm Theoretical calculations
per channel). For the excitation, the supercontinuum laser, a Ab initio calculations were performed using the Gaussian16
Fianium Whitelase Micro and a PicoQuant 375 nm laser diode programme package.66 In the first step, the molecules were

8660 | Dalton Trans., 2019, 48, 8642–8663 This journal is © The Royal Society of Chemistry 2019
View Article Online

Dalton Transactions Paper

optimised in their ground states using B3LYP (DFT)67–70 The total drift mobility (μ) and relative drift mobility (%μi)
method with a 6-311++G(g,p)71 basis set. After geometry optim- of each dimer pairs were obtained from the equations:
ization, the vibrational frequencies were calculated, and the e e X 2
results showed that the optimized structures were stable geo- μ¼ D¼ ri Wi Pi
kB T 6kB T i
metric structures (no imaginary frequencies). The natural tran-
sition orbitals72 were calculated based on optimised molecular ri 2 Wi Pi
%μi ¼ P 2
geometries at the same level of theory. The graphical represen- r i W i Pi
Published on 09 May 2019. Downloaded by Università degli Studi dell'Insubria on 3/3/2021 11:33:48 AM.

tations of all calculated orbitals can be found in the ESI i

(Fig. S40–48†). They were visualized with the Avogadro


programme.73,74 In the next step, the excited state geometries
along with the emission spectra, were obtained with TD-DFT
methods (at CAM-B3LYP75/6-311++G(d,p)) using geometries Conflicts of interest
obtained from ground state optimizations as starting points.
To take into account the conditions of the fluorescence There are no conflicts to declare.
measurements (diluted DCM solution), we carried out the
DFT and TD-DFT calculations in the presence of the solvent
(DCM) with the polarizable continuum model (PCM) using
the CPCM polarizable conductor calculation model.76 The Acknowledgements
result of the τP1–P4-constrained optimisation energy scan (the
This work was financed by the National Science Centre
τP1–P4 angle varies from 0° to 90° in steps of 10° with all other
(Narodowe Centrum Nauki, Grant No. UMO-2015/19/D/ST5/
parameters fully optimised) is discussed in detail in the ESI
00735). The X-ray measurements were recorded at the
(Fig. S51†).
Crystallographic Unit of the Physical Chemistry Laboratory,
The charge transport rate of the electron (We) and hole (Wh)
Chemistry Department, University of Warsaw. We would like to
can be calculated using the Marcus–Hush theory:
thank prof. Krzysztof Woźniak from the University of Warsaw.
 12   The authors thank the Wroclaw Centre for Networking and
Ve=h 2 π Λe=h
We=h ¼ exp  Supercomputing (http://www.wcss.pl), grant No. 285. for pro-
ℏ λkB T 4kB T
viding computer facilities.
where Λe and Λh are the reorganization energy for electron and
hole, respectively, T is the temperature and Ve/h is the elec-
tronic coupling between the donor and acceptor calculated
from the orbital splitting energies according to Koopmans’ References
approximation.77
The reorganization energy for electrons (Λe) and holes (Λh) 1 D. Li, H. Zhang and Y. Wang, Chem. Soc. Rev., 2013, 42,
are given by the expressions: 8416–8433.
2 D. Frath, J. Massue, G. Ulrich and R. Ziessel, Angew. Chem.,
Λe ¼ ðE0 GSðÞ  E0 GSð0Þ Þ þ ðE GSð0Þ  E GSðÞ Þ Int. Ed., 2014, 53, 2290–2310.
3 Y.-L. Rao and S. Wang, Inorg. Chem., 2011, 50, 12263–
Λe ¼ ðE0 GSðþÞ  E0 GSð0Þ Þ þ ðEþ GSð0Þ  Eþ GSðþÞ Þ 12274.
where E0GS(0), E+GS(0) and E−GS(0) are the energies of the neutral, 4 D. Li, H. Zhang, C. Wang, S. Huang, J. Guo and Y. Wang,
cationic and anionic states with the optimized geometry of J. Mater. Chem., 2012, 22, 4319–4328.
neutral species, respectively. E0GS(+) and E+GS(+) are the energies 5 M. Santra, H. Moon, M.-H. Park, T.-W. Lee, Y. K. Kim and
of neutral and cationic states with the optimized geometry of K. H. Ahn, Chem. – Eur. J., 2012, 18, 9886–9893.
the cationic species. E0GS(−) and E−GS(−) are the energies of 6 H.-Y. Chen, Y. Chi, C.-S. Liu, J.-K. Yu, Y.-M. Cheng,
neutral and anionic states with the optimized geometry of the K.-S. Chen, P.-T. Chou, S.-M. Peng, G.-H. Lee, A. J. Carty,
anionic species. S.-J. Yeh and C.-T. Chen, Adv. Funct. Mater., 2005, 15, 567–
In the next step, the diffusion coefficient (D) was evaluated: 574.
7 J. Dobkowski, P. Wnuk, J. Buczyńska, M. Pszona,
1X 2
D¼ r i W i Pi G. Orzanowska, D. Frath, G. Ulrich, J. Massue,
6 i S. Mosquera-Vázquez, E. Vauthey, C. Radzewicz, R. Ziessel
and J. Waluk, Chem. – Eur. J., 2015, 21, 1312–1327.
where ri represents the distance between the centres of mass
8 A. M. Grabarz, B. Jędrzejewska, A. Zakrzewska, R. Zaleśny,
of two neighbouring molecules and Pi is the hopping prob-
A. D. Laurent, D. Jacquemin and B. Ośmiałowski, J. Org.
ability:
Chem., 2017, 82, 1529–1537.
Ni Wi 9 B. Jędrzejewska, A. Skotnicka, A. D. Laurent, M. Pietrzak,
Pi ¼ P
Ni Wi D. Jacquemin and B. Ośmiałowski, J. Org. Chem., 2018, 83,
i
7779–7788.

This journal is © The Royal Society of Chemistry 2019 Dalton Trans., 2019, 48, 8642–8663 | 8661
View Article Online

Paper Dalton Transactions

10 Q. Wu, M. Esteghamatian, N.-X. Hu, Z. Popovic, G. Enright, 34 H. Höpfl, V. Barba, G. Vargas, N. Farfan, R. Santillan and
Y. Tao, M. D’Iorio and S. Wang, Chem. Mater., 2000, 12, 79– D. Castillo, Chem. Heterocycl. Compd., 1999, 35, 912–927.
83. 35 J. Ugolotti, S. Hellstrom, G. J. P. Britovsek, T. S. Jones,
11 Y. Hong, J. W. Y. Lam and B. Z. Tang, Chem. Soc. Rev., 2011, P. Hunt and A. J. P. White, Dalton Trans., 2007, 1425–1432.
40, 5361–5388. 36 M. H. Kolář and P. Hobza, Chem. Rev., 2016, 116, 5155–
12 K. Durka, I. Głowacki, S. Luliński, B. Łuszczyńska, 5187.
J. Smętek, P. Szczepanik, J. Serwatowski, U. E. Wawrzyniak, 37 K. Durka, K. Gontarczyk, S. Luliński, J. Serwatowski and
Published on 09 May 2019. Downloaded by Università degli Studi dell'Insubria on 3/3/2021 11:33:48 AM.

G. Wesela-Bauman, E. Witkowska, G. Wiosna-Sałyga and K. Woźniak, Cryst. Growth Des., 2016, 16, 4292–4308.
K. Woźniak, J. Mater. Chem. C, 2015, 3, 1354–1364. 38 M. E. Brezgunova, J. Lieffrig, E. Aubert, S. Dahaoui,
13 Z. Zhang, Z. Zhang, K. Ye, J. Zhang, H. Zhang and Y. Wang, P. Fertey, S. Lebègue, J. G. Ángyán, M. Fourmigué and
Dalton Trans., 2015, 44, 14436–14443. E. Espinosa, Cryst. Growth Des., 2013, 13, 3283–3289.
14 M. F. Smith, S. J. Cassidy, I. A. Adams, M. Vasiliu, 39 F. Zhou, R. Liu, P. Li and H. Zhang, New J. Chem., 2015, 39,
D. L. Gerlach, D. A. Dixon and P. A. Rupar, Organometallics, 1611–1618.
2016, 35, 3182–3191. 40 F. Würthner, T. E. Kaiser and C. R. Saha-Möller, Angew.
15 A. Iida, A. Sekioka and S. Yamaguchi, Chem. Sci., 2012, 3, Chem., Int. Ed., 2011, 50, 3376–3410.
1461–1466. 41 M. Guerrini, C. Cocchi, A. Calzolari, D. Varsano and
16 A. Hübner, A. M. Diehl, M. Bolte, H.-W. Lerner and S. Corni, J. Phys. Chem. C, 2019, 123, 6831–6838.
M. Wagner, Organometallics, 2013, 32, 6827–6833. 42 F. S. Santos, E. Ramasamy, V. Ramamurthy and
17 C. Bonnier, W. E. Piers, A. Al-Sheikh Ali, A. Thompson and F. S. Rodembusch, J. Photochem. Photobiol., A, 2016, 317,
M. Parvez, Organometallics, 2009, 28, 4845–4851. 175–185.
18 S. Yamaguchi, T. Shirasaka, S. Akiyama and K. Tamao, 43 J. Jayabharathi, V. Thanikachalam and K. Jayamoorthy,
J. Am. Chem. Soc., 2002, 124, 8816–8817. Photochem. Photobiol. Sci., 2013, 12, 1761–1773.
19 C. J. Berger, G. He, C. Merten, R. McDonald, M. J. Ferguson 44 Z. Zhang, Z. Zhang, H. Zhang and Y. Wang, Dalton Trans.,
and E. Rivard, Inorg. Chem., 2014, 53, 1475–1486. 2018, 47, 127–134.
20 Z. Zhang, H. Zhang, C. Jiao, K. Ye, H. Zhang, J. Zhang and 45 Y. Qi, R. Kang, J. Huang, W. Zhang, G. He, S. Yin and
Y. Wang, Inorg. Chem., 2015, 54, 2652–2659. Y. Fang, J. Phys. Chem. B, 2017, 121, 6189–6199.
21 K. Yuan, X. Wang, S. K. Mellerup, I. Kozin and S. Wang, 46 P. Krishnamoorthy, B. Ferreira, C. S. B. Gomes, D. Vila-Viçosa,
J. Org. Chem., 2017, 82, 13481–13487. A. Charas, J. Morgado, M. J. Calhorda, A. L. Maçanita and
22 B. M. Bell, T. P. Clark, T. S. De Vries, Y. Lai, D. S. Laitar, P. T. Gomes, Dyes Pigm., 2017, 140, 520–532.
T. J. Gallagher, J.-H. Jeon, K. L. Kearns, T. McIntire, 47 D. Li, K. Wang, S. Huang, S. Qu, X. Liu, Q. Zhu,
S. Mukhopadhyay, H.-Y. Na, T. D. Paine and A. A. Rachford, H. Zhang and Y. Wang, J. Mater. Chem., 2011, 21, 15298–
Dyes Pigm., 2017, 141, 83–92. 15304.
23 D.-G. Chen, R. Ranganathan, J.-A. Lin, C.-Y. Huang, 48 R. A. Marcus and N. Sutin, Biochim. Biophys. Acta, Rev.
M.-L. Ho, Y. Chi and P.-T. Chou, J. Phys. Chem. C, 2019, Biomembr., 1985, 811, 265–322.
123, 4022–4028. 49 P. F. Barbara, T. J. Meyer and M. A. Ratner, J. Phys. Chem.,
24 M. Urban, P. Górka, K. Nawara, K. Woźniak, K. Durka and 1996, 100, 13148–13168.
S. Luliński, Dalton Trans., 2018, 47, 15670–15684. 50 E. Laborda, M. C. Henstridge, C. Batchelor-McAuley and
25 P. A. Chase, W. E. Piers and B. O. Patrick, J. Am. Chem. Soc., R. G. Compton, Chem. Soc. Rev., 2013, 42, 4894–4905.
2000, 122, 12911–12912. 51 M. Mamada, H. Fujita, K. Kakita, H. Shima, Y. Yoneda,
26 I. Gorokhovik, S. Rieder, G. Povie and P. Renaud, ARKIVOC, Y. Tanaka and S. Tokito, New J. Chem., 2016, 40, 1403–
2014, 2014, 274–286. 1411.
27 A. Iida and S. Yamaguchi, J. Am. Chem. Soc., 2011, 133, 52 D. Li, Y. Yuan, H. Bi, D. Yao, X. Zhao, W. Tian, Y. Wang and
6952–6955. H. Zhang, Inorg. Chem., 2011, 50, 4825–4831.
28 S. Luliński, J. Smętek, K. Durka and J. Serwatowski, 53 A. Wakamiya, T. Taniguchi and S. Yamaguchi, Angew.
Eur. J. Org. Chem., 2013, 8315–8322. Chem., Int. Ed., 2006, 45, 3170–3173.
29 S. Brend’amour, J. Gilmer, M. Bolte, H.-W. Lerner and 54 R. Filler, A. E. Fiebig and M. Y. Pelister, J. Org. Chem., 1980,
M. Wagner, Chem. – Eur. J., 2018, 24, 16910–16918. 45, 1290–1295.
30 M. Urban, K. Durka, P. Jankowski, J. Serwatowski and 55 Y.-H. Cho, C.-Y. Lee and C.-H. Cheon, Tetrahedron, 2013,
S. Luliński, J. Org. Chem., 2017, 82, 8234–8241. 69, 6565–6573.
31 G. Wesela-Bauman, M. Urban, S. Luliński, J. Serwatowski 56 H. Kadri, C. S. Matthews, T. D. Bradshaw, M. F. G. Stevens
and K. Woźniak, Org. Biomol. Chem., 2015, 13, 3268–3279. and A. D. Westwell, J. Enzyme Inhib. Med. Chem., 2008, 23,
32 N. G. Kim, C. H. Shin, M. H. Lee and Y. Do, J. Organomet. 641–647.
Chem., 2009, 694, 1922–1928. 57 D. Pfeifer, K. Hoffmann, A. Hoffmann, C. Monte and
33 Z. Zhang, H. Bi, Y. Zhang, D. Yao, H. Gao, Y. Fan, U. Resch-Genger, J. Fluoresc., 2006, 16, 581–587.
H. Zhang, Y. Wang, Y. Wang, Z. Chen and D. Ma, Inorg. 58 C. Würth, M. Grabolle, J. Pauli, M. Spieles and U. Resch-
Chem., 2009, 48, 7230–7236. Genger, Nat. Protoc., 2013, 8, 1535–1550.

8662 | Dalton Trans., 2019, 48, 8642–8663 This journal is © The Royal Society of Chemistry 2019
View Article Online

Dalton Transactions Paper

59 U. Resch-Genger and K. Rurack, Pure Appl. Chem., 2013, 85, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin,
2005–2013. K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador,
60 K. Rurack and M. Spieles, Anal. Chem., 2011, 83, 1232– J. J. Dannenberg, S. Dapprich, A. D. Daniels, Ö. Farkas,
1242. J. B. Foresman, J. V. Ortiz, J. Cioslowski and D. J. Fox,
61 A. M. Brouwer, Pure Appl. Chem., 2011, 83, 2213–2228. Gaussian 09, Gaussian, Inc., Wallingford, CT, USA,
62 J. C. de Mello, H. F. Wittmann and R. H. Friend, Adv. 2009.
Mater., 1997, 9, 230–232. 67 A. D. Becke, J. Chem. Phys., 1993, 98, 5648–5652.
Published on 09 May 2019. Downloaded by Università degli Studi dell'Insubria on 3/3/2021 11:33:48 AM.

63 CrysAlis Pro Software, Oxford Diffraction Ltd, Oxford, U.K., 68 C. Lee, W. Yang and R. G. Parr, Phys. Rev. B: Condens.
2010. Matter Mater. Phys., 1988, 37, 785–789.
64 G. M. Sheldrick, Acta Crystallogr., Sect. A: Found. 69 S. H. Vosko, L. Wilk and M. Nusair, Can. J. Phys., 1980, 58,
Crystallogr., 2008, 64, 112–122. 1200–1211.
65 G. M. Sheldrick, Acta Crystallogr., Sect. C: Struct. Chem., 70 P. J. Stephens, F. J. Devlin, C. F. Chabalowski and
2015, 71, 3–8. M. J. Frisch, J. Phys. Chem., 1994, 98, 11623–11627.
66 M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, 71 R. Krishnan, J. S. Binkley, R. Seeger and J. A. Pople,
M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, J. Chem. Phys., 1980, 72, 650–654.
B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, 72 R. L. Martin, J. Chem. Phys., 2003, 118, 4775–4777.
X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, 73 Avogadro: an open-source molecular builder and visualization
J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, tool. Version 1.2.0. http://avogadro.cc/.
J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, 74 M. D. Hanwell, D. E. Curtis, D. C. Lonie,
H. Nakai, T. Vreven, J. A. Montgomery Jr., J. E. Peralta, T. Vandermeersch, E. Zurek and G. R. Hutchison,
F. Ogliaro, M. J. Bearpark, J. Heyd, E. N. Brothers, J. Cheminf., 2012, 4, 17.
K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, 75 T. Yanai, D. P. Tew and N. C. Handy, Chem. Phys. Lett.,
K. Raghavachari, A. P. Rendell, J. C. Burant, S. S. Iyengar, 2004, 393, 51–57.
J. Tomasi, M. Cossi, N. Rega, N. J. Millam, M. Klene, 76 J. Tomasi, B. Mennucci and R. Cammi, Chem. Rev., 2005,
J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, 105, 2999–3094.
R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, 77 T. Koopmans, Physica, 1934, 1, 104–113.

This journal is © The Royal Society of Chemistry 2019 Dalton Trans., 2019, 48, 8642–8663 | 8663

You might also like