Journal Pre-proof Towards better understanding of photophysical properties of rhenium(I) tricarbonyl complexes with terpy-like ligands
Magdalena Małecka, Barbara Machura, Anna Świtlicka, Sonia Kotowicz, Grażyna Szafraniec-Gorol, Mariola Siwy, Marcin Szalkowski, Sebastian Maćkowski, Ewa Schab-Balcerzak PII:
S1386-1425(20)30101-3
DOI:
https://doi.org/10.1016/j.saa.2020.118124
Reference:
SAA 118124
To appear in:
Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy
Received date:
2 January 2020
Revised date:
31 January 2020
Accepted date:
1 February 2020
Please cite this article as: M. Małecka, B. Machura, A. Świtlicka, et al., Towards better understanding of photophysical properties of rhenium(I) tricarbonyl complexes with terpylike ligands, Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy(2020), https://doi.org/10.1016/j.saa.2020.118124
This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
© 2020 Published by Elsevier.
Journal Pre-proof Towards better understanding of photophysical properties of rhenium(I) tricarbonyl complexes with terpy-like ligands
Magdalena Małecka a, Barbara Machura* a, Anna Świtlickaa, Sonia Kotowicza, Grażyna Szafraniec-Gorola, Mariola Siwyb, Marcin Szalkowskic, Sebastian Maćkowskic, Ewa
of
Schab-Balcerzak*a,b
Institute of Chemistry, University of Silesia, 9th Szkolna Street, 40006 Katowice, Poland
b
Centre of Polymer and Carbon Materials, Polish Academy of Sciences, 34 M.
ro
a
c
-p
Curie-Sklodowska Str., 41-819 Zabrze, Poland
Nanophotonics Group, Institute of Physics, Faculty of Physics, Astronomy and Informatics,
re
Nicolaus Copernicus University, 5 Grudziadzka Str., 87-100 Torun, Poland
na
lP
*Corresponding authors: e-mail:
[email protected],
[email protected]
Jo ur
Abstract: Series of Re(I) carbonyls complexes were designed and synthesized to explore the impact of the triimine skeleton and number of methoxy groups attached to aryl substituents on their optoelectronic and thermal properties. The chemical structures of the prepared complexes were confirmed by 1H and 13C NMR spectroscopy, HR-MS, elemental anlsysis, and X-ray measurements. DSC measuremtns showed that they melted in the range of 198–325 C. Some of them form stable molecular glasses with high glass transition temperatures (158–173C). Experimentally obtained optical properties were supported by DFT calculations. The UV-Vis spectra display a series of overlapping absorption bands in the range 200–350 nm, and much weaker broad band in the visible spectral region, due to intraligand and charge transfer transitions, respectively. All synthesized complexes were emissive in solution and in solid state as powder. Moreover, when applied in diodes, some of them exhibited ability for emission of light under external voltage with maximum of electroluminescence band located at 591–630 nm. Keywords: Re(I) carbonyl • terpy-like ligands • photoluminescence • electroluminescence
Journal Pre-proof Introduction Re(I) carbonyls [ReL(CO)3(NN)]n+ (n = 0 or 1, L - ancillary ligand) with derivatives of 1,10-phenanthroline and 2,2-bipyridine (NN) have received widespread scientific attention since 1974, when Wrighton and Morse first examined the luminescence behavior of [ReCl(CO)3(4,7-(Ph)2phen)] and [ReCl(CO)3(5-R-phen)] (R – H, CH3, Cl, Br, NO2) [1]. It has been demonstrated that the character and energies of HOMO and LUMO, and hence photo- and electroluminescent properties of [ReL(CO)3(NN)]n+, may be fine-tuned by varying the ancillary ligand (L) and introducing electron withdrawing or donating groups into the diimine skeleton [2-42]. The development in this area has been mostly driven by the potential
of
application of [ReL(CO)3(NN)]n+ in photocatalysis [43-47], biological imaging [48-57], and
ro
organic light-emitting devices [14, 58-65].
The related [ReCl(CO)3(terpy-κ2N)] compound, first examined in 1988 by Juris et al. [66],
-p
was found to be non-emissive in solution at room temperature, which limited further studies of Re(I) tricabonyls with 2,2′:6′,2′′-terpyridines. Scientific interest in these systems was renewed
re
in the early 2000s [67-72], especially when Wang and co-workers demonstrated that
lP
introduction of hole-transporting moieties, carbazole or diphenylamine-based substituents, into 2,2′:6′,2′′-terpyridine at 4-position, leads to significantly enhanced phosphorescence of
na
[ReCl(CO)3(4′-R-terpy-κ2N)] compared to [ReCl(CO)3(terpy-κ2N)] [70]. Additionally, theoretical investigations based on the density functional theory (DFT) performed for these compounds, revealed that they facilitate electron transfer and hole extraction, making them
Jo ur
very attractive as electrophosphorescence materials for organic light emitting diodes (OLEDs) construction [73]. Potential usefulness of [ReCl(CO)3(L-2N)] with modified terpyridine-like ligands for use in the OLED technology was further verified by the research results obtained in our group [74-79].
In the current work, series of phenyl and naphthyl-substituted 2,2′:6′,2′′-terpyridine (terpy), 2,6-di(thiazol-2-yl)pyridine (dtpy) and 2,6-di(pyrazin-2-yl)pyridine (dppy) derivatives decorated with methoxy groups (scheme 1) were used for the synthesis of [ReCl(CO)3(L-2N)]. Beneficial role of a single methoxy group attached to aryl substituents was proved by our previous studies. In solid state, the Re(I) complex with 2,2′:6′,2′′-terpyridine substituted with 4-methoxy-1-phenyl exhibited a remarkably enhanced photoluminescence quantum yield of ∼30% in relation to other [ReCl(CO)3(4-R-terpy-κ2N)] with 4-Cl-phenyl (∼21%), 4-Br-phenyl (∼17%),
4-biphenyl
(∼12%)
substituents
[75,78].
The
Re(I)
complexes
with
Journal Pre-proof 2,2′:6′,2′′-terpyridine or 2,6-di(thiazol-2-yl)pyridine functionalized with 4-methoxy-1-phenyl and 6-methoxy-2-naphthyl were also found to be very promising for OLED fabrication [78]. The aim of the current was to explore the impact of additional methoxy groups attached to the phenyl or naphthyl-substituent of terpy, dtpy and dppy on the thermal, spectral and electrochemical properties of the rhenium(I) carbonyls [ReCl(CO)3(Ln-2N)] (scheme 1). As the methoxy unit is known to lower oxidation potential and enhance the electrochemical stability, its presence in compounds designed for optoelectronic can be beneficial [80]. The presence of methoxy substituent at the para position of the phenyl ring adds electron density into the triimine skeleton, while meta-positioned methoxy groups withdraw electron density
of
from the trisheterocyclic moiety [81-82].
ro
The photophysical properties of 1–9 were explored by carrying out absorption and emission studies in solution and solid state, as well as a possibility of light emission under
-p
external voltage in diodes with configuration ITO/PEDOT:PSS/compound/Al and
Jo ur
na
lP
compounds was investigated using DSC.
re
ITO/PEDOT:PSS/PVK:PBD:compound/Al was tested. The thermal behavior of the prepared
Journal Pre-proof O
O
H3C
O CH3
H3C
N L
N L
CH3
CH3
O
O
O
H3C
CH3
3
O
CH3
O
H3C
CH3
S
4
N
L
7
N
re
5
L
lP
O
na
N
N
N
CH3
O
6
CH3
O
O
O
CH3
CH3
CH3
Jo ur
N
N
N
L
CH3
O
N
-p
N
N
L
S
N
N N
ro
of
H3C
N L
O
O
N
N
2
CH3
O
N
N
1
O
CH3
N
S
N
O
H3C
CH3
S
N N
O
O
S
S
N N
N
N L
8
N
N
N
N L
9
Scheme 1. Ligands employed in this study.
Results and Discussion Synthesis, structural and thermal characterisation The complexes [ReCl(CO)3(Ln-2N)] (1–9) were prepared according to the standard procedure via direct exchange of two carbonyl ligands in Re(CO)5Cl by the corresponding L1-9 ligand (Scheme 1). The molecular formulae of 1–9 were confirmed by elemental analysis, HRMS (ESI) spectrometry, as well as NMR and IR spectroscopy. In the IR spectra, a sharp and intense high-energy carbonyl stretching band (2019–2029 cm–1) accompanied with two
Journal Pre-proof overlapping lower-energy (CO) absorptions (1955–1864 cm–1) (Figures S1-S9) is supportive of facial arrangement of CO groups in the moiety [Re(CO)3]+. Consistent with the bidentate coordination mode of Ln, the protons of the outer pirydyl/thiazolyl/pyrazinyl rings are magnetically inequivalent, showing separate signals in the 1H NMR spectra of 1–9, each integrated for 1 proton (Figures S10S19). Using the multidimensional techniques 1H–13C HMBC, 1H–13C HMQC, 1H–1H COSY, the full assignment of the signals in the 1H and
13
C
NMR spectra was performed for 1, 4, 7 (Figure S20). Crystals suitable for X-ray diffraction studies were obtained for the complexes 5, 7 and 9, and the structural analysis confirmed distorted octahedral geometry of Re(I) ion, defined by three
na
lP
re
-p
ro
of
fac-arranged carbonyl groups, two nitrogen atoms of Ln, and one chlorine atom (Figure 1).
7
9
Jo ur
5
Figure 1. Molecular structures of 5, 7 and 9 with thermal ellipsoids set at 50% probability for non-hydrogen atoms.
As expected for [ReCl(CO)3(4′-R-terpy-κ2N)] and related systems [74-79], the complexes 5, 7 and 9 show a small bite angle N(2)–Re(1)–N(1) angle of 74.91(12) in 5, 74.87(11) in 7, 74.50 (13)° in 9, significant enlargement of C(1)–Re(1)–N(2) angle of 100.37(16) in 5, 99.98(15) in 7 and 102.53(17)° in 9 as well as elongation of Re(1)–N(2) bond length to the central pyridine ring [2.250(3)Å for 5, 2.204(3)Å for 7, 2.207(3) Å for 9] in comparison with Re(1)–N(1) distance to the peripheral pyridine ring [2.165(3)Å in 5, 2.165(3)Å in 7, 2.148(4) Å in 9]. All these structural features are attributed to κ2N-coordination of the tiiminie moiety and steric interaction of the uncoordinated heterocycle ring with the carbonyl group C(1)–O(1). The non-coordinated peripheral ring is inclined to the central pyridine at 41.622(3) in 5, 50.96(15) in 7 and 54.69(17) in 9, while the dihedral angle between the central pyridine and pendant
Journal Pre-proof phenyl ring ring is 10.4465(7)° in 5, 19.95(14)° in 7 and 15.19(14) in 9. Additional structural data of 5, 7 and 9 are included in Tables S1–S5 and shown in Figure S21 (in ESI). Thermal properties of the prepared 1–9 complexes were investigated using differential scanning calorimetry (DSC). The registered temperatures of melting (Tm), crystallization (Tc), and glass transition (Tg) are presented in Experimental section. During the first heating scan in DSC thermograms only the melting endotherm with maximum between 198 and 325 C was seen, except for complex 9. In the case of 9, apart from Tm at 233 C, also crystallization exotherm giving new crystals, which melted with degradation at 287 C were observed. It was found that the presence of three methoxy substituents raises the melting temperature of the
of
complexes (4–6). Moreover, these complexes (4–6) melted with thermal decomposition. Other
ro
complexes, that is, with two methoxy groups (1, 3, 7) formed amorphous material after rapid cooling with Tg in the range of 158 – 173 C, which is beneficial for formation of
-p
morphologically stable thin films. Complexes 3 and 7 form stable molecular glasses, because they do not undergo crystallization even upon heating above their Tg, contrary to 1, which
re
during further heating above Tg crystallized and again formed crystalline material. The results
lP
indicate that all of the investigated complexes exhibited high enough Tm values from the point
na
of view of applications in optoelectronic devices.
DFT evaluation of the electronic structures and transport properties To gain better understanding of the photophysical properties of 1–9, the frontier molecular
Jo ur
orbitals and their energy levels have been calculated at DFT/PBE1PBE/DEF2– TZVPD/6-31+G** level theory, and the results are presented in Figures 2-3 and S22-S24. Geometry optimization yielded bond angles and lengths in good agreement with crystal structures (Table S1). For all of the compounds, the LUMO and LUMO+1 are composed predominantly of π* orbitals localized on the triimine skeleton, and the energies of these orbitals are essentially influenced by the type of the triimine ligand, decreasing in the order terpy dtpy dppy. The HOMOs of 7–9 are energetically destabilized relative to the respective 1–6 and previously reported Re(I) tricarbonyls with the corresponding triimines substituted with 4-methoxy-1-naphthyl [78]. The HOMO energy levels of 7–9, raised by the attachment of 4,7-dimethoxy-1-naphthyl unit to the triimine moiety (terpy, dtpy and dppy) seem to be favorable for hole-transporting ability of the resulting complexes. On the other hand, decrease of the LUMO energy levels of the Re(I) complexes with dtpy- and dppy-based ligands relative to those bearing terpy derivatives is expected to improve the electron injection ability [83]. On
Journal Pre-proof the contrary, introduction of additional methoxy groups into the phenyl group attached to the central pyridine of terpy, dtpy and dppy leads to stabilization of the HOMO energy level of 1– 6 relative to the corresponding Re(I) tricarbonyls with 4-MeO-phenyl-subsitued triimines, which results in higher HOMO–LUMO energy gaps. Structural modifications of the triimine ligand also tune the distribution of the HOMO orbitals, which can result in differences of electronic transition character upon excitation. The HOMOs of 1, 3 and 7–9 are principally centered on the substituent part of the organic ligand, while 5d rhenium, *CO and Cl make a major contribution into the HOMO of 2. The HOMOs of 4, 5
Jo ur
na
lP
re
-p
ro
of
and 6 spread over the whole molecule, except the peripheral triimine ring orbitals.
Figure 2. Partial molecular orbital energy DFT/PBE1PBE/DEF2–TZVPD/6-31+G** level.
levels
of
1–9
computed
at
the
Journal Pre-proof Compound
1
2
3
4
5
6
7
HOMO
f o
LUMO
n r u
l a
Figure 3. Frontier molecular orbitals of the complexes 1–9.
o J
r P
e
o r p
8
9
Journal Pre-proof The values of ionization potential (IP) and electronic affinity (EA) were computed to estimate energy barriers for injecting holes and electrons, while reorganization energies (λhole and λelectron), which reflect the geometric relaxation associated with transition between neutral and ionized states, were calculated to estimate the charge transport properties of 1–9 (Table 1). For photoluminescent materials, lower IP values mean the easier injection of holes from the hole-transporting layer to the emitter. On the contrary, the higher EA, the easier the entrance of electrons from the electron-transporting layer. The low reorganization energies are required for efficient charge transporting process [84-85].
of
Compared to the Re(I) tricarbonyls bearing the terpy/dtpy/dppy ligands substituted with 4-methoxy-1-naphthyl [78], the compounds 7–9 showed lower IP values, which may
ro
imply improved hole injection abilities. The reorganization energies (λhole and λelectron)
-p
were found to be almost unchanged after the introduction of additional methoxy groups into the naphthyl ring. Most importantly however, the values of λhole and λelectron are
re
comparable, which may make the complexes 7–9 good ambipolar materials, likewise as systems with terpy/dtpy/dppy substituted with 4-methoxy-1-naphthyl [78]. The
lP
introduction of additional methoxy groups into the phenyl group attached to the terpy, dtpy and dppy effectively lowers the values of λhole, but it was found unfavourable for
na
hole-transporting ability of 1–6, leading to the increase of IP values relative to the Re(I) tricarbonyls bearing the corresponding triimines with 4-methoxy-1-phenyl substituent
Jo ur
[78]. For all three series 1–3, 4–6 and 7–9, the calculated EAs and EEPs values increase in the following order terpy dtpy dppy.
Table 1. Calculated ionization potentials and electron affinities (vertical and adiabatic), energy gap, as well as hole and electrons reorganization energies and extraction potentials (DFT/PBE1PBE/DEF2-TZVPD/DEF2-TZVP) of 1–9 in acetonitrile. Complex
IP(v) [eV]
IP(a) [eV]
EA(v) [eV]
EA(a) [eV]
λhole [eV]
λelectron [eV]
HEP [eV]
EEP [eV]
energy gap(a) [eV]
1
6.26
5.93
2.79
2.96
0.71
0.34
5.55
3.13
2.96
2
6.30
5.97
3.04
3.21
0.70
0.33
5.60
3.38
2.76
3
6.40
6.09
3.19
3.36
0.68
0.32
5.72
3.52
3.52
4
6.26
5.92
2.81
2.97
0.72
0.33
5.54
3.14
2.95
5
6.29
5.96
3.03
3.20
0.49
0.34
5.59
3.37
2.76
Journal Pre-proof 6
6.33
5.84
3.19
3.35
0.77
0.32
5.56
3.51
2.49
7
5.77
5.63
2.74
2.92
0.27
0.36
5.49
3.10
2.71
8
5.81
5.68
2.98
3.16
0.27
0.35
5.54
3.34
2.52
9
5.81
5.67
3.15
3.32
0.27
0.34
5.54
3.49
2.35
;
;
;
Absorption properties and TDDFT calculations Absorption properties of 1–9 were investigated in solid state and two solvents of
of
different polarity, chloroform (ε = 4.8) and acetonitrile (ε = 37.5) (Table S6 and Figures S25-S27). The absorption profiles of 1–9 in acetonitrile are shown in Figure 4. The
ro
electronic spectra of 1–9 display a series of overlapping intense absorption bands in the
-p
range 200–350 nm, which is typical of intraligand transitions, and much weaker and
Jo ur
na
lP
re
broader bands extending well into the visible region (Figure 4).
(a)
(b) Figure 4.The UV-Vis absorption spectra for 1–9 in (a) acetonitrile and (b) thin films on glass. Right y axis in Fig. 4b is referred to compounds 2 and 5. According to the TDDFT calculations, the low-energy absorption bands of 1–9 are composed of several transitions of different origins (Table 2 and Figure S28). In addition to the metal-to-ligand charge transfer (1MLCT) and ligand-to-ligand charge transfer (1LLCT), assigned as metal-ligand-to-ligand charge transfer (1MLLCT) transitions, also intraligand charge transfer (1ILCT) transitions make contributions into
Journal Pre-proof the visible absorption spectra. The latter ones are the prevailing components for compounds bearing 4,7-dimethoxy-1-naphthyl substituent (7–9). The visible absorption spectra of these compounds (7–9) are principally contributed by HOMOLUMO and HOMO LUMO+1 transitions, which can be interpreted as charge delocalization from the 4,7-dimethoxy-1-naphthyl unit to π-conjugated trisheterocyclic acceptor moiety (1ILCT). For complexes 1–6, the transitions the transitions assigned to the lowest wavelength absorption are of 1MLLCT origin or mixed 1MLLCT/1ILCT character (Table 3). With respect to the corresponding Re(I) tricarbonyls with metoxyphenyl-substituted
of
triimines (1–6), the lowest energy absorption bands of 7–9 are bathochromically
ro
shifted, consistent with extension of π conjugation of the substituent. Conversely, substituent pattern (1–3 in relation to the corresponding 4–6) has little impact on the
-p
location of the longest-wavelength absorption. The effect of the triimine skeleton is seen in the bathochromic shift of the lowest energy absorption band in the order terpy
re
tpy dppy for all three series 1–3, 4–6 and 7–8 (Figure S27). The absorption profiles of 1–9 are also affected by the changes in the solvent polarity. Typically for Re(I)
lP
carbonyls[ReL(CO)3(NN)]n+, the longest wavelength absorption band in the spectra of
na
1–9 is negatively solvatochromic, with maxima blue-shifted of 25 nm for 1, 22 nm for 2, 24 nm for 3, 17 nm for 4, 28 nm for 5, 21 nm for 6, 12 nm for 7, 17 nm for 8, and 18 nm for 9.
Jo ur
In solid state as powder, the absorption spectra of 1–9 are bathochromically shifted in comparison with those obtained in solution, most likely due to stronger intermolecular interactions in the solid state (Figure S25). On the other hand, complexes as thin film on glass substrate showed absorption range similar to those in solution (Figure 4). Table 2. The energies and characters of spin-allowed electronic transitions assigned to the lowest wavelength absorption band of complexes 1-9 in MeCN. Experimental absorption ; nm/eV
Calculated transitions Major contribution (%)
Character
E [eV]
λ [nm]
Oscillator strength
3.05 3.19 3.24 3.51
406.4 389.0 383.0 353.4
0.0112 0.0046 0.1278 0.0071
1
370 (3.35)
H-1→L (98%) H→L (98%) H-2→L (97%) H-3→L (94%)
MLLCT ILCT MLLCT MLLCT
Journal Pre-proof 2 MLLCT MLLCT/ILCT ILCT MLLCT 3
2.86 3.05 3.21 3.32
434.1 406.5 385.9 373.8
0.0034 0.1426 0.0041 0.0216
MLLCT ILCT MLLCT MLLCT ILCT MLLCT 4
2.77 2.97
447.4 417.1
0.0131 0.0627
401 (3.09)
H-1→L (94%) H→L (51%) H-2→L (45%) H-2→L (48%) H→L (45%) H-4→L (96%)
3.00
413.6
0.0543
3.25
381.1
0.0068
376 (3.30)
H→L (94%) H-1→L (89%) H-3→L (92%)
MLLCT MLLCT/ILCT MLLCT 5
3.05 3.22 3.50
406.8 385.2 353.8
0.0144 0.1851 0.0122
MLLCT ILCT/MLLCT ILCT/MLLCT MLLCT/ILCT MLLCT ILCT MLLCT/ILCT ILCT/MLLCT ILCT MLLCT/ILCT MLCT 6
2.87
432.6
0.0041
3.03
408.5
0.2174
387 (3.20)
H-1→L (73%) H→L (23%) H→L (41%) H-3→L (35%) H-1→L (22%) H-2→L (54%) H-3→L (21%) H→L (19%) H-2→L (44%) H-3→L (35%) H-4→L (86%)
3.26
380.6
0.0475
3.28
378.3
0.0468
3.34
371.1
0.0168
H-1→L (64%) H→L (32%) H-2→L (36%) H→L (32%) H-1→L (29%) H-2→L (57%) H→L (34%)
MLLCT ILCT/MLLCT MLLCT/ILCT ILCT/MLLCT MLLCT MLLCT/ILCT ILCT/MLLCT 7
2.77
447.8
0.0234
2.96
419.2
0.1745
3.14
394.8
0.038
ILCT MLLCT MLLCT/ILCT ILCT ILCT MLCT ILCT MLLCT/ILCT 8
2.86 3.10 3.31
433.3 399.3 374.7
0.2119 0.0171 0.0391
389 (3.19)
H→L (97%) H-1→L (97%) H-2→L (80%) H-3→L (17%) H→L+1 (96%) H-4→L(96%) H-3→L (77%) H-2→L (17%)
3.47 3.55 3.63
357.3 349.2 341.5
0.1386 0.0058 0.0349
402 (3.08)
H→L (98%) H-1→L (98%) H-2→L (78%) H-3→L (19%)
ILCT MLLCT ILCT MLLCT
2.69 2.91 3.13
461.2 426.33 395.65
0.2315 0.011 0.0419
ro
-p
re
lP
na
Jo ur
399 (3.11)
of
391 (3.17)
H→L (98%) H-2→L (78%) H-1→L (87%) H-4→L (83%)
Journal Pre-proof
408 (3.04)
H→L+1 (98%) H-4→L (90%) H-3→L (78%) H-2→L (19%)
ILCT MLLCT/IL ILCT MLLCT 9
3.25 3.37 3.42
381.35 368.08 362.26
0.1356 0.0119 0.0236
H→L (97%) H-1→L (96%) H-2→L (83%) H→L+1 (95%) H-4→L (98%) H-3→L (81%) H→L+2 (95%)
ILCT MLLCT MLLCT ILCT MLLCT ILCT ILCT
2.49 2.76 3.01 3.18 3.24 3.39 3.47
497.2 450.0 412.4 390.0 382.2 365.3 356.9
0.1464 0.0118 0.0607 0.1637 0.0043 0.0247 0.0421
Emission properties
of
Photoluminescence behavior of 1–9 was studied in solution and in solid state as
ro
powder, as well as in the form of a thin film on a glass substrate. The relevant photophysical data are summarized in Table 3, and excitation and emission spectra of
-p
1–9 are shown in Figure S29.
Upon excitation in the lowest-energy absorption band, the Re(I) complexes 1–9 exhibit
re
emission band with the maximum falling in the range 654 – 744 nm in solution, and
lP
580 – 686 in solid state as powder. Upon varying the solvent polarity from chloroform to more polar acetonitrile, the shift of the emission is smaller than 15 nm (Table 3). Consistent with the rigidochromic effect leading to raising the energy of the emissive
na
CT states due the lack of solvent mobility following excitation, the emission maxima of 1–9 in solid state as powder occur in higher energies in relation to the corresponding
Jo ur
solutions. The blue-shift of emission when going from solution to the solid state is accompanied by the increase of lifetimes, which are significantly longer in solid state as compared to those in solution. Upon excitation in the lowest-energy absorption band, the Re(I) complexes 1, 4, 5 and 9 as thin films deposited on glass were not emissive. In the case of compound 3, 6 and 7 a weak emission with the em at 621, 616 and 573 nm, respectively, was observed (cf. Table 3). Table 3. Summary of photoluminescent properties of complexes 1–9. Maximum of excitation wavelength (ex) [nm]
Emission [nm]
Stokes shift [cm-1]
τ[ns]
χ2
φ [%]
MeCN
370, 273
672
12140
1.087
0.5
CHCl3 solid
393, 315, 258
660
10290
1.152
0.1
457
600
5215
0.977
1.4
392
nd
––
1.99 (82.86 %) 4.67 (17.14%) 3.08 23.92 (33.42%) 93.97 (66.58%) ––
––
––
Compound
1 O
O H3C
CH3
N N
Cl
N Re
OC
CO CO
film
Journal Pre-proof 2 O
O
H3C
CH3
S
378, 331, 297 394, 297 416
735 730 630
12850 12610 8165
4.44 6.29 362.08
0.881 1.049 1.128
0.9 0.4 3.8
MeCN CHCl3 solid film
385, 285 425, 316 503, 406 412
743 744 667 621
12515 10090 4890 ––
3.63 4.99 63.70 ––
0.989 0.916 1.080 ––
0.3 0.4 1.8 ––
MeCN CHCl3 solid
369, 285 393, 317 462, 380, 265
666 657 580
12085 10225 4400
1.098 1.082 1.013
1.1 1.3 16.3
392
nd
––
2.14 3.86 67.87 (4.28%) 466.79 (95.72%) ––
––
––
MeCN CHCl3 solid film
398, 335, 303 401, 345, 257 495 416
703 702 620 nd
10900 10690 4070 ––
4.79 6.28 208.12 ––
1.062 1.057 1.060 ––
0.4 0.3 9.5 ––
MeCN CHCl3 solid
394, 295 418, 296 480, 371
725 719 666
11590 10015 5820
0.908 1.095 1.013
0.7 0.8 1.6
616
––
3.66 5.12 13.88 (18.94%) 119.95 (81.06%) ––
––
––
1.037
0.7
1.164
0.6
1.060
0.6
–– 1.083 1.037 1.128
–– 1.0 1.3 8.2
S
N Cl
N
MeCN CHCl3 solid
N Re
OC
CO CO
3 O
O H3C
CH3
N
N
N
N Cl
N Re
OC
CO
4 H3C
O O
O H3C
CH3
N N
Cl
film
N Re
OC
CO
H3C
O
O
O
H3C
CH3
S
S
N Cl
N
N Re CO
re
OC
H3C O O
O H3C
CH3
N
N
N
N Cl
film
N Re
CH3
O
O
CH3
MeCN
387, 303, 239
656
10595
CHCl3
399, 302, 260
654
10415
477, 326
633
5165
406 397, 369, 297 419, 328, 248 481
573 722 717 612
–– 11335 9920 4450
2.88 (48.14%) 7.13 (51.86%) 4.14 (89.72%) 30.36 (10.28%) 39.66 (48.33%) 137.77(51.67%) –– 4.79 6.99 216.54
413, 334, 292, 245 431, 337, 295 539, 388 432
734
10590
3.93
1.022
0.9
739 686 nd
9670 3975 ––
5.66 158.14 ––
1.032 1.233 ––
0.6 0.8 ––
solid
N N Cl
N Re
OC
CO CO
8 O
CH3
O
Cl
film MeCN CHCl3 solid
S
N N
CH3
na
7
S
410
CO CO
Jo ur
OC
lP
CO
6
-p
5
ro
CO
of
CO
N Re
OC
CO CO
MeCN
9 O
CH3
O
N
N N Cl
N N
Re OC
CO CO
CH3
CHCl3 solid film
The underlined wavelengths were taken to register the emission spectrum; nd- emission was not detected
The emission of 1–9 is dependent on the imine acceptor core demonstrating red-shift in the order dppy dtpy terpy for all three series 1–3, 4–6 and 7–9 (Table 3).
Journal Pre-proof Exemplarily, in the solid state, the emission maxima of 1, 2 and 3 appeared at 600, 630 and 667 nm, respectively. Solid-state emission of terpyridyl Re(I) complexes is also clearly modulated by the substitution pattern of methoxy groups and π conjugation of aryl group attached to the central pyridine. The red-shift follows the order 3,4,5-trimethoxy-1-phenyl (4, em=580 nm ) < 3,5-dimethoxy-1-phenyl (1, em=600 nm) < 4,7-dimethoxy-1-naphthyl (7, em=633 nm). Conversely, in solution, the emission maxima of terpyridyl Re(I) complexes (1, 4 and 7) appear in rather narrow range 654–672 nm, demonstrating little impact of substitution pattern of methoxy groups and π conjugation of the aryl group. Stronger substituent influence was
of
observed for Re(I) bearing dtpy- and dppy-based ligands. As shown in Table 3, the
occurs
in
higher
energy
region
in
ro
emission of Re(I) with dtpy/dppy substituted with 3,4,5-trimethoxy-1-phenyl (5 and 6) relation
to
those
with
attached
-p
3,5-dimethoxy-1-phenyl (2 and 3) and 4,7-dimethoxy-1-naphthyl (8 and 9). A hypsochromic shift of the emission after introduction of methoxy group at the para
re
position of the phenyl ring may imply MLCT character of the excited states for Re(I)
lP
complexes bearing triimine ligands with phenyl-based substituents (1–6). The electron-donating groups are expected to destabilize MLCT excited states, which is manifested in blue-shift of the emission band [86]. A red-shift reported for 8 and 9 with
na
4,7-dimethoxy-1-naphthyl can be attributed to the effect of extended π conjugation of
process.
Jo ur
naphthyl substituent and contribution of ILCT excited states in the deactivation
In solution, the Re(I) carbonyls are weakly emissive, with quantum yields below 1.3%. The photoluminescence lifetimes of 1–9 fall in the nanosecond time regime, which is characteristic of vast majority of rhenium(I) carbonyls with terpy-like ligands [74-79]. Larger values of
have been reported for rhenium(I) carbonyls with
2,2′:6′,2″-terpyridine functionalized with carbazole and diphenylamine substituents [70]. In the powdered solid-state form, the quantum yield for 1–9 was found to be strongly structure-dependent, varying from 0.6% to 16.3 %. Particularly significant enhancement of the quantum yield was found for 4 (16.3 %) and 5 (9.5 %) in comparison to that of 1 (1.4 %) and 2 (3.8 %), while the solid state emission quantum yield of Re(I) bearing dppy-derivatives was only slightly affected by ligand modification (0.8-1.8 %). The data clearly indicate beneficial influence of the electron-donating methoxy group at the para position of the phenyl ring attached to the
Journal Pre-proof central pyridine ring of terpy and dtpy. Referring to the previously examined [ReCl(CO)3(4-R-terpy-κ2N)] (30.55 %) and [ReCl(CO)3(4-R-dtpy-κ2N)] (5.87 %) with 4-methoxy-1-phenyl substituent, however, different trends are observed for 4 and 5 complexes. While the emission quantum yield of terpyridyl Re(I) complex (4) decreases, the compound 5 shows an enhancement of photoluminescence, demonstrating that the substituent effect depends on the triimine core. A remarkably improved emission efficiency in solid state was also found for 8. Its emission quantum yield (8.2 %) was significantly higher in relation to [ReCl(CO)3(4-R-dtpy-κ2N)] with 4-methoxy-1-naphthyl (3.43 %) [78].
of
Additionally, the possibility of light emission induced by external voltage of
thin
films
were
applied
as
active
ro
prepared Re(I) complexes was tested. Compounds 3, 6 and 7, which exhibited PL as layers
in
a
diode
structure
-p
ITO/PEDOT:PSS/complex/Al. However, only device with 7 emitted light with maximum emission band (EL) located at 666 nm (cf. Fig. 5). In the next step, diodes
re
with guest-host configuration were fabricated. As matrix, a mixture of (PVK)
(50
wt
%)
and
lP
poly(9-vinylcarbazole)
(2-tert-butylphenyl-5-biphenyl-1,3,4-oxadiazole) (PBD) (50 wt %) was utilized. Re(I) complexes 1, 3, 4–7 and 9 as luminophore with concentration of 2 wt% in the matrix
na
were dispersed. All diodes were emissive and emitted light with EL in the range of 594–630 nm. It was found that the devices with complexes bearing terpyridine skeleton
Jo ur
(1, 4 and 7) emitted light with EL bathochromically shifted compared to the others (cf. Table 4). The effect of substituent on the EL ability can be seen in the series of diodes with 1, 4 and 7. The highest EL intensity was found for the device with compound containing 3,5- dimetoxy-1-phenyl unit (1). However, the lowest turn-on voltage ws measured
for
the
diode
based
on
triimine
skeleton
substituted
with
3,4,5-trimethoxy-1-phenyl unit (4). Additionally, diodes with lower luminophore (6 and 7) content, that is, 1 wt % in the matrix, were prepared. It can be seen that together with reduction of compound content, significant increase of the EL intensity was observed (cf. Table 4)
Journal Pre-proof
200
16V
17V
18V
14V
150 640 nm
100
50
0 400
500
600
ITO/PEDOT:PSS/PVK:PBD:7 (2wt%)/Al
19V
Intensity [counts]
Intensity [counts]
900
ITO/PEDOT:PSS/7/Al
700
800
600
500
6000
700
800
900
ITO/PEDOT:PSS/PVK:PBD:7(1wt%)/Al
of
Intensity [counts]
5000
23V
4000
ro
590 nm
2000
500
600
700
800
Wavelength [nm]
900
3000 2000
-p
4000
re
Intensity [counts]
23V
0 400
600
Wavelength [nm]
ITO/PEDOT:PSS/PVK:PBD:7(1wt%)/Al
6000
18V
300
Wavelength [nm]
8000
17V
594 nm
0 400
900
15V
1000 0
0
10
20
30
40
Voltage [V]
lP
Figure 5. Electroluminescence spectra of the diodes based on Re(I) complex 7.
na
Table 4. Position of maximum of electroluminescence band (EL) obtained under external voltage (V) and its maximal reached EL intensity.
Jo ur
Code ITO/PEDOT:PSS/complex/Al
λEL [nm] (V)
Intensity [a.u]
–– –– 1 nd nd 3 –– –– 4 –– –– 5 nd nd 6 640 (19V) 150 7 –– –– 9 ––: not measured; nd: EL not detected
ITO/PEDOT:PSS/PVK:PBD:complex/Al content of complex in the matrix 2 wt% 1 wt% λEL [nm] (V) Intensity λEL [nm] Intensity [a.u] (V) [a.u] –– –– 605 (26V) 9055 –– –– 625 (10V) 1068 –– –– 594 (6V) 614 –– –– 630 (16V) 1505 625 (25V) 1913 615 (20V) 5233 594 (15V) 504 591 (23V) 5032 –– –– 630 (26V) 111
Electrochemistry To investigate the redox behavior and estimate HOMO/LUMO energy levels and band gap energy (Eg) of 1–9, cyclic voltammetry (CV) and differential pulse voltammetry
Journal Pre-proof (DPV) on a glassy carbon working electrode in CH2Cl2 with 0.1 M n-Bu4NPF6 as the supporting electrolyte were used. The potentials were referenced to the ferrocene/ferrocenium redox couple, and onsets of the first oxidation and reduction waves were used to estimate the values of IP and EA, which can be regarded as closely related to the HOMO and LUMO levels, respectively [87]. The electrochemical data are gathered in Table 5, while the CVs and DPVs are shown in Figure S30.
Table 5. Relevant electrochemical data.
-1.70 -1.49 -1.37 -1.88 -1.60 -1.39 -1.71 -1.60 -1.34
0.71 0.90 1.06 0.58 0.82 1.02 0.82 0.87 1.08
0.62 0.78 0.86 0.50 0.73 0.79 0.69 0.73 0.89
IP [eV]
Eg [eV]
of
-1.80 -1.60 -1.50 -1.99 -1.69 -1.48 -1.82 -1.70 -1.46
EA [eV]
-5.72 -5.88 -5.96 -5.60 -5.83 -5.89 -5.79 -5.83 -5.99
2.32 2.27 2.23 2.38 2.33 2.18 2.40 2.33 2.35
-3.40 -3.61 -3.73 -3.22 -3.50 -3.71 -3.39 -3.50 -3.64
ro
1 2 3 4 5 6 7 8 9
Ered1(onset) Eox1(onset) E ox1 [V] [V] [V]
-p
Ered1 [V]
re
Code
lP
IP = -5,1 - Eox1(onset), EA = -5,1 - Ered1(onset), Eg = Eutl(onset) - Ered(onset). Solution: CH2Cl2, concentration: 10-3 mol/L, electrolyte: 0.1M Bu4NPF6. GC as working electrode.
na
Consistent with the theoretical studies, the LUMO of these systems is predominately localized on the triimine core, thus the first wave, reversible (2) or quasi-reversible (1,
Jo ur
3-9), can be assigned to triimine-centered reduction process. Introduction of an additional N- or S-donor atom into the -deficient trisheterocyclic unit of dtpy and dppy leads to increase of -acceptor properties of dtpy and dppy ligands and stabilization of the LUMO orbital Re(I) complexes bearing 2,6-di(thiazol-2-yl)pyridine and 2,6-di(pyrazin-2-yl)pyridine derivatives, relative to those with 2,2′:6′,2′′-terpyridines, which is reflected in occurring the first reduction peak at less negative potentials in the order terpy dtpy dppy. The first oxidation potentials of 1–9, assigned to the ReI/ReII couple, were slightly sensitive to the type of the triimine skeleton. For all three series 1–3, 4–6 and 7–9, the easiest to oxidize were the Re(I) complexes with the 2,2′:6′,2′′-terpyridine ligands, while those bearing with 2,6-di(pyrazin-2-yl)pyridines showed the highest values of Eox1. A small variation of the oxidation potential can be also correlated with the substitution pattern of phenyl ring. Introduction of the methoxy group into the para
Journal Pre-proof position resulted in slightly decreased first oxidation potential of 4–6 relative to corresponding 1–3.
Conclusions Concluding the results obtained in the context of impact of both the chemical structure of the triimine skeleton and the substituent:
all synthesized complexes showed high enough values of Tm, which is necessary for applications in optoelectronic devices;
Re(I) complexes of 2,2′:6′,2′′-terpyridines (terpy) functionalized with
of
2,6-di(pyrazin-2-yl)pyridine
substituted
with
ro
carbonyl
of
3,5-dimethoxy-1-phenyl (1) and 4,7-dimethoxy-1-naphthyl (7) units and Re(I)
3,5-dimethoxy-1-phenyl group (3) formed amorphous material with high Tg; replacement of terpy core by dtpy and dppy resulted in bathochromic shift of the
-p
series 1–3, 4–6 and 7–9;
methoxy substituent pattern had weaker impact on the location of the
lP
re
low-energy absorption bands and emission maxima in solution for all three
longest-wavelength absorption and emission in solution than π conjugation of the attached aryl group;
in film, upon excitation into the lowest-energy absorption band, only the Re(I) complexes
of
na
2,6-di(pyrazin-2-yl)pyridines
(dppy)
substituted
with
Jo ur
3,5-dimethoxy-1-phenyl (3) and 3,4,5-trimethoxy-1-phenyl (6) and the Re(I) carbonyl of 2,2′:6′,2′′-terpyridine with 4,7-dimethoxy-1-naphthyl (7) were emissive and emitted light with the em in the range of 573–621 nm. Among them only 7 was applied as an active layer in a diode, and exhibited weak electroluminescence. On the other hand, all the devices with Re(I) complexes dispersed molecularly in a matrix emitted light under external voltage.
Experimental section Materials Re(CO)5Cl and poly(9-vinylcarbazole) PVK (Mn= 25000–50000) were commercially available (Sigma Aldrich) and they were used without further purification. Poly(3,4-(ethylenedioxy)thiophene): poly-(styrenesulfonate) (PEDOT:PSS) (0.1–1.0 S/cm) and substrates with pixilated ITO anodes were supplied by Ossila. All solvents
Journal Pre-proof for synthesis were of reagent grade, while HPLC grade solvents were used for spectroscopy studies. The ligands L1–L9 were synthesized following the methodology based on condensation of 2-acetylpyridine (L1, L4, L7), 2-acetylthiazole (L2, L5, L8) and 2-acetylpyrazine (L3, L6, L9) with 3,5-dimethoxybenzaldehyde (L1-L3), 3,4,5-trimethoxy-1-naphthaldehyde (L4-L6) and 4,8-dimethoxy-2-naphthaldehyde (L7-L9) [75, 78, 79]. Preparations of [ReCl(CO)3(Ln-κ2N)] complexes (1–9) The precursor [Re(CO)5Cl (0.10 g, 0.27 mmol) and molar equivalent of the corresponding Ln ligand (0.27 mmol) were dissolved in argon-saturated acetonitrile (1,
of
2, 4, 5, 6, 7, 9) or toluene (3, 8) (50 mL) and heated under reflux for 8h. The resulting
ro
yellow (1, 4, 5, 7), red (3) or orange (2, 6, 9, 8) solid was collected by filtration, washed with diethyl ether and dried.
-p
[ReCl(CO)3(L1)] (1): Yield: 65 %. IR (KBr, cm-1): 2021(vs), 1914(vs) and 1890(vs) ν(C≡O); 1606 (m), 1541 (m) ν(C=N) and ν(C=C). 1H NMR (400 MHz, DMSO) δ 9.11
re
(d, J = 8.2 Hz, 1H, HC4), 9.07 (d, J = 5.2 Hz, 1H, HC1), 9.04 (s, 1H, HB4), 8.81 (d, J = 4.3 Hz, 1H, HA1), 8.39 (t, J = 7.8 Hz, 1H, HC3), 8.22 (s, 1H, HB2), 8.07 (t, J = 7.1 Hz, 1H,
lP
HA3), 7.91 (d, J = 7.7 Hz, 1H, HA4), 7.81 – 7.76 (m, 1H, HC2), 7.66 – 7.61 (m, 1H, HA2), 7.31 (d, J = 2.1 Hz, 2H, HD2), 6.74 (t, J = 2.0 Hz, 1H, HD4), 3.89 (s, 6H, HD5).13C NMR
na
(100 MHz, DMSO) δ 197.81 (CCO), 194.41 (CCO), 190.94 (CCO), 161.40 (CB1), 161.23 (CD3), 157.84 (CA5), 157.03 (CB5), 156.22 (CC5), 152.69 (CC1), 150.74 (CD1), 149.21
Jo ur
(CA1), 139.93 (CC3), 136.98 (CB3), 136.93 (CA3), 127.49 (CC2), 125.60 (CA4), 125.20 (CC4), 124.94 (CA2), 124.73 (CB2), 120.98 (CB4), 105.94 (CD2), 102.79 (CD4), 55.68 (CD5). HR-MS (ESI): calcd. for C26H19N3O5Re [M-Cl]+ 640.0884, found 640.0878. Anal. Calcd. for C26H19O5N3ClRe (675.11 g/mol): C, 46.26; H, 2.84; N, 6.22 %; found: C, 46.38; H, 2.98; N, 6.31 %. DSC: I heating scan: Tm=297 C, II heating scan Tg=173 C, Tc=266 C, Tm= 290 C. [ReCl(CO)3(L2)] (2): Yield: 60%. IR (KBr, cm-1): 2019 (vs), 1907(vs) ν(C≡O); 1606 (m), 1537 (m) ν(C=N) and ν(C=C). 1H NMR (400 MHz, Acetone) δ 8.79 (s, 1H), 8.36 (d, J = 2.8 Hz, 1H), 8.29 – 8.23 (m, 2H), 8.13 (d, J = 2.6 Hz, 1H), 8.07 (d, J = 2.5 Hz, 1H), 7.25 (s, 2H), 6.72 (s, 1H), 3.90 (s, 6H). 13C NMR not recorded due to insufficient complex solubility. HR-MS (ESI): calcd. for C22H15ClN3NaO5ReS2 [M+Na]+ 709.9580, found 709.9583. Anal. Calcd. for C22H15O5N3ClReS2 (687.15 g/mol): C, 38.4; H, 2.20; N 6.12%; found C, 38.52; H, 2.55; N, 6.06 %.
Journal Pre-proof [ReCl(CO)3(L3)] ∙ C6H5CH3, (3∙ C6H5CH3): Yield: 70%. IR (KBr, cm-1): 2029(vs), 1933(vs) and 1909(vs) ν(C≡O); 1603 (m), 1594 (m) ν(C=N) and ν(C=C). 1H NMR (400 MHz, DMSO) δ 10.37 (d, J = 0.9 Hz, 1H), 9.28 (d, J = 1.8 Hz, 1H), 9.15 (d, J = 1.2 Hz, 1H), 9.14 (dd, J = 3.1, 1.1 Hz, 1H), 8.99 (d, J = 3.1 Hz, 1H), 8.94 – 8.92 (m, 2H), 8.46 (d, J = 1.8 Hz, 1H), 7.37 (d, J = 2.2 Hz, 2H), 6.75 (t, J = 2.1 Hz, 1H), 3.89 (s, 6H). 13
C NMR (100 MHz, DMSO) δ 196.55, 194.55, 189.49, 161.27, 158.70, 155.41,
153.51, 150.95, 150.91, 147.93, 147.20, 145.85, 145.79, 145.33, 144.07, 136.43, 125.80, 121.95, 106.05, 103.10, 55.70. HR-MS (ESI): calcd. for C24H17ClN5NaO5Re [M+Na]+ 700.0360, found 700.0363. Anal. Calcd. for C25H18O5N5ClRe (690.11
ro
heating scan: Tm=258 C, II heating scan Tg=164 C.
of
g/mol): C, 43.51; H, 2.63; N, 10.15 %; found: C, 43.43; H, 2.81; N, 10.18 %. DSC: I [ReCl(CO)3(L4)] (4): Yield: 70%.IR (KBr, cm-1): 2023(vs), 1944(vs) and 1864 (vs)
-p
ν(C≡O); 1611 (m), 1587 (m) ν(C=N) and ν(C=C). 1H NMR (400 MHz, DMSO) δ 9.09 – 9.03 (m, 2H, HC1+C4), 9.00 (s, 1H, HB4), 8.80 (d, J = 4.2 Hz, 1H, HA1), 8.40 (t, J = 7.6
re
Hz, 1H, HC3), 8.23 (s, 1H, HB2), 8.06 (t, J = 6.3 Hz, 1H, HA3), 7.88 (d, J = 7.4 Hz, 1H, HA4), 7.79 (t, J = 6.4 Hz, 1H, HC2), 7.66 – 7.61 (m, 1H, HA2), 7.40 (s, 2H, HD2), 3.95 (s, 13
C NMR (100 MHz, DMSO) δ 197.82 (CCO), 194.41
lP
6H, HD5), 3.76 (s, 3H, HD6).
(CCO), 190.98 (CCO), 161.31 (CB1), 157.97 (CA5), 156.84 (CB5), 156.27 (CC5), 153.58
na
(CD3), 152.71 (CC1), 150.89 (CD1), 149.23 (CA1), 139.90 (CC3), 136.94 (CA3), 130.47 (CB3), 127.47 (CC2), 125.51 (CC4), 125.19 (CA4), 124.92 (CA2), 124.51 (CB2), 120.89
Jo ur
(CB4), 105.70 (CD2), 60.21 (CD6), 56.44 (CD5). HRMS (ESI): calcd. for C27H22N3O6Re [M+H]+ 706.0741; found: 706.0742. Anal. Calcd. for C27H21O6N3ClRe (705.14 g/mol): C, 45.99; H, 3.00; N, 5.96 %; found: C, 46.07; H, 3.16; N, 6.07 %. DSC: I heating scan: Tm=322 C with degradation.
[ReCl(CO)3(L5)] (5): Yield: 60%. IR (KBr, cm-1): 2019(vs), 1920(vs) and 1910(vs) ν(C≡O); 1612 (m), 1589 (m) ν(C=N) and ν(C=C). 1H NMR (400 MHz, DMSO) δ 9.23 (s, 1H), 8.80 (d, J=3.1 Hz 1H), 8.72 (s, 1H), 8.69 (d, J = 3.0 Hz, 1H), 8.58 (d, J = 2.7 Hz, 1H), 8.51 (d, J = 3.0 Hz, 1H), 7.90 (s, 2H), 4.41 (s, 6H), 4.28 (s, 3H). 13C NMR not recorded due to insufficient complex solubility. HR-MS (ESI): calcd. for C23H17ClN3NaO6ReS2 [M+Na]+ 739.9686, found 739.9699. Anal. Calcd. for C23H17O6N3ClReS2 (717.18 g/mol): C, 38.52; H, 2.39; N, 5.86 %; found: C, 38.39; H, 2.72; N, 5.88. DSC: I heating scan: Tm=325 C with degradation.
Journal Pre-proof [ReCl(CO)3(L6)].H2O (6.H2O): Yield: 70%. IR (KBr, cm-1): 2026(vs), 1956(vs) and 1898(vs) ν(C≡O); 1608 (m), 1585 (m) ν(C=N) and ν(C=C). 1H NMR (400 MHz, DMSO) δ 10.34 (s, 1H), 9.23 (d, J = 0.7 Hz, 1H), 9.15 – 9.13 (m, 2H), 8.99 (d, J = 3.0 Hz, 1H), 8.94 – 8.92 (m, 2H), 8.45 (d, J = 1.7 Hz, 1H), 7.46 (s, 2H), 3.95 (s, 6H), 3.77 (s, 3H).
13
C NMR (100 MHz, DMSO) δ 196.58, 194.58, 189.56, 158.61, 155.26,
153.63, 151.10, 150.98, 147.93, 147.13, 145.91, 145.82, 145.34, 144.14, 140.21, 129.89, 125.54, 121.77, 105.82, 60.25, 56.49. HR-MS (ESI): calcd. for C25H19ClN5NaO6Re [M+Na]+ 730.0465, found 730.0466. Anal. Calcd. for C25H21O7N5ClRe (725.13 g/mol): C, 42.46; H, 2.71; N, 9.90 %; found: C, 41.41; H,
of
2.92; N, 9.66 %. DSC: I heating scan: Tm=281 C with degradation. [ReCl(CO)3(L7)] (7.CH3CN.H2O): Yield: 70%. IR (KBr, cm-1): 2020(vs), 1919(vs)
ro
and 1880(vs) ν(C≡O); 1622(m), 1606 (m), 1582 (m) ν(C=N) and ν(C=C). 1H NMR
-p
(400 MHz, DMSO) δ 9.08 (d, J = 5.2 Hz, 1H, HC1), 9.00 (s, 1H, HB4), 8.97 (d, J = 8.2 Hz, 1H, HC4), 8.78 (d, J = 4.6 Hz, 1H, HA1), 8.35 – 8.30 (m, 1H, HC3), 8.23 (d, J = 9.1
re
Hz, 1H, HD6), 8.05 (dt, J = 7.7, 0.7 Hz, 1H, HA3), 7.98 (s, 1H, HB2), 7.93 (d, J = 7.8 Hz, 1H, HA4), 7.79 – 7.73 (m, 2H, HC2+D2), 7.63 – 7.59 (m, 1H, HA2), 7.31 – 7.26 (m, 2H,
lP
HD7+D9), 7.04 (d, J = 8.2 Hz, 1H, HD3), 4.04 (s, 3H, HD11), 3.76 (s, 3H, HD12). 13C NMR (100 MHz, DMSO) δ 197.73 (CCO), 194.51 (CCO), 191.06 (CCO), 161.12 (CB1), 158.61
na
(CD8), 157.72 (CA5), 156.74 (CB5), 156.59 (CD4), 156.39 (CC5), 152.72 (CD1), 151.94 (CC1), 149.28 (CA1), 140.02 (CC3), 136.87 (CA3), 132.30 (CD10), 129.93 (CD2), 127.84
Jo ur
(CB2), 127.37 (CC2), 125.75 (CC4), 125.31 (CA4), 125.13 (CA2), 124.91 (CB3), 124.27 (CD6), 124.11 (CB4), 120.07 (CD5), 117.42 (CD7), 103.66 (CD9), 102.57 (CD3), 55.92 (CD11), 55.08 (CD12). HR-MS (ESI): calcd. for C30H22ClN3O5Re [M+H]+ 726.0792, found 726.0799. Anal. Calcd. for C32H26O6N4ClRe (784.23 g/mol): C, 49.01; H, 3.34; N, 7.14 %; found: C, 49.34; H, 3.48; N, 7.12 %. DSC: I heating scan: Tm= 198 C, II heating scan Tg=158 C. [ReCl(CO)3(L8)]∙½ C6H5CH3 (8∙½C6H5CH3): Yield: 70%. IR (KBr, cm-1): 2023(vs), 1924(vs) and 1898(vs) ν(C≡O); 1618 (m), 1609 (m), 1584 (m) ν(C=N) and ν(C=C) 1H NMR (400 MHz, Acetone) δ 8.73 (s, 1H), 8.38 (d, J = 3.3 Hz, 1H), 8.29 (d, J = 9.2 Hz, 1H), 8.25 (d, J = 3.3 Hz, 1H), 8.15 – 8.11 (m, 2H), 8.06 (d, J = 3.2 Hz, 1H), 7.77 (d, J = 8.1 Hz, 1H), 7.44 (d, J = 2.4 Hz, 1H), 7.26 – 7.25 (m, 1H), 7.04 (d, J = 8.1 Hz, 1H), 4.10 (s, 3H), 3.83 (s, 3H).
13
C NMR not recorded due to insufficient complex solubility.
HR-MS (ESI): calcd. for C26H17ClN3NaO5ReS2 [M+Na]+ 759.9737, found 759.9741.
Journal Pre-proof Anal. Calcd. for C26H19O5N3ClRe (675.11 g/mol): C, 46.26; H, 2.84; N, 6.22 %; found: C, 46.38; H, 2.98; N, 6.31 %. [ReCl(CO)3(L9)].CHCl3 (9): Yield: 65%. IR (KBr, cm-1): 2023(vs), 1920(vs) and 1893(vs) ν(C≡O); 1622 (m), 1608 (m), 1574 (m) ν(C=N) and ν(C=C). 1H NMR (400 MHz, DMSO) δ 10.24 (s, 1H), 9.28 (d, J = 1.2 Hz, 1H), 9.16 (s, 1H), 9.15 (d, J = 2.9 Hz, 1H), 8.97 (d, J = 3.0 Hz, 1H), 8.90 (s, 2H), 8.26-8.21 (m, 2H), 7.80 (d, J = 8.1 Hz, 1H), 7.36 (d, J = 2.3 Hz, 1H), 7.29 (dd, J = 9.2, 2.4 Hz, 1H), 7.07 (d, J = 8.2 Hz, 1H), 4.05 (s, 3H), 3.78 (s, 3H).
13
C NMR (100 MHz, DMSO) δ 196.50, 194.59, 189.63,158.76,
158.48, 156.85, 155.15, 153.45, 152.36, 151.16, 147.83, 146.99, 145.88, 145.78,
of
145.36, 144.13, 132.27, 130.44, 128.96, 125.37,125.28, 124.13, 120.06, 117.56,
ro
103.72, 102.66, 55.99, 55.16. HR-MS (ESI): calcd. for C28H19ClN5NaO5Re [M+Na]+ 750.0516, found 750.0508. Anal. Calcd. for C29H20O5N5Cl4Re (846.52 g/mol) calcd.
-p
41.15; H, 2.38; N, 8.27 %; found: C, 41.87; H, 2.62; N, 8.50 %. DSC: I heating scan:
re
Tm=233 C, Tc=244 C, Tm= 287 C with degradation. Crystal structure determination and refinement
lP
X-ray quality crystals of 5, 7 and 9 were obtained by recrystallization from mixture of acetonitrile/chloroform. The X-ray intensity data of 5, 7 and 9 were collected on a
na
Gemini A Ultra diffractometer equipped with Atlas CCD detector and graphite monochromated MoKα radiation (λ= 0.71073 Å) at room temperature. The unit cell
Jo ur
determination and data integration were carried out using the CrysAlis package of Oxford Diffraction, and Lorentz polarization and empirical absorption correction were applied [88]. The structures were solved by the Patterson method using SHELXS-2014 and refined with SHELXL-2014 using full-matrix least-squares on F2 [89]. All the non-hydrogen atoms were refined anisotropically, and hydrogen atoms were placed in calculated positions refined using idealized geometries (riding model) and assigned fixed isotropic displacement parameters, d(C–H) = 0.93 Å, Uiso(H) = 1.2 Ueq(C) (for aromatic) and d(C–H) = 0.96 Å, Uiso(H) = 1.5 Ueq(C) (for methyl). The methyl groups were allowed to rotate about their local threefold axis. The crystallographic data of 5, 7 and 9 are summarized in Table S1 in ESI. CCDC- 1974623 (5), 1974625 (7) and 1974624 (9) contains the supplementary crystallographic data. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif.
Journal Pre-proof Physical measurements The IR spectra were recorded with a Nicolet iS5 FTIR spectrophotometer in the spectral range 4000–400 cm–1 with the samples in the form of KBr pellets. The electronic spectra were measured using ThermoScientific Evolution 220 UV/Vis Spectrometer (in MeCN and CHCl3 solution) and Jasco V570
UV-V-NIR
Spectrometer (in solid state as film deposited on a glass substrate). The 1H NMR and 13
C NMR spectra were recorded (295 K) on Bruker Avance 400 NMR spectrometer at a
resonance frequency of 400 MHz for 1H NMR spectra and 100 MHz for
13
C NMR
spectra using DMSO-d6 or acetone-d6 as solvent. The HRMS measurements were
of
performed using Q-TOF MaXis Impact Bruker mass spectrometer equipped with an
ro
electrospray (ESI) ion source and q-TOF type mass analyzer. The recorded data were processed using Data Analylis 4.1 software package. The analyzed samples were
-p
dissolved in a mixture of MeCN CHCl3 (1/1). Elemental analysis was registered by Vario EL Cube (Elemental company) using acetanilide as a standard. Steady-state
re
luminescence spectra of solid state and solution samples were measured with FLS-980 fluorescence spectrophotometer equipped with a 450 W Xe lamp and high gain
lP
photomultiplier PMT+250nm (Hamamatsu, R928P) detector. The PL lifetime measurement was performed with a time correlated single photon counting (TCSPC) or
na
multi-channel scaling (MCS) method. Excitation wavelength (375 nm, 410 nm) for TCSPC was obtained using the TCSPC diode with various pulse periods as light
Jo ur
source. For MCS excitation wavelength was obtained using 60W microsecond Xe flash lamp. Differential Scanning Calorimetry (DSC) was performed with a TA-DSC 2010 apparatus, under nitrogen atmosphere using sealed aluminum pans with heating rate 20 °C/min. Electrochemical measurements were performed with Eco Chemie Autolab PGSTAT128n potentiostat device. The measurements were investigated in dichloromethane (Sigma-Aldrich, 99.8%) solution (c = 10-3 mol/L) using glassy carbon electrode (diam. 2.0 mm) as a working electrode with 0.1M Bu4NPF6 (Sigma-Aldrich, 99%) as a electrolyte. The platinum coil and silver wire as auxiliary and reference electrode were used. Cyclic voltammetry (CV) were recorded with moderate scan rate equal to 0.1 V/s and differential pulse voltammetry (DPV) with moderate scan rate equal to 0.01 V/s. The solution was purged with argon for about 10 min before every measurement. All potentials were referenced with respect to ferrocene couple (Fc/Fc+) which was used as the internal standard. The IP of Fc/Fc+
Journal Pre-proof was calculated to be equal to -5.1 eV as shown in literature data [86]. All measurements were performed at 25 ± 1 ˚C. Photoluminescence spectra in solid state as films deposited on a glass substrate and as blends with PVK:PBD were registered using Hitachi F-2500 spectrometer. In order to collect electroluminescence (EL) spectra the voltage was applied using a precise voltage supply (GwInstek PSP-405) and the sample was fixed to an XYZ stage. Light from the OLED device was collected through a 30 mm lens, focused on the entrance slit (50 μm) of a monochromator (Shamrock SR-303i) and detected using a CCD detector (Andor iDus 12305). Typical acquisition times were equal to 10 seconds. The
of
pre-alignment of the setup was done using a 405 nm laser.
ro
Computational details
-p
The calculations were performed using the GAUSSIAN-16 program package [90]. The geometries of the singlet ground state (S0) of 1–9 were fully optimized without any
re
symmetry restrictions at the DFT level with the PBE1PBE hybrid exchange-correlation functional [91-92]. The calculations were performed using def2-TZVPD basis set for
lP
rhenium and 6-31g* for carbon, 6-31g for hydrogen and 6-31g** basis set for other elements (Cl, N, O) [93-97]. The starting point for geometry optimization was taken
na
from the X-ray structure, and all the subsequent calculations were performed based on the optimized geometries. To verify that each of the geometries is a minimum on the
Jo ur
potential energy surface, vibrational frequencies were calculated on the basis of the optimized geometry. Polarized continuum model (PCM) was employed to recreate media effects in acetonitrile and chloroform (MeCN and CHCl3) [98-100]. The predicted bond lengths and angles for the ground state are within the range of error expected for DFT calculations of rhenium(I) complexes, and the general trends observed in the experimental data are well reproduced in the calculations, providing confidence on the reliability of the chosen method to reproduce the geometry of studied complexes (Table S2).
Acknowledgments The research was co-financed by National Science Centre of Poland grants no. DEC2017/25/B/ST5/01611 and 2017/27/B/ST3/02457. Magdalena Małecka thanks to „PIK”- program for facilitating scientific self-development for PhD students,
Journal Pre-proof co-financed by the European Union. The calculations were carried out in Wroclaw Centre for Networking and Supercomputing (http://www.wcss.wroc.pl). Authors thank dr. Henryk Janeczek for DSC measurements.
References [1] M.S. Wrighton, D.L. Morse, The Nature of the Lowest Excited State in Tricarbonylchloro-1,10-phenanthrolinerhenium(I) and Related Complexes, J. Am. Chem. Soc. 96 (1974) 998-1003. [2] J. K.Hiño, L. D. Ciana, W. J. Dressick, P.Sullivan, Substituent Constant
of
Correlations as Predictors of Spectroscopic, Electrochemical, and Photophysical
ro
Properties in Ring-Substituted 2,2'-Bipyridine Complexes of Rhenium(I), Inorg. Chem. 31 (1992) 1072-1080.
-p
[3] L. Sacksteder, M. Lee, J.N. Demas, B.A. DeGraf, Long-Lived, Highly Luminescent Rhenium(I) Complexes as Molecular Probes: Intra- and Intermolecular Excited-State
re
Interactions, J. Am. Chem. Soc. 115 (1993) 8230-8238.
lP
[4] B. D. Rossenaar, D. J. Stufkens, A. Vlcek Jr., Halide-Dependent Change of the Lowest-Excited-State Character from MLCT to XLCT for the Complexes
na
Re(X)(CO)3(-diimine) (X = Cl, Br, I; -diimine = bpy, iPr-PyCa, iPr-DAB) Studied by Resonance Raman, Time-Resolved Absorption, and Emission Spectroscopy, Inorg. Chem. 35 (1996) 2902-2909.
Jo ur
[5] D. J. Stufkens, A. Vlcek Jr., Ligand-dependent excited state behaviour of Re(I) and Ru(II) carbonyl–diimine complexes, Coord.Chem. Rev. 177 (1998) 127-179. [6] S. Ranjan, S.–Y. Lin, K.–C. Hwang, Y. Chi, W.–L. Ching, C.–S. Liu, Realizing Green Phosphorescent Light-Emitting Materials from Rhenium(I) Pyrazolato Diimine Complexes, Inorg. Chem. 42 (2003) 1248-1255. https://doi.org/10.1021/ic025918. [7] M. Busby, A. Gabrielsson, P. Matousek, M. Towrie, A. J. Di Bilio, H. B. Gray, A. Vlcek
Jr.,
Excited-State
fac-[ReI(L)(CO)3(5-NO2-phen)]+ 1,10-Phenanthroline)
Dynamics (L
Complexes,
=
of
fac-[ReI(L)(CO)3(phen)]+
Imidazole,
Inorg.
Chem.
4-Ethylpyridine; 43
(2004)
Phen
and =
4994−5002.
https://doi.org/10.1021/ic035471b. [8] K. K.–W. Lo, K. H.–K. Tsang, W.–K. Hui, N. Zhu, Synthesis, Characterization, Crystal Structure, and Electrochemical, Photophysical, and Protein-Binding Properties
Journal Pre-proof of Luminescent Rhenium(I) Diimine Indole Complexes, Inorg. Chem. 44 (2005) 6100−6110. https://doi.org/10.1021/ic050531u. [9] A. A. Martı, G. Mezei, L . Maldonado, G. Paralitici, R. G. Raptis, J. L. Colon, Structural
and
Photophysical
Characterisation
of
fac-[Tricarbonyl(chloro)-
(5,6-epoxy-1,10-phenanthroline)rhenium(I)], Eur. J. Inorg. Chem. (2005) 118-124. https://doi.org/10.1002/ejic.200400531. [10] C.–C. Ko, W.–M. Kwok, V. W.–W. Yam, D. L. Phillips, Triplet MLCT Photosensitization of the Ring-Closing Reaction of Diarylethenes by Design and Synthesis of a Photochromic Rhenium(i) Complex of a Diarylethene-Containing Ligand,
Chem.
Eur.
J.
(2006)
ro
https://doi.org/10.1002/chem.200501325.
12
of
1,10-Phenanthroline
5840
–
5848.
[11] M.–J. Li, C.–C. Ko, G.–P. Duan, N. Zhu, V. W.–W. Yam, Functionalized Complexes
with
Crown
Ether
Pendants
Derived
from
-p
Rhenium(I)
1,10-Phenanthroline: Selective Sensing for Metal Ions, Organometallics 26 (2007)
re
6091-6098. https://doi.org/10.1021/om700481f.
[12] S.–T. Lam, N. Zhu, V. W.–W. Yam, Synthesis and Characterization of
lP
Luminescent Rhenium(I) Tricarbonyl Diimine Complexes with a Triarylboron Moiety and the Study of Their Fluoride Ion-Binding Properties, Inorg. Chem. 48 (2009)
na
9664-9670. https://doi.org/10.1021/ic900803a. [13] R. A. Kirgan, B.P. Sullivan, D. P. Rillem, Photochemistry and Photophysics of
Jo ur
Coordination Compounds: Rhenium, Top. Curr. Chem. 281 (2007) 45–100. https://doi.org/10.1007/128_2007_143. [14] A. Kumar, S.-S. Sun, A. J. Lees, Photophysics and Photochemistry of Organometallic Rhenium Diimine Complexes, Top Organomet Chem. 29 (2010) 1–35. https://doi.org/10.1007/3418_2009_2. [15] A. Vlček Jr., Ultrafast Excited-State Processes in Re(I) Carbonyl-Diimine Complexes: From Excitation to Photochemistry, Top Organomet Chem. 29 (2010) 73– 114. https://doi.org/10.1007/3418_2009_4. [16] H. Takeda, K. Koike, T. Morimoto, H. Inumaru, O. Ishitani, Photochemistry and Photocatalysis of Rhenium (I) Diimine Complexes, Adv. Inorg. Chem. 63 (2011) 137– 186. https://doi.org/10.1016/B978-0-12-385904-4.00007-X. [17] C.-C. Ko, A. W.-Y. Cheung, L.T.-L. Lo, J. W.-K. Siu, C.-O. Ng, S.-M. Yiu, Syntheses and photophysical studies of new classes of luminescent isocyano
Journal Pre-proof rhenium(I) diimine complexes, Coord. Chem. Rev. 256 (2012) 1546–1555. https://doi.org/10.1016/j.ccr.2012.01.006. [18] C.–C. Ko, L. T.-L. Lo, C.-O. Ng, S.-M. Yiu, Photochemical Synthesis of Intensely Luminescent Isocyano Rhenium(I) Complexes with Readily Tunable Structural Features,
Chem.
Eur.
J.
16
(2010)
–
13773
13782.
https://doi.org/10.1002/chem.201000793. [19] A. Otavio, T. Patrocinio, M. K. Brennaman, T. J. Meyer, N. Y. Murakami Iha, Excited-State Dynamics in fac-[Re(CO)3(Me4phen)(L)]+, J. Phys. Chem. A 114 (2010) 12129–12137. https://doi.org/10.1021/jp104692w.
of
[20] C.–C. Ko, C.–O. Ng, H. Feng, W.–K. Chu, Synthesis, characterisation and
tricarbonyl
diimine
complexes,
Dalton
Trans.
39
(2010)
6475–648.
-p
https://doi.org/10.1039/b927167h.
ro
photophysical studies of leucotriarylmethanes-containing ligands and their rhenium(I)
[21] C.–O. Ng, S.–W. Lai, H. Feng, S.–M. Yiu, C.–C. Ko, Luminescent rhenium(I)
ligands:
Anion-sensing
study,
Dalton
Trans.
40
(2011)
10020–10028.
lP
https://doi.org/10.1039/c1dt10831j.
re
complexes with acetylamino- and trifluoroacetylamino-containing phenanthroline
[22] J. E. Jones, B. M. Kariuki, B. D. Ward, S. J. A. Pope, Amino-anthraquinone
na
chromophores functionalised with 3-picolyl units: structures, luminescence, DFT and their coordination chemistry with cationic Re(I) di-imine complexes, Dalton Trans. 40
Jo ur
(2011) 3498–3509. https://doi.org/10.1039/c0dt01383h. [23] M. V. Werrett, D. Chartrand, J. D. Gale, G.S. Hanan, J. G. MacLellan, M. Massi,.S. Muzzioli, P. Raiteri, B. W. Skelton, M. Silberstein, S. Stagni, Synthesis, Structural, and Photophysical Investigation of Diimine Triscarbonyl Re(I) Tetrazolato Complexes, Inorg. Chem. 50 (2011) 1229–1241. https://doi.org/10.1021/ic1015516. [24] F. Zhao, W.-q. Liu, H.–y. Xia, Y.–b. Wang, Towards an understanding of the absorption and emission spectra of rhenium(I) tricarbonyl polypyridine complexes containing NO2 group: A density functional theory study, Comput. Theor. Chem. 997 (2012) 49–54. http://dx.doi.org/10.1016/j.comptc.2012.07.035. [25] X. Liu, H. Xia, W. Gao, Q. Wu, X. Fan, Y. Mu, C. Ma, New rhenium(I) complexes with substituted diimine ligands for highly efficient phosphorescent devices fabricated by
a
solution
process,
J.
https://doi.org/10.1039/c2jm14222h.
Mater.
Chem.
22
(2012)
3485–3492.
Journal Pre-proof [26] M. Kuss-Petermann, H. Wolf, D. Stalke, O. S. Wenger, Influence of Donor−Acceptor Distance Variation on Photoinduced Electron and Proton Transfer in Rhenium(I)−Phenol Dyads, J. Am.
Chem. Soc. 134 (2012) 12844−12854.
https://doi.org/10.1021/ja3053046. [27] C.–C. Ko, C.–O. Ng, S.–M. Yiu, Luminescent Rhenium(I) Phenanthroline Complexes with a Benzoxazol-2-ylidene Ligand: Synthesis, Characterization, and Photophysical
Study,
Organometallics
31
(2012)
7074−7084.
https://doi.org/10.1021/om300526e. [28] P.J. Wright, S. Muzzioli, M. V. Werrett, P. Raiteri, B. W. Skelton, D. S. Silvester,
of
S. Stagni, M. Massi, Synthesis, Photophysical and Electrochemical Investigation of Dinuclear Tetrazolato-Bridged Rhenium Complexes, Organometallics 31 (2012)
ro
7566−7578. https://doi.org/10.1021/om300870a.
-p
[29] R. Horvath, M. G. Fraser, S. A. Cameron, A. G. Blackman, P. Wagner, D. L. Officer, K. C. Gordon, Synthesis, Characterization, and Photophysics of Oxadiazole-
re
and Diphenylaniline-Substituted Re(I) and Cu(I) Complexes, Inorg. Chem. 52 (2013) 1304−1317. https://doi.org/10.1021/ic3018387.
lP
[30] W.–K. Chung, K. M.–C. Wong, W. H. Lam, X. Zhu, N. Zhu, H.–S. Kwok, V. W.– W. Yam, Syntheses, photophysical, electroluminescence and computational studies of
na
rhenium(I) diimine triarylamine-containing alkynyl complexes, New J. Chem. 37 (2013) 1753—1767. https://doi.org/10.1039/c3nj00121k.
Jo ur
[31] K. Potgieter, P. Mayer, T. Gerber, N. Yumata, E. Hosten, I. Booysen, R. Betz, M. Ismail, B. van Brecht, Rhenium(I) tricarbonyl complexes with multidentate nitrogen-donor
derivatives,
Polyhedron
49
(2013)
67–73.
http://dx.doi.org/10.1016/j.poly.2012.08.078. [32] S. F. Sousa, R. N. Sampaio, N. M. Barbosa Neto, A. E. H. Machado, A. O. T. Patrocinio,
The
photophysics
of
fac-[Re(CO)3(NN)(bpa)]+
complexes:
a
theoretical/experimental study, Photochem. Photobiol. Sci. 13 (2014) 1213–1224. https://doi.org/10.1039/c4pp00074a. [33] C. Daniel, Photochemistry and photophysics of transition metal complexes: Quantum
chemistry,
Coord.
Chem.
Rev.
282–283
(2015)
19–32.
http://dx.doi.org/10.1016/j.ccr.2014.05.023. [34] A. Woźna, A. Kapturkiewicz, The luminescence properties of the heteroleptic [Re(CO)3(N-N)Cl] and [Re(CO)3(N-N)(CH3CN)]+ complexes in view of the combined
Journal Pre-proof Marcus–Jortner and Mulliken–Hush formalism, Phys. Chem. Chem. Phys. 17 (2015) 30468—30480. https://doi.org/10.1039/c5cp05167c. [35] L. D. Ramos, R. N. Sampaio, F. F. de Assis, K. T. de Oliveira, P. Homem-de-Mello, A. O. T. Patrocinio, K. P. M. Frin, Contrasting photophysical properties of rhenium(I) tricarbonyl complexes having carbazole groups attached to the polypyridine
ligand,
Dalton
Trans.
45
(2016)
11688–11698.
https://doi.org/10.1039/c6dt01112h. [36] M. R. Gonçalves, K. P.M. Frin, Synthesis, characterization, photophysical and properties
phenanthroline)rhenium(I)
of
fac-tricarbonyl(4,7-dichloro-1,10-
complexes,
Polyhedron
97
of
electrochemical
112–117.
ro
http://dx.doi.org/10.1016/j.poly.2015.05.007.
(2015)
[37] M. R. Gonçalves, K. P.M. Frin, Synthesis, characterization, photophysical and
triphenylphosphine
-p
electrochemical properties of rhenium(I) tricarbonyl diimine complexes with ligand,
Polyhedron
132
(2017)
20–27.
re
http://dx.doi.org/10.1016/j.poly.2017.04.029.
[38] M. Pinto, I. Chakraborty, J. Martinez-Gonzalez, P. Mascharak, Synthesis and of
photoactive
rhenium
lP
structures
2-(pyridin-2-yl)-1,3-benzothiazole, Acta
na
1,10-phenanthroline,
carbonyl
complexes
derived
2-(quinolin-2-yl)-1,3-benzothiazole
Cryst.
C73
(2017)
from and
923–929.
https://doi.org/10.1107/S2053229617014644.
Jo ur
[39] A. Carreño, E. Solis-Céspedes, D. Páez-Hernández, R. Arratia-Pérez, Exploring the geometrical and optical properties of neutral rhenium (I) tricarbonyl complex of 1,10-phenanthroline-5,6-diol using relativistic methods, Chem. Phys. Lett. 685 (2017) 354–362. http://dx.doi.org/10.1016/j.cplett.2017.07.058. [40] W.–K. Chung, M. Ng, N. Zhu, S. K.–L. Siu, V. W.–W. Yam, Synthesis, characterization and computational studies of luminescent rhenium(I) tricarbonyl diimine complexes with 8-hydroxyquinolinecontaining alkynyl ligands, J. Organomet. Chem. 847 (2017) 278-288. http://dx.doi.org/10.1016/j.jorganchem.2017.04.016. [41] L. A. Faustino, A. E. H. Machado, A. O. T. Patrocinio, Photochemistry of fac-[Re(CO)3(dcbH2)(trans-stpy)]+: New Insights on the Isomerization Mechanism of Coordinated
Stilbene-like
Ligands,
Inorg.
https://doi.org/10.1021/acs.inorgchem.8b00093.
Chem.
57
(2018)
2933−2941.
Journal Pre-proof [42] R. C. Amaral, L. S. Matos, K. P. S. Zanoni, N. Y. M. Iha, Photoreversible Molecular Motion of stpyCN Coordinated to fac-[Re(CO)3(NN)]+ Complexes, J. Phys. Chem. A 122 (2018) 6071−6080. https://doi.org/10.1021/acs.jpca.8b02630. [43] P. Kurz, B. Probst, B. Spingler, R. Alberto, Ligand Variations in [ReX(diimine)(CO)3] Complexes: Effects on Photocatalytic CO2 Reduction, Eur. J. Inorg. Chem. (2006) 2966–2974. https://doi.org/10.1002/ejic.200600166. [44] L. T.-L. Lo, S.-W. Lai, S.-M. Yiu, C.-C. Ko, A new class of highly solvatochromic dicyano rhenate(I) diimine complexes – synthesis, photophysics and photocatalysis, Chem. Commun. 49 (2013) 2311-2313. https://doi.org/10.1039/c3cc39189b.
of
[45] Y.Kuramochi, O. Ishitani, H. Ishida, Reaction mechanisms of catalytic
ro
photochemical CO2 reduction using Re(I) and Ru(II) complexes, Coord. Chem. Rev. 373 (2018) 333–356. https://doi.org/10.1016/j.ccr.2017.11.023.
-p
[46] N.M. Orchanian, L. E. Hong, J. A. Skrainka, J. A. Esterhuizen, D. A. Popov, S. C. Marinescu, Surface-Immobilized Conjugated Polymers Incorporating Rhenium
re
Bipyridine Motifs for Electrocatalytic and Photocatalytic CO2 Reduction, ACS Appl. Energy Mater. 2 (2019) 110−123. https://doi.org/10.1021/acsaem.8b01745.
lP
[47] D. R. Whang, D. H. Apaydin, S. Y. Park, N. S. Sariciftci, An electron-reservoir Re(I) complex for enhanced efficiency for reduction of CO2 to CO, Journal of Catalysis
na
363 (2018) 191–196. https://doi.org/10.1016/j.jcat.2018.04.028. [48] K. K. Lo, W. Hui, D. C. Ng, K. Cheung, Synthesis, Characterization,
Jo ur
Photophysical Properties, and Biological Labeling Studies of a Series of Luminescent Rhenium(I) Polypyridine Maleimide Complexes, Inorg. Chem. 41 (2002) 40-46. https://doi.org/10.1021/ic010602z. [49] L. Liu, X. Li, S. Hou, Y. Xue, Y. Yao, Y. Ma, X. Feng, S. He, Y. Lu, Y. Wang, X. Zeng, Nanoparticles from block copolymer encapsulating Re(phen) complexes as bifunctional agents for cell imaging and gene transfection, Chem. Commun. (2009) 6759–6761. https://doi.org/10.1039/b912659g. [50] R.-R. Ye, C.-P. Tan, M.-H. Chen, L. Hao, L.-N. Ji, Z.-W. Mao, Mono- and Dinuclear Phosphorescent Rhenium(I) Complexes: Impact of Subcellular Localization on
Anticancer
Mechanisms,
Chem.
Eur.
J.
22
(2016)
7800
–
7809.
https://doi.org/10.1002/chem.201505160. [51] V. Fernandez-Moreira, F. L. Thorp-Greenwood and M. P. Coogan, Application of d6 transition metal complexes in fluorescence cell imaging, Chem. Commun. 46 (2010) 186–202. https://doi.org/10.1039/b917757d.
Journal Pre-proof [52] S. Clède, C. Policar, Metal–Carbonyl Units for Vibrational and Luminescence Imaging: Towards Multimodality, Chem. Eur. J. 21 (2015) 942 – 958. https://doi.org/10.1002/chem.201404600. [53] V.Fernandez-Moreira, F. L. Thorp-Greenwood, A. J. Amoroso, J. Cable, J. B. Court, V. Gray, A. J. Hayes, R. L. Jenkins, B. M. Kariuki, D. Lloyd, C. O. Millet, C. Ff. Williams, M. P. Coogan, Uptake and localisation of rhenium fac-tricarbonyl polypyridyls in fluorescent cell imaging experiments, Org. Biomol. Chem. 8 (2010) 3888–3901. https://doi.org/10.1039/c004610h. [54] K. K-W Lo , M -W. Louie, K. Y. Zhang, Design of luminescent iridium(III) and
Coord.
Chem.
Rev.
254
(2010)
2603–2622.
ro
probes,
of
rhenium(I) polypyridine complexes as in vitro and in vivo ion, molecular and biological
https://doi.org/10.1016/j.ccr.2010.01.014.
-p
[55] S. Hostachy, C. Policar, N. Delsuc, Re(I) carbonyl complexes: Multimodal platforms for inorganic chemical biology, Coord. Chem. Rev. 351 (2017) 172–188.
re
http://dx.doi.org/10.1016/j.ccr.2017.05.004.
[56] J. Skiba, A. Kowalczyk, Paweł St czek, T. erna , D. Trzybi ski, K.Woźniak, U.
lP
Schatzschneider, R. Czerwieniec, K. Kowalski, Luminescent fac-[Re(CO)3(phen)] carboxylato complexes with non-steroidal anti-inflammatory drugs: synthesis and insights
into
the
na
mechanistic
fac-[Re(CO)3(phen)(aspirin)],
in
New
vitro
J.Chem.
anticancer 43
activity
(2019)
of
573-583.
Jo ur
https://doi.org/10.1039/c8nj05494k.
[57] J. Skiba , T. erna , D. Trzybi ski, K. Woźniak , G. Ferraro , D. Marasco, A. Merlin, M. Z. Shafikov, R. Czerwieniec, K. Kowalski, Mitochondria Targeting with Luminescent
Rhenium(I)
Complexes,
Molecules
22
(2017)
809.
https://doi.org/10.3390/molecules22050809. [58] X. Li , D. Zhanga , W. Li , B. Chua, L. Hana, T. Li , Z. Su , J. Zhu , S. Wu , Y.Chen , P. Lei , Z. Hu , Z. Zhang, New rhenium complexes containing 4,5-diazafluorene ligand for high-efficiency green electrophosphorescence, Synthetic Metals 159 (2009) 1340–1344. https://doi.org/10.1016/j.synthmet.2009.03.004. [59] S. K. Mizoguchi, A. O. Toledo Patrocínio, N. Y. Murakami Iha, On the energy transfer from a polymer host to the rhenium(I) complex in OLEDs, Synthetic Metals 159 (2009) 2315–2317. https://doi.org/10.1016/j.synthmet.2009.08.046. [60] X. Li, D. Zhang, G. Lu, G. Xiao, H. Chi, Y. Dong, Z. Zhang, Z. Hu, Synthesis and characterization of novel rhenium (I) complexes with large Stokes shift for applications
Journal Pre-proof in organic electroluminescent devices, J. Photoch. Photobio A 241 (2012) 1–7. http://dx.doi.org/10.1016/j.jphotochem.2012.05.017. [61] T. Yu, D. P -K. Tsang, V. K -M. Au, W. H.Lam, M -Y. Chan, V. W -W. Yam, Deep
Red
to
Near-Infrared
Emitting
Rhenium(I)
Complexes:
Synthesis,
Characterization, Electrochemistry, Photophysics, and Electroluminescence Studies, Chem. Eur. J. 19 (2013) 13418 – 13427. https://doi.org/10.1002/chem.201301841. [62] M. Rui, W.Yuhong, W.Yinting, Z. Na, Phosphorescent rhenium emitters based on two
electron-withdrawing
electroluminescent
diamine
performance,
ligands: J.
Structure,
Lumin.
153
(2014)
and
264–271.
of
http://dx.doi.org/10.1016/j.jlumin.2014.03.051.
characterization
ro
[63] G -W. Zhao, J -H. Zhao ,Y -X. Hu ,D -Y. Zhang , X. Li, Recent advances of neutral rhenium(I) tricarbonyl complexes for application in organic light-emitting diodes, Metals
212
(2016)
131–141.
-p
Synthetic
http://dx.doi.org/10.1016/j.synthmet.2015.12.014.
re
[64] X. Liu, H. Xia, W. Gao, Q. Wu, X. Fan, Y. Mu, C. Ma, New rhenium(I) complexes with substituted diimine ligands for highly efficient phosphorescent devices fabricated a
solution
proces,
J.
Mater.
lP
by
Chem.
22
(2012)
3485.
https://doi.org/10.1039/c2jm14222h.
na
[65] Y -X. Hu , G -W. Zhao , Y. Dong , Y -L. Lü , X. Li , D -Y. Zhang, New rhenium(I) complex with thiadiazole-annelated 1,10-phenanthroline for highly efficient OLEDs,
Dyes
Jo ur
phosphorescent
and
Pigments
137
(2017)
569-575.
https://doi.org/10.1016/j.dyepig.2016.10.048. [66] A. Juris,
S. Campagna, I.Bidd, J -M. Lehn, R. Ziessel, Synthesis and
Photophysical and Electrochemical Properties of New Halotricarbony1 (poly pyridine) rhenium (I) Complexes, Inorg.Chem. 27 (1988) 4007-4011. [67] M. P. Coogan, V. Fernandez-Moreira, B. M. Kariuki, S. J. A. Pope, F L. Thorp-Greenwood, A Rhenium Tricarbonyl 4’-Oxo-terpy Trimer as a Luminescent Molecular Vessel with a Removable Silver Stopper, Angew. Chem. Int. Ed. 48 (2009) 4965 –4968. https://doi.org/10.1002/anie.200900981. [68] A. J. Amoroso, A. Banu, M. P. Coogan, P. G. Edwards, G. Hossain, K. M. A. Malik, Functionalisation of terpyridine complexes containing the Re(CO)3+ moiety, Dalton Trans. 39 (2010) 6993–700. https://doi.org/10.1039/c0dt00174k.
Journal Pre-proof [69] D. R. Black, S. E. Hightower, Preparation and characterization of rhenium(I) dicarbonyl complexes based on the meridionally-coordinated terpyridine ligand, Inorg. Chem. Commun. 24 (2012) 16–19. https://doi.org/10.1016/j.inoche.2012.07.034. [70] D. Wang, Q.–L. Xu, S. Zhang, H.–Y. Li, C.–C. Wang, T.–Y. Li, Y.–M. Jing, W. Huang, Y.–X. Zheng, G. Accors, Synthesis and photoluminescence properties of rhenium(I)
complexes
hole-transporting
based
units,
on
2,2’:6’,2’’-terpyridine
Dalton
Trans.
42
derivatives
(2013)
with
2716–2723.
https://doi.org/10.1039/c2dt32154h. [71] B.A. Frenzel, J. E. Schumaker, D. R. Black, S. E. Hightower, Synthesis,
of
spectroscopic, electrochemical and computational studies of rhenium(I) dicarbonyl
ro
complexes based on meridionally-coordinated 2,2’:6’,2’’-terpyridine, Dalton Trans. 42 (2013) 12440–12451. https://doi.org/10.1039/c3dt51251g.
-p
[72] P. Bulsink, A. Al-Ghamdi, P. Joshi, I. Korobkov, T. Woo, D. Richeson, Capturing Re(I) in an neutral N,N,N pincer Scaffold and resulting enhanced absorption of visible
re
light, Dalton Trans. 45 (2016) 8885–8896. https://doi.org/10.1039/c6dt00661b. [73] G. Velmurugan, P. Venuvanalingam, Luminescent Re(I) terpyridine complexes
lP
for OLEDs: what does the DFT/TD-DFT probe reveal?, Dalton Trans. 44 (2015) 8529– 8542. https://doi.org/10.1039/c4dt02917h.
na
[74] T. Klemens, A. Świtlicka-Olszewska, B. Machura, M. Grucela, E. Schab-Balcerzak, K. Smolarek, S. Mackowski, A. Szlapa, S. Kula, S. Krompiec, P.
Jo ur
Lodowski, A. Chrobok, Rhenium(I) terpyridine complexes – synthesis, photophysical properties and application in organic light emitting devices, Dalton Trans. 45 (2016) 1746–1762. https://doi.org/10.1039/c5dt04093k. [75] T. Klemens, A. Switlicka-Olszewska, B. Machura, M. Grucela, H.Janeczek, E.Schab-Balcerzak, A. Szlapa, S. Kula, S. Krompiec, K. Smolarek, D.Kowalska, S. Mackowski, K. Erfurt, P. Lodowski, Synthesis, photophysical properties and application in organic light emitting devices of rhenium(I) carbonyls incorporating functionalized 2,20:60,200-terpyridines,
RSC
Adv. 6 (2016) 56335–56352.
https://doi.org/10.1039/c6ra08981j. [76] T. Klemens, K. Czerwi ska, A. Szlapa-Kula, S. Kula, A. Świtlicka, S. Kotowicz, M. Siwy, K. Bednarczyk, S.Krompiec, K. Smolarek, S. Maćkowski, W. Danikiewicz, E. Schab-Balcerzak, B. Machura, Synthesis, spectroscopic, electrochemical and computational
studies
of
rhenium(I)
tricarbonyl
complexes
based
on
Journal Pre-proof bidentate-coordinated 2,6-di(thiazol-2-yl)pyridine derivatives,
Dalton Trans. 46
(2017) 9605–9620. https://doi.org/10.1039/c7dt01948c. [77] T. Klemens, A. Świtlicka, A.Szlapa‐ Kula, S.Krompiec, P. Lodowski, A. Chrobok, M. Godlewska, S. Kotowicz, M. Siwy, K. Bednarczyk, M. Libera, S. Maćkowski, T. Pędzi ski, E. Schab‐ Balcerzak, B. Machura, Experimental and computational exploration of photophysical and electroluminescent properties of modified 2,2′:6′,2″‐ terpyridine, 2,6‐ di(thiazol‐ 2‐ yl)pyridine and 2,6‐ di(pyrazin‐ 2‐ yl)pyridine ligands and their Re(I) complexes, Appl. Organometal. Chem. 32 (2018) e4611. https://doi.org/10.1002/aoc.4611. . Machura , S. Kula, S. Krompiec, K. Łaba, M.
of
[78] T. Klemens, A. Świtlicka,
Korzec, M. Siwy, H. Janeczek, E. Schab-Balcerzak, M. Szalkowski, J. Grzelak, S.
ro
Maćkowski, A family of solution processable ligands and their Re(I) complexes 163 (2019) 86–101.
-p
towards light emitting applications, Dyes and Pigments https://doi.org/10.1016/j.dyepig.2018.11.035.
re
[79] T. Klemens, A. Świtlicka, S. Kula, M. Siwy, K. Łaba, J. Grzelak, M. Szalkowski, S. Maćkowski, E. Schab-Balcerzak, B. Machura, The effect of 2-, 3- and 4-pyridyl on
photophysics
of
fac-[ReCl(CO)3(n-pytpy-κ2N)]
lP
substituents
complexes:
Experimental and theoretical insights, Journal of Luminescence 209 (2019) 346–356.
na
https://doi.org/10.1016/j.jlumin.2019.01.045. [80] K. Sua, N. Suna, X. Tiana, S. Guob, Z. Yana, D. Wanga, H. Zhoua, X. Zhaoa, C.
Jo ur
Chen, Highly soluble polyimide bearing bulky pendant diphenylamine-pyrene for fast-response electrochromic and electrofluorochromic applications, Dyes and Pigments 171 (2019) 107668. https://doi.org/10.1016/j.dyepig.2019.107668. [81] F. Wu, J. Liu, G. Wang, Q. Song, L. Zh, m-Methoxy Substituents in a Tetraphenylethylene-Based Hole-Transport Material for Efficient Perovskite Solar Cells, Chem. Eur. J. 22 (2016) 16636–1664. https://doi.org/10.1002/chem.201603672. [82] B. J. Neyhouse, T. A. White, Modifying the steric and electronic character within Re(I)-phenanthroline complexes for electrocatalytic CO2 reduction, Inorg. Chim. Acta 479 (2018) 49–57. https://doi.org/10.1016/j.ica.2018.04.008. [83] X. Li, X. Liu, Z. Wu, H. Zhang, DFT/TDDFT Studies on the Electronic Structures and Spectral Properties of Rhenium(I) Pyridinybenzoimidazole Complexes, J. Phys. Chem. A 112 (2008) 11190–11197. https://doi.org/10.1021/jp805758n. [84] X.-K. Chen, L.-Y. Zou, A.-M. Ren, J.-X. Fan, How dual bridging atoms tune structural and optoelectronic properties of ladder-type heterotetracenes?—a theoretical
Journal Pre-proof study,
Phys.
Chem.
Chem.
Phys.
13
(2011)
19490–19498.
https://doi.org/10.1039/C1CP22227A. [85] G. Velmurugan, B. K. Ramamoorthi, P. Venuvanalingam, Are Re(I) phenanthroline complexes suitablecandidates for OLEDs? Answers from DFT and TD-DFT investigations, Phys. Chem. Chem. Phys. 16 (2014) 21157—21171. https://doi.org/10.1039/c4cp01135j. [86] L. D. Ramos, R. N. Sampaio, F. F. de Assis, K. T. de Oliveira, P. Homem-de-Mello, A. O. T. Patrocinio, K. P. M. Frin, Contrasting photophysical properties of rhenium(I) tricarbonyl complexes having carbazole groups attached to the ligand,
Dalton
Trans.
45
(2016)
of
polypyridine
ro
https://doi.org/10.1039/c6dt01112h.
11688–11698.
[87] P. Bujak, I. Kulszewicz-Bajer, M. Zagorska, V. Maurel, I. Wielgus, A. Pron,
https://doi.org/10.1039/c3cs60257e.
-p
Polymers for electronics and spintronics, Chem. Soc. Rev. 42 (2013) 8895—8999.
Yarnton, Oxfordshire, England 2014.
re
[88] Agilent Technologies, CrysAlisPRO and CrysAlisRED, Oxford Diffraction Ltd.
lP
[89] G. M. Sheldrick, Crystal structure refinement with SHELXL, Acta Cryst. C 71 (2015) 3–8. https://doi.org/10.1107/S2053229614024218.
na
[90] Gaussian 16, Revision C.01, M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, G. A. Petersson, H.
Jo ur
Nakatsuji, X. Li, M. Caricato, A. V. Marenich, J. Bloino, B. G. Janesko, R. Gomperts, B. Mennucci, H. P. Hratchian, J. V. Ortiz, A. F. Izmaylov, J. L. Sonnenberg, D. Williams-Young, F. Ding, F. Lipparini, F. Egidi, J. Goings, B. Peng, A. Petrone, T. Henderson, D. Ranasinghe, V. G. Zakrzewski, J. Gao, N. Rega, G. Zheng, W. Liang, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, K. Throssell, J. A. Montgomery, Jr., J. E. Peralta, F. Ogliaro, M. J. Bearpark, J. J. Heyd, E. N. Brothers, K. N. Kudin, V. N. Staroverov, T. A. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. P. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, J. M. Millam, M. Klene, C. Adamo, R. Cammi, J. W. Ochterski, R. L. Martin, K. Morokuma, O. Farkas, J. B. Foresman, and D. J. Fox, Gaussian, Inc., Wallingford CT, 2016. [91] C. Adamo, V. Barone, Toward reliable density functional methods without adjustable parameters: The PBE0 model, https://doi.org/10.1063/1.478522.
J. Chem. Phys. 110 (1999) 6158-69.
Journal Pre-proof [92] M. Ernzerhof, G. E. Scuseria, Assessment of the Perdew-Burke-Ernzerhof exchange-correlation
functional,
J
Chem.
Phys.
110
(1999)
5029-5036.
https://doi.org/10.1063/1.478401. [93] R. Ditchfield, W. J. Hehre, and J. A. Pople, Self-Consistent Molecular Orbital Methods. 9. Extended Gaussian-type basis for molecular-orbital studies of organic molecules, J. Chem. Phys. 54 (1971) 724. https://doi.org/10.1063/1.1674902. [94] W. J. Hehre, R. Ditchfield, J. A. Pople, Self-Consistent Molecular Orbital Methods. 12. Further extensions of Gaussian-type basis sets for use in molecular-orbital studies of organic-molecules, J. Chem. Phys. 56 (1972) 2257.
of
https://doi.org/10.1063/1.1677527.
ro
[95] M. M. Francl, W. J. Pietro, W. J. Hehre, J. S. Binkley, D. J. DeFrees, J. A. Pople, M. S. Gordon, Self-Consistent Molecular Orbital Methods. 23. A polarization-type
-p
basis set for 2nd-row elements, J. Chem. Phys. 77 (1982) 3654-3665. https://doi.org/10.1063/1.444267.
re
[96] T. Clark, J. Chandrasekhar, G. W. Spitznagel, P. v. R. Schleyer, Efficient diffuse function-augmented basis-sets for anion calculations. 3. The 3-21+G basis set for
lP
1st-row elements, Li-F, J. Comp. Chem., 4 (1983) 294-301. [97] V. A. Rassolov, M. A. Ratner, J. A. Pople, P. C. Redfern, L. A. Curtiss, 6-31G*
na
Basis Set for Third-Row Atoms, J. Comp. Chem. 22 (2001) 976-984. https://doi.org/10.1063/1.1674902.
Jo ur
[98] E. Cancès, B. Mennucci, J. Tomasi, A new integral equation formalism for the polarizable continuum model: theoretical background and applications to isotropic and anisotropic
dielectrics,
J.
Chem.
Phys.
107
(1997)
3032-3041.
http://dx.doi.org/10.1063/1.474659. [99] B. Mennucci, J. Tomasi, Continuum solvation models: a new approach to the problem of solute's charge distribution and cavity boundaries, J Chem Phys, 106 (1997) 5151-5158. https://doi.org/10.1063/1.473558.
[100] M. Cossi, V. Barone, B. Mennucci, J. Tomasi, Ab initio study of ionic solutions by a polarizable continuum dielectric model, Chem Phys Lett, 286 (1998) 253-260. https://doi.org/10.1016/S0009-2614(98)00106-7.
Journal Pre-proof
Author Contribution:
Magdalena Małecka recorded NMR and IR spectra, took part into the studies concerning electronic absorption and emission spectroscopy in solution and in solid as powder as well as participated in the discussion of the optical properties. She also carried out part of DFT computations and participated in the discussion of the correlation between computation results and optical properties.
of
Anna Świtlicka carried out the synthesis of rhenium(I) complexes and performed the X-Ray measurements.
-p
ro
Grażyna Szafraniec-Gorol - synthesized 2,2′:6′,2′′-terpyridine, 2,6-di(thiazol-2-yl)pyridines and 2,6-di(pyrazin-2-yl)pyridine ligands.
na
lP
re
Sonia Kotowicz - carried out the electrochemical measurements and discussed the electrochemical properties of the Re(I) carbonyl complexes, as well as she took part into the studies concerning electronic absorption and emission spectroscopy in solution and in solid as powder.
Jo ur
Mariola Siwy prepared the film and blend from complexes, registered their UV-vis and PL spectra, as well as fabricated diodes for EL measurements and recorded EL spectra
Marcin Szalkowski and Sebastian Madkowski are responsible for electroluminescence studies.
Ewa Schab-Balcerzak designed as well as discussed the results from UV-Vis, PL and EL measurements of the synthesized ligands and complexes in solid state as film and blend.
Barbara Machura designed as well as discussed the results from X-Ray analysis and investigations of optical properties of solution and solid state.
Barbara Machura and Ewa Schab-Balcerzak co-produced the manuscript.
Journal Pre-proof
Declaration of interests
☒ The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.
re
-p
ro
of
☐The authors declare the following financial interests/personal relationships which may be considered as potential competing interests:
Highlights
lP
Graphical abstract
Jo ur
na
Series of Re(I) carbonyls complexes with terpy-like ligands were investigated. Their photophysical properties were explored by carrying out absorption and emission studies Impact of the triimine core and aryl substituent decorated with methoxy groups was discussed. The capacity of the obtained Re(I) complexes for electroluminescence was tested.