Anna Świtlickaa,
Tomasz Klemensa,
Barbara Machura*a,
Ewa Schab-Balcerzak*ab,
Katarzyna Lababc,
Mieczyslaw Lapkowskibc,
Marzena Grucelab,
Jacek Nycza,
Marcin Szalaa and
Magdalena Kaniad
aInstitute of Chemistry, University of Silesia, 9 Szkolna Str., 40-006 Katowice, Poland. E-mail: bmachura@poczta.onet.pl
bCentre of Polymer and Carbon Materials, Polish Academy of Sciences, 34 M. Curie-Sklodowska Str., 41-819 Zabrze, Poland. E-mail: ewa.schab-balcerzak@us.edu.pl
cSilesian University of Technology, Faculty of Chemistry, 9 Strzody Str., 44-100 Gliwice, Poland
dMass Spectrometry Group, Institute of Organic Chemistry, Polish Academy of Sciences, Kasprzaka 44/52, PO Box 58, 01-224 Warszawa 42, Poland
First published on 18th November 2016
Five rhenium(I) tricarbonyl complexes incorporating 1,10-phenanthroline derivatives with electron-withdrawing Cl and electron-donating CH3 substituents were synthesized, and their photophysical, thermal, and electrochemical properties, with electroluminescence ability were examined. The melting temperature of these complexes was found to decrease from 391 to 281 °C upon adding methyl groups and decreasing the number of chlorine atoms. Compounds bearing methyl substituents could form amorphous molecular materials with glass transition temperatures in the range 118–127 °C. Cyclic voltammetry measurements demonstrated that the complexes are electrochemically active with low energy band gaps in the range 2.24–2.46 eV. All complexes are photoluminescent, form films in the solid state, and blend with poly(9-vinylcarbazole) (PVK) and PVK:(2-(4-tert-butylphenyl)-5-(4-biphenylyl)-1,3,4-oxadiazole) (PBD), emitting light with λem from 560 to 599 nm. They exhibited large Stokes shifts of up to 323 nm in solutions and up to 220 nm as film on a glass substrate. They were tested as guests in organic light-emitting diodes. As hosts, PVK and a binary matrix consisting of PVK with PBD were applied. The ability of the investigated complexes to emit light under an applied voltage was shown.
The continuing development of luminescent coordination compounds has largely been motivated by their possible use in catalysis,3 solar energy conversion,4 luminescence sensing,5 light-emitting diodes (LEDs),6 nonlinear optical materials,7 and molecular sensors.8
Since the first observation of electronically excited luminescence in [ReCl(CO)3(phen)], reported by Wrighton and Morse in 1974,9 particular attention has been devoted to rhenium(I) diimine tricarbonyls [Re(CO)3La(N–N)]n+ (La – axial ligand; n = 0 or 1). So far, a great number of investigations on the [Re(CO)3La(bipy)]n+ and [Re(CO)3La(phen)]n+ chromophores have appeared in the literature.10 The influence of structural variations of diimines, and the effect of the axial ligand on the photophysical and photochemical properties of rhenium(I) tricarbonyls have been investigated, both experimentally and computationally.11 For example, Rillema et al.12 examined the photobehavior of a series of Re(I) tricarbonyl complexes containing 2,6-dimethylphenylisocyanide and 5- and 6-derivatized phenanthroline ligands. Cardinaels13 and Pope14 demonstrated the effect of substitution patterns on the photoluminescent properties of rhenium(I) complexes incorporating substituted imidazo[4,5-f]-1,10-phenanthrolines.
Due to their remarkable photochemical and photophysical properties, diimine rhenium(I) tricarbonyls are still considered as excellent candidates for various applications in areas ranging from electronic and energy transfer to catalysis and medicinal chemistry. In this context, there is a need for the development of new luminescent rhenium(I) coordination compounds.
Herein, we report the synthesis, characterization, photophysical, thermal, and electrochemical properties of three novel rhenium(I) tricarbonyl complexes incorporating 1,10-phenanthroline derivatives: 2,5,9-trimethyl-4,7-dichloro-1,10-phenanthroline (2,5,9-(Me)3-4,7-Cl2phen), 2,9-dimethyl-4,7-dichloro-1,10-phenanthroline (2,9-(Me)2-4,7-Cl2phen), and 4,5,7-trichloro-1,10-phenanthroline (4,5,7-Cl3phen). For a better understanding of the effect of the number and location of chloro substituents in 1,10-phenanthroline, rhenium(I) tricarbonyls with 4,7-dichloro-1,10-phenanthroline (4,7-Cl2phen)2a,2b,9,15e and 5-chloro-1,10-phenanthroline (5-Clphen)2a,2b,9 were also included in the studies.
The luminescent properties of Re(I) complexes were studied in solution (CHCl3 and CH3CN) and solid state at room temperature, as well as in 4:
1 EtOH–MeOH glass matrix at 77 K. To obtain a detailed insight into their electronic structures and spectroscopic properties, density functional theory (DFT) and time-dependent DFT (TD-DFT) calculations were performed.
The infrared spectra of 1–5 display three intense bands ν(CO) in the region 1876–2025 cm−1 with a typical pattern for complexes incorporating the fac-[Re(CO)3]+ moiety, two overlapping lower-energy bands (1937–1876 cm−1), and a sharp intense band at higher wavenumbers (2018–2025 cm−1). The presence of the phenanthroline ligand is shown in the IR spectra by medium intensity bands in the range 1617–1560 cm−1, corresponding to the ν(CN) and ν(CC) modes.
The 13C NMR spectra of 1–5 exhibit characteristic resonances around 198 and 189 ppm, corresponding to the carbonyl groups. Distinctive signals of the alkyl protons of the ligands 2,5,9-(Me)3-4,7-Cl2phen (in 1) and 2,9-(Me)2-4,7-Cl2phen (in 2) occur in the range 3.23–3.14 ppm, while the aromatic protons of the complexes of 1–5 were detected between 8.10 and 9.52 ppm (details are given in the Experimental part).
![]() | ||
Fig. 1 A perspective view showing the asymmetric units of 1, 3, and 5 with the atom numbering. Displacement ellipsoids are drawn at 50% probability. |
The structures 1, 3, and 5 are stabilized by short contacts, C–H⋯Cl, C–H⋯O, and C–H⋯N (Table S3†), which can be considered as weak hydrogen bonds as well as by π–π type interactions. As demonstrated by the crystal packing analysis (Mercury 2.4 program16), the molecules in structures 1 and 3 are linked into dimers by π–π type interactions with centroid-to-centroid separations of 3.883(5) Å and 3.986(5) Å in 1 and 3.869(7) Å in 3 (Fig. S1†). In the crystal lattice of 5, the molecules of [ReCl(CO)3(5-Clphen)] are extended into a 1D supramolecular network through two different π⋯π type interactions with centroid-to-centroid separations of 3.755(3) Å and 3.744(3) Å (Fig. S1†).
Code | Eox [V] | Eox,onset [V] | Ered1 [V] | Ered2 [V] | Ered3 [V] | Ered,onset [V] | IP [eV] | EA [eV] | Eg [eV] |
---|---|---|---|---|---|---|---|---|---|
a IP = −5.1 − Eox,onset, EA = −5.1 − Ered,onset, Eg = IP − EA. | |||||||||
1 | 1.06 | 0.95 | −1.56 | −1.73 | −1.86 | −1.47 | −6.05 | −3.63 | 2.42 |
2 | 1.06 | 0.95 | −1.59 | −2.19 | — | −1.51 | −6.05 | −3.59 | 2.46 |
3 | 1.03 | 0.94 | −1.38 | −1.89 | −2.06 | −1.30 | −6.04 | −3.80 | 2.24 |
4 | 1.02 | 0.93 | −1.49 | −2.09 | — | −1.42 | −6.03 | −3.68 | 2.35 |
5 | 0.99 | 0.91 | −1.64 | −2.12 | — | −1.56 | −6.01 | −3.54 | 2.47 |
The data from Table 1 indicate that the oxidation potentials were relatively insensitive to changes in the phenanthroline substituents, and a slight increasing tendency was observed in the order 5 < 4 < 3 < 2 < 1. The oxidation onsets of the Re(I) complexes are very close to each other. Based on the obtained data, a reduction process can be associated with the phenanthroline ligand, while the anodic wave can be associated with the Re(II/I) oxidation process. The second oxidation peak of rhenium Re(III/II) was not observed, probably due to the electrochemical window of acetonitrile. The oxidation process of the Re(I) complexes was irreversible, excluding the complexes with electron-donating group (CH3), that is, compounds 1 and 2. The first reduction process was quasi-reversible for all Re(I) complexes. To provide insight into the electroluminescent device fabrication of the investigated complexes, the IP/HOMO and the EA/LUMO were estimated. The EA decreases as the number of electron-withdrawing substituents increases, resulting in a decreased band gap in the order 5 > 4 > 3. Electron-donating –CH3 groups increase the EA of 1 and 2 compared to 4; however, the energy band gaps (Eg) obtained for complexes 1, 2, and 4 are very similar. The band-gap decreases in the order 5 > 2 > 1 > 4 > 3. The HOMO is determined mainly by the orbitals of the metal unit, while the LUMO is determined mainly by the orbitals of the ligand unit.15e It should be stressed that all Re(I) complexes exhibited a value of Eg useful for optoelectronic applications.
Complex | Medium | λ [nm] (103 ε [M−1 cm−1]) |
---|---|---|
1 | CH3CN | 387.2 (6.16), 332.5 (9.10), 285.0 (28.57), 215.9 (55.60), 200.1 (53.08) |
CHCl3 | 404.6 (2.98), 339.6 (3.92), 282.2 (17.80), 246.2 (18.88) | |
Film | 396 | |
2 | CH3CN | 382.1 (5.66), 330.3 (7.15), 270.7 (38.85), 231.4 (42.39), 215.6 (66.61), 196.1 (53.04) |
CHCl3 | 373.8 (3.68), 272.8 (24.72) | |
Film | 364 | |
3 | CH3CN | 393.1 (8.85), 313.7 (12.5), 274.9 (43.77), 229.1 (46.99), 214.9 (64.11), 197.5 (55.48) |
CHCl3 | 408.9 (4.48), 313.9 (6.8), 277.8 (28.04) | |
Film | 405 | |
4 | CH3CN | 380.4 (6.30), 268.3 (47.30), 212.8 (63.24) |
CHCl3 | 399.9 (4.60), 270.9 (36.92) | |
Film | 393 | |
5 | CH3CN | 378.8 (6.05), 290.5 (19.72), 269.4 (40.40), 218.0 (48.16), 200.3 (63.46) |
CHCl3 | 393.9 (3.32), 268.0 (37.24), 240.3 (37.04) | |
Film | 385 |
With reference to previous spectroscopic studies for diimine rhenium(I) tricarbonyls, the intense absorptions at high energy (200–320 nm) are predominately attributable to intraligand (IL) transitions of the diimine moieties, while the moderately intense lowest energy absorption band is typically from the spin-allowed metal dπ(Re) to ligand charge transfer.
As can be expected, the position and intensity of the lowest energy absorption of 1–5 is solvent dependent, shifting to higher energy in more polar media by 14 nm for 1, 10 nm for 2, 16 nm for 3, 16 nm for 4, and 15 nm for 5, and the MLCT band is sensitive to the number and nature of substituents in the 1,10-phenanthroline ring. The extent of the red shift was in the order 3 > 4 > 5, with an increasing number of chloro substituents, which draw electron density from the phenanthroline ring.
To obtain a deeper understanding of the nature of excited states involved in absorption processes, time-dependent DFT (TDDFT) calculations at the B3LYP/def2-TZVPD level were performed for 1–5. The calculated and measured absorption spectra in acetonitrile solution are graphically compared in Fig. 3. It is observed that the main spectral features of 1–5 are predicted to a good accuracy, both in position and relative intensities, by the TDDFT calculations. Also, the structural parameters of the optimized structures of 1, 3, and 5 coincide well with the experimental bond lengths and angles (see Table S2†). The calculated absorption energies associated with their oscillator strengths, their main contributions, and their assignments to the experimental results of 1–5 are given in Tables S4–S8 (see ESI†). Fig. 4 presents the molecular orbital energy level graph for the examined rhenium(I) complexes, and the selected molecular orbitals of 1–5 are shown in Fig. S4 and characterized in Tables S9–S13.†
![]() | ||
Fig. 3 Experimental (black) and calculated (red) electronic absorption spectra of 1–5 in CH3CN solution. |
As can be seen from Fig. 4, the LUMO energy decreases as the number of chloro substituents increases, resulting in a decreased band gap in the order 5 > 4 > 3, which then leads to an increase in the λmax of MLCT absorption (Table 2). On the other hand, electron-releasing –CH3 groups increase the LUMO energy of 1 and 2 compared to 4, which leads to an increased HOMO–LUMO band gap. The DFT/B3LYP/def2-TZVPD calculations showed a negligible effect of the third methyl group on the energy of the HOMO and LUMO orbitals of 1 in relation to 2; the HOMO–LUMO band gap is the same for 1 and 2.
The HOMO and HOMO−1 levels of 1–5 are very close in energy, and these orbitals are predominately constituted by 5dyz and 5dxz rhenium orbitals in a bonding relation to the carbonyl π* orbitals and an antibonding arrangement with chlorine occupied p orbitals. The HOMO−2 orbitals of 1–5 contain significant 5dxy rhenium (45–60%) character, along with ∼30% contributions from the CO component, and they are about ∼0.5 eV lower in energy in relation to the highest occupied molecular orbital. The LUMO and LUMO+1 levels are predominately centered on the phenanthroline ligand and particularly the π* antibonding orbital. These results are consistent with previously published studies on Re-phen carbonyl complexes.1a,18
According to TDDFT calculations, the lowest energy absorption of 1–5 arises from one-electron excitations H−1 → L, H → L+1, and H−1 → L+1, assigned as metal-ligand-to-ligand charge transfer transitions (MLLCT). With an increasing number of chloro substituents, the calculated low-energy absorption shift to lower energy is in the order 5 > 4 > 3. On the contrary, electron-donating groups lead to blue shifts of the H−1 → L excitations for 1 and 2 in relation to 4. For complexes 1–3, transitions of MLLCT character (H−2 → L+1) were also calculated to contribute to the higher energy absorption bands, at 333 nm for 1, 327 nm for 2, and 314 nm for 3. Generally, however, the intense bands in the high energy region are largely attributed to intraligand transitions (IL) and ligand–ligand charge transfer (LLCT).
Complex | Medium | λex (nm) | λem (nm) | Stokes shift (cm−1) | τ, ns (weight) | χ2 | Φ (%) |
---|---|---|---|---|---|---|---|
a (4 : 1 v/v). | |||||||
1 | CHCl3 | 408 | 640 | 8885 | 42.80 | 1.054 | 0.69 |
CH3CN | 400 | 665 | 9962 | 15.38 | 1.114 | 0.14 | |
Solid state | 449 | 594; 716 | 5437, 8305 | 130.64 (37.39%), 386.69 (62.61%), 2.34 (29.87%), 41.84 (14.69%), 237.03 (55.43%) | 1.179, 0.975 | 0.85 | |
EtOH–MeOHa (77 K) | 402 | 540 | 6357 | 7070 (53.88%), 25![]() |
1.030 | — | |
Film | 390 | 599 | 8558 | — | — | — | |
Blend with PVK | 330 | 386; 566 | — | — | |||
Blend with PVK:PBD | 330 | 384; 566 | — | — | — | — | |
2 | CHCl3 | 374 | 640 | 11![]() |
39.72 | 1.124 | 0.58 |
CH3CN | 375 | 665 | 11![]() |
13.50 | 1.129 | 0.48 | |
Solid state | 473 | 618 | 4960 | 113.05 | 1.202 | 3.46 | |
EtOH–MeOHa (77 K) | 385 | 543 | 7558 | 1970 (12.80%), 6660 (59.94%), 16![]() |
1.097 | — | |
Film | 390 | 586 | 10![]() |
— | — | — | |
Blend with PVK | 330 | 387; 582 | — | — | — | ||
Blend with PVK:PBD | 330 | 386; 560 | — | — | — | ||
3 | CHCl3 | 408 | 680 | 9804 | 38.58 (76.48%), 141.89 (23.54%) | 1.145 | 0.57 |
CH3CN | 390 | 716 | 11![]() |
14.14 (78.38%), 57.25 (21.62%) | 0.974 | 0.50 | |
Solid state | 490 | 643 | 4856 | 119.09 | 1.133 | 4.11 | |
EtOH–MeOHa (77 K) | 373 | 569 | 9235 | 2590 (44.22%), 8720 (41.00%), 62![]() |
1.151 | — | |
Film | 390 | 610 | 8298 | — | — | — | |
Blend with PVK | 330 | 378; 589 | — | — | — | — | |
Blend with PVK:PBD | 330 | 378; 589 | — | — | — | — | |
4 | CHCl3 | 400 | 651 | 9639 | 69.53 | 1.198 | 0.89 |
CH3CN | 380 | 690 | 11![]() |
24.47 | 1.063 | 0.46 | |
Solid state | 442 | 591 | 5704 | 120.00 (11.30%), 459.99 (88.70%) | 1.013 | 7.32 | |
EtOH–MeOHa (77 K) | 360 | 567 | 10![]() |
1610 (14.95%), 5520 (75.89%), 17![]() |
1.097 | — | |
Film | 390 | 585 | 8351 | — | — | — | |
Blend with PVK | 330 | 384; 582 | — | — | — | — | |
Blend with PVK:PBD | 330 | 384; 577 | — | — | — | — | |
5 | CHCl3 | 397 | 634 | 9416 | 98.53 | 1.004 | 1.28 |
CH3CN | 376 | 650 | 11![]() |
44.81 | 1.123 | 0.48 | |
Solid state | 463 | 566 | 3930 | 525.33 (26.87%), 1259 (73.13%) | 1.104 | 11.38 | |
EtOH–MeOHa (77 K) | 362 | 548 | 9376 | 3080 (23.54%), 7750 (69.24%), 38![]() |
1.178 | — | |
Film | 390 | 561 | 8552 | — | — | — | |
Blend with PVK | 330 | 385; 575 | — | — | — | — | |
Blend with PVK:PBD | 330 | 383; 565 | — | — | — | — |
Complexes 1–5 were found to be typical MLCT emitters. The irradiation of 1–5 in solution with wavelengths of 374–408 nm gave rise to emissions with maxima and Stokes' shifts in the ranges 634–716 nm and 8885–11823 cm−1, respectively. The emission bands are broad and structureless, and they show negative solvatochromism, reflected in blue-shifted emission with increasing solvent polarity. On going from CH3CN to CHCl3, the emissions of 1, 2, 3, 4, and 5 appeared at 25, 25, 36, 39, and 16 nm shorter wavelengths, which is attributed to a reduced (and reversed) molecular dipole in their MLCT excited states.7a,19 As can be seen from Table 3, complexes 1–5 also exhibit shortened lifetimes and decreasing emission quantum yields upon changing the solvent from CHCl3 to CH3CN. In similarity to the absorption behavior, the emission maximum wavelengths of 3, 4, and 5 show red-shifts with an increasing number of chloro substituents, in the order 3 > 4 > 5. On the other hand, the emissions of 1 and 2 appear in a lower-energy region compared to 4, but there is no marked difference in the fluorescence maximum wavelengths of 1 and 2.
Another characteristic feature of these complexes (except for 1) is the large hypsochromic shift of their emission maxima on going from a fluid environment to a rigid medium, 566–643 nm in the solid state and 540–569 nm upon cooling to low temperature (77 K), described as the rigidochromic effect. This is responsible for raising the energy of the emissive 3MLLCT due to the lack of solvent reorganization following excitation.8a,20
For 1, bimodal emission was observed, at 594 and 716 nm. A significant increase in lifetimes in the solid state and in a 77 K EtOH–MeOH matrix can be attributed to the intermolecular interactions in a more rigid environment, and decreased thermal deactivation.
The relatively long emission lifetimes, together with large absorption–emission shifts, indicate the phosphorescent nature of the emission.
The normalized emission spectra in solution, solid state as powder, and in alcohol glass at 77 K for representative compound 2 are shown in Fig. 5.
![]() | ||
Fig. 5 Normalized emission spectra in solution, solid state, and in alcohol glass at 77 K for compound 2. |
The emission spectra of the remaining compounds are included in the ESI (Fig. S2†).
Additionally, PL spectra of the complexes as thin films on glass substrates were recorded under λex = 390 nm. It was found that the λem of films was shifted to a higher energetic region with respect to the chloroform solution, as for complexes in the form of powder. In film, they exhibited large Stokes shifts of up to 220 nm, as in solution. The large Stokes shifts are desirable for emitting materials in OLEDs. A small Stokes shift can cause self-quenching and consequently decrease the emission intensity.21
The emission properties of 1–5 were also examined computationally at the DFT/UB3LYP/def2-TZVPD level. The energies of phosphorescence emissions of 1–5 were calculated from the energy difference between the ground singlet and the triplet state ΔET1–S0, as well as being computed using TD-DFT (Table 4, Fig. S5 and Table S14†).
Compound | DFT | TD-DFT | |||||
---|---|---|---|---|---|---|---|
ΔET1–S0 | Character | Major contribution (CI coefficient) | E [eV] | λcal [nm] | Character | λexp [nm] [eV]−1 | |
1 | 1.97/629.4 | 3MLLCT | L → H (0.649) | 1.75 | 707.08 | 3MLLCT | 665/1.86 |
2 | 1.98/626.2 | 3MLLCT | L → H (0.663) | 1.87 | 661.55 | 3MLLCT | 665/1.86 |
3 | 1.90/652.5 | 3MLLCT | L → H (0.662) | 1.79 | 691.39 | 3MLLCT | 716/1.73 |
4 | 1.97/629.4 | 3MLLCT | L → H (0.666) | 1.86 | 665.16 | 3MLLCT | 690/1.79 |
5 | 2.09/593.2 | 3MLLCT | L → H (0.655) | 1.97 | 627.63 | 3MLLCT | 650/1.90 |
As can be seen from Table 4, the calculated emission energies are in reasonably good agreement with the experimental emission band maxima of 1–5, indicating that the luminescence has characteristics of phosphorescence and occurs from the low-lying triplet state (T1). The nature of the transitions resulting in the emissions of 1–5 is 3MLLCT. Compared to the experimental data, the excitation energies calculated by the TD-DFT method lie closer than those of ΔET1–S0 for complexes 2–5, whereas ΔET1–S0 is more accurate than the excitation energies calculated by TDDFT for 1.
The next step of our research covers a study of the possible applications of the complexes as guest components in guest–host light-emitting diodes prepared by solution processing. Therefore, the photoluminescence ability of these complexes in the guest–host configuration as blends with poly(9-vinylcarbazole) (PVK) and in mixtures of PVK with (2-(4-tert-butylphenyl)-5-(4-biphenylyl)-1,3,4-oxadiazole) (PBD) (50:
50 as weight%) were investigated. The content of the complexes in blends was 15 wt%. The layers of such compositions were used for light-emitting diodes (see the section on Electroluminescence). The emission spectra of the blends with complexes were recorded under excitation at 330 nm and 390 nm, corresponding to the λmax values of the matrices and the complexes, respectively. Under excitation at 330 nm, both matrices, that is, PVK and the mixture of PVK:PBD, exhibited the highest intensity emission (cf. Fig. S3 in ESI†). All blends under λex = 390 nm did not show emission, except of blends with complex 5. In Fig. 6, the representative photoluminescence spectra of the matrices (PVK and PVK:PBD) and blend with dispersed compound 5 are compared.
In the emission spectra of all complexes dispersed in matrices under excitation at 330 nm, two emission bands were observed. A band originating from the emission of both matrices was seen in a higher energy region around 400 nm, in relation to the band ascribed to the luminescence of the complexes from about 570–600 nm. In the guest–host diode configuration, two luminescence mechanisms are possible, that is, Förster energy transfer and charge trapping.22 The obtained results suggest that in PVK and PVK:PBD blends, Förster energy transfer from host to guest occurs, which is possible only if the emission spectrum of the matrix–host overlaps with the absorption spectrum of the guest molecule. As can be seen in Fig. 6b, the emission spectra of PVK and PVK:PBD overlap over a large range of the absorption spectra of the complexes. However, the emission of the matrices is still present in the PL spectra of the blends, indicating that the energy transfer is not complete. A higher intense emission was observed for complexes dispersed in a binary host matrix (PVK:PBD).
In these matrices, the complexes were dispersed with 15 wt% concentration. At the beginning of our research, devices with an emitting layer as a neat complex film with structure ITO/PEDOT:PSS:Re(I) complex/Al were prepared. Next the EL of the synthesized compounds was tested in devices with the configuration ITO/PEDOT:PSS/PVK:Re(I) complex/Al and ITO/PEDOT:PSS/PVK:PBD:Re(I) complex/Al.
For the prepared diodes, current density–voltage (J–V) characteristics were registered. Fig. 7 presents an example of the J–V characteristics, together with photos of diodes under an applied voltage, whereas the data obtained for all devices are summarized in Table 5. The thicknesses of the selected active layer and active layer surface were investigated by atomic force microscopy (AFM). The layers were characterized by similar thicknesses around 100–140 nm. The AFM images of the emitting layers show a uniform and flat surface with a surface root-mean-square roughness (RMS) in the range 1.8–4.9 nm.
![]() | ||
Fig. 7 Current density–voltage (J–V) characteristics of the prepared devices based on complexes 3 and 5, and photos of diodes under an applied voltage. |
Compound | Active layer | Von [V] | Jmax [mA cm−2] |
---|---|---|---|
1 | Film | 3.0 | 0.6 |
PVK blend | 2.8 | 18 | |
PVK:PBD blend | 2.6 | 19 | |
2 | Film | — | — |
PVK blend | 2.6 | 16 | |
PVK:PBD blend | 2.6 | 40 | |
3 | Film | 3.5 | 2.6 |
PVK blend | 2.1 | 35 | |
PVK:PBD blend | 2.6 | 76 | |
4 | Film | 2.0 | 6 |
PVK blend | 1.7 | 14 | |
PVK:PBD blend | 1.2 | 17 | |
5 | Film | — | — |
PVK blend | 2.2 | 65 | |
PVK:PBD blend | 1.6 | 46 |
In the case of diodes with the neat complex as emission layer, a low current density (J below 5 mA cm−2) was obtained and no light emission under the applied voltage was seen. The devices with an active layer consisting of complexes dispersed in matrices were characterized by low values of the turn-on voltage (Von) in the range 1.2–3.0 V, indicating efficient injection from the electrodes and transport of holes and electrons to the emission layers. The maximal achieved current density (Jmax) was in the range 14–76 mA cm−2. The highest Jmax values were found for devices based on compounds 3 and 5, used as emitters. The utilization of an additional electron-transporting compound (PBD) as a component of the emitting layer resulted in an improvement in the device performance. It should be emphasized that all fabricated diodes containing complexes in the blend emitted light under the applied voltage, which was more intense in the case of devices with a binary matrix. Thus, these complexes seem to be promising for applications, and further investigations covering measurements of electroluminescence parameters, together with diode architecture modification, are necessary and will be carried out.
Poly(9-vinylcarbazole) (PVK) (Mn = 25000–50
000) and 2-(4-tert-butylphenyl)-5-(4-biphenylyl)-1,3,4-oxadiazole (PBD) were purchased from Sigma Aldrich and used without additional purification. Poly(3,4-(ethylenedioxy)thiophene):poly-(styrenesulfonate) (PEDOT:PSS) (0.1–1.0 S cm−1) and substrates with pixelated ITO anodes were supplied by Ossila. All solvents for synthesis were of reagent grade and were used as received. For spectroscopy studies, HPLC grade solvents were used.
Footnote |
† Electronic supplementary information (ESI) available: Crystal data and structure refinement, short intra- and intermolecular contacts, comparison of experimental and theoretical bond lengths [Å] and angles [°], the energies and characters of the selected eletronic transitions with their assignment to the experimental absorption bands, percentage contributions of Re, CO, phen and Cl units in the selected occupied and unoccupied molecular orbitals in the ground state, 1H and 13C NMR spectra, DSC thermograms, excitation and emission spectra, emission spectra of PVK and PVK![]() ![]() ![]() ![]() |
This journal is © The Royal Society of Chemistry 2016 |