Next Article in Journal
Evaluation of Organophosphate Pesticide Residues in Food Using the Partial Least Squares Method
Previous Article in Journal
1′-Homocarbocyclic Nucleoside Analogs with an Optically Active Substituted Bicyclo[2.2.1]Heptane Scaffold
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Proceeding Paper

Anionic Low Band Gap-Conjugated Polyelectrolytes as Hole-Transporting Layer in Optoelectronics Devices †

by
Benedetta Maria Squeo
*,
Wojciech Mróz
,
Umberto Giovanella
and
Mariacecilia Pasini
*
Istituto di Scienze e Tecnologie Chimiche (SCITEC), Consiglio Nazionale delle Ricerche, via Corti 12, 20133 Milano, Italy
*
Authors to whom correspondence should be addressed.
Presented at the 24th International Electronic Conference on Synthetic Organic Chemistry, 15 November–15 December 2020; Available online: https://ecsoc-24.sciforum.net/.
Chem. Proc. 2021, 3(1), 18; https://doi.org/10.3390/ecsoc-24-08406
Published: 14 November 2020

Abstract

:
In the last years, interfacial engineering has played a critical role in promoting the performance of optoelectronic devices as organic solar cells (OSC) and organic light-emitting diodes (OLEDs) since interfacial layers help to form an ohmic contact between the electrodes and the active layers, which is of great importance for charge collection/injection. Conjugated polyelectrolytes (CPEs), which are conjugated polymers bearing side-chain ionic functionalities such as anionic, cationic, or zwitterionic groups, have emerged as a new class of interfacial materials in thin film-based electronic devices thanks to their ability to reduce the barrier between electrode and active layer. In view of this, we designed and synthesized two novel low bandgap anionic copolymers with different anionic pendant groups and different conjugated backbones to obtain hole-transporting layer (HTL) materials as an alternative to commonly used poly(3,4-ethylenedioxythiophene):poly(styrenesulfonate). The functional behavior of these copolymers as anode modifiers is herein preliminarily investigated in an OLED prototype.

1. Introduction

The control of the interfaces has proved to be one of the crucial elements for the engineering of different types of optoelectronic devices such as OLED (organic light-emitting diodes), OSC (organic solar cells), and OFET (field effect transistor), which are based on organic and inorganic multilayer systems [1,2].
There are different strategies for the engineering of the interfaces, but certainly conjugated polyelectrolytes [3] are playing an increasingly important role in organic [4] and hybrid systems, including colloidal semiconductor nanocrystals [5,6] and perovskites [7,8].
The conjugated polyelectrolytes are formed by an organic semiconductor backbone and by polar side pendants that can be anionic, cationic, or neutral.
Similarly, in traditional semiconductor polymers, it is possible to modify the main conjugated chain and consequently modulate the energy bandgap of the system. The presence of polar groups on the side chains, on the other hand, ensures solubility in polar solvents such as water.
The conjugated polyelectrolytes (CPEs) have proved to be particularly effective in solid-state optoelectronic applications. In fact, as solution-processed materials with respect to the evaporated counterpart, they are ideal for low-cost, large-scale technologies such as roll-to-roll or ink-jet printing; their orthogonal solubility simplifies the preparation procedures of the devices, preventing re-dissolution of the hydrophobic semiconducting layer during device fabrication; they are compatible with flexible devices; they are environmental friendly thanks to water solubility; they have unique solid-state interface properties, which increase the adhesion between organic and metallic layers; and they have the ability to reduce the barrier between electrode and active layers thanks to the formation of an aligned interfacial dipole at the metal/organic semiconductor interface.
There are mainly two kinds of interfacial materials that are commonly used in optoelectronic devices—if they are at the interface with the anode, they are called the anode interfacial layer, or AIL, and cathode interfacial layer, or CIL, when they are at the interface with the cathode (Figure 1).
Considering a traditional direct geometry device, several CILs have been developed, which generally have a fluorenic, phenylene, or benzodithiazole-based backbone with cationic ammonium or polar substituent groups such as amine [9,10] or phosphonates [11], while the development of AILs is lagging behind.
The most commonly used AIL material in OSCs is acidic poly(3,4-ethylenedioxythiophene):poly(styrenesulfonate) (PEDOT:PSS), thanks to its high optical transparency in the visible NIR (Near InfraRed) regime, suitable energy level, and sufficient conductivity. However its hygroscopic and acidic nature induces the etching of the indium from the indium tin oxide (ITO)/PEDOT:PSS interface, then causing the diffusion of indium and ultimately leading to fast deterioration of the devices [12,13].
Recently, Bazan and coworkers [14] demonstrated an anionic low bandgap CPE copolymer based on the acceptor (A) benzodithiazole (BT) and on the donor (D) cyclopentadithiophene (CPDT) bearing sulfonated side groups. This copolymer can be easily doped after dialysis in water and, thanks to its pH-neutral nature, is a promising alternative to acidic PEDOT:PSS.
Although this material has been successfully incorporated in different types of devices [5,15,16,17], from a chemical point of view, it is important to understand the role of the backbone and of the lateral anionic substituents.
Herein, we present the synthesis and preliminary optoelectronic characterization of two new low bandgap D-A anionic CPEs as a possible anternative to PEDOT:PSS. We copolymerized the BT monomer with a dithienopyrrolo (DTP) unit that has been reported to be a stronger donor with respect to CPDT moiety. DTP has never been used, to our knowledge, in CPE, and for comparison, we synthetized a CPDT–BT polymer with only a lateral substituent. The DTP–BT polymer has a carboxylate lateral group with a lower acidity than the sulfonate group but with a stronger donor [18] in the backbone (polymer P1, Figure 2), and the CPDT–BT polymer is similar to the CPE proposed by Bazan et al., but bearing only one sulfonate group. We aimed to understand the combined role of the backbone, in particular the energy bandgap modulation, as well as the lateral polar groups in the self-doping process of these materials
The UV–VIS–NIR optical characterization combined with the electron paramagnetic resonance (EPR) showed that, regardless of the bandgap and the strength of the donors, the sulfonate group seemed to have a fundamental role in the self-doping process of the polymer.
Finally, P2 was also tested in a very preliminary OLED prototype.

2. Materials and Methods

All glassware was oven-dried. Unless specifically mentioned, all chemicals were commercially available and were used as received.
Ethyl 6-(4H-dithieno[3,2-b:2’3’-d]pyrrole-4-yl)hexanoate (1) was synthesized following the procedure reported by Zecchin et al. [19].
The dialysis membrane (MWCO: 3500–5000 Da) was purchased from Membrane Filtration Products Inc. NMR spectra were recorded at 600 MHz.
Dibromo-ethyl 6-(4H-dithieno[3,2-b:2’3’-d]pyrrole-4-yl)hexanoate (2). Compound 1 (150 mg, 0.511 mmol, 1 eq.) was suspended in dimethylformamide (DMF) (1 mL) and cooled down at 0 °C. A solution of N-bromosuccinimide (NBS; 182 mg2, 1.022 mmol, 2 eq.) in chloroform (CHCl3, 5.11 mL) was added dropwise under stirring. The reaction was allowed to reach room temperature in the dark by shielding the flask with aluminum foil. After 1 h at room temperature, the crude product was diluted with CHCl3 and washed 3 times with water. The organic solvent was removed under reduced pressure and the product was used without further purification. 1H NMR (D2O) δ: 7.00 (s, 2H), 4.05 (t, 2H), 2.30 (m, 2H), 1.85 (m, 2H), 1.65 (m, 2H), 1.35 (m, 2H).
Polymer P1. A mixture of compound 2 (230.5 mg, 0.511 mol, 1 eq.), 2,1,3-benzothiadiazole-4,7-bis(boronic acid pinacol ester) (198.32 mg, 0.511 mmol, 1 eq.), and tetrakis(triphenylphosphine)palladium(0) (Pd(PPh3)4, 11.8 mg, 2% mol) was added in a pre-degassed Schlenk flask, followed by 3 vacuum/nitrogen cycles. Then, degassed DMF (4.8 mL) and degassed potassium carbonate aqueous solution (1.2 mL) were added. The mixture was stirred at 110 °C for 3 h. The reaction mixture was then poured in acetone and the dark blue precipitate was collected by filtration and washed with copious amounts of acetone. The precipitate was all dissolved in deionized H2O and transferred into a dialysis tube (MWCO: 3500–5000). The dialysis tube was placed in a large beaker with H2O stirring for 3 days, and the H2O was changed every 12 h. Evaporation of H2O provided the title product as a dark blue solid (109 mg, 50%) after drying under vacuum overnight.
Synthesis of 4-potassium butanylsulfonate-4H-cyclopenta-[2,1-b;3,4-b’]-dithiophene (3). 4H-cyclopenta-[2,1-b;3,4-b’]-dithiophene (CPDT, 300 mg, 1.68 mmol, 1.0 equiv) and tetrabutylammonium bromide (27 mg, 0.084 mmol, 0.05 equiv) were dissolved in anhydrous DMSO (8.2 mL), and the solution was degassed by bubbling with Ar for 5 min. 50% KOH in H2O (1.8 g) was added via syringe, followed by the addition of 1,4-butanesultone (924 µL, 9.02 mmol, 1.2 equiv). After stirring at room temperature for 3 h, the reaction mixture was poured into acetone (50 mL), and the yellowish precipitate was collected by filtration and washed with acetone. The crude was used in the next step without further purification. 1H NMR (D2O) δ: 7.05 (d, 2H), 6.95 (d, 2H), 3.42 (t, 1H) 2.73 (t, 2H), 1.63 (m, broad, 2H), 1.56 (m, broad, 2H), 1.3 (m, broad, 2H).
Synthesis of 2,6-dibromo-4-potassium butanylsulfonate-4H-cyclopenta-[2,1-b;3,4-b’]-dithiophene (4). The crude product 3 was suspended in DMF (6.7 mL), and H2O (≈1 mL) was added while stirring until it was all dissolved. NBS (747 g, 4.2 mmol, 2.5 equiv) was added in the dark by shielding the flask with aluminum foil. The brown solution was stirred at room temperature for 1 h and poured into acetone. The yellowish precipitate was collected by filtration and washed with acetone (640 mg, 75% yield). 1H NMR (D2O) δ 7.23 (s, 2H), 3.74 (t, broad, 1H), 2.80 (m, broad, 2H), 2.60–2.50 (m, broad, 2H), 1.85–1.80 (m, broad, 2H), 1.55–1.45 (m, broad, 2H).
Synthesis of polymer P2. A mixture of compound 4 (117 mg, 0.230 mol, 1 eq.), 2,1,3-benzothiadiazole-4,7-bis(boronic acid pinacol ester) (89 mg, 0.230 mmol, 1 eq.), and Pd(PPh3)4 (5.2 mg, 2% mol) were added in a pre-degassed Schlenk flask, followed by 3 vacuum/nitrogen cycles. Then, degassed DMF (2.2 mL) and degassed potassium carbonate aqueous solution (0.55 mL) were added. The mixture was stirred at 110 °C for 3 h. The reaction mixture was then poured in acetone and the dark blue precipitate was collected by filtration and washed with copious amounts of acetone. The precipitate was all dissolved in deionized H2O and transferred into a dialysis tube (MWCO: 3500–5000). The dialysis tube was placed in a large beaker with H2O stirring for 3 days, and the H2O was changed every 12 h. Evaporation of H2O provided the title product as a dark blue solid (65 mg, 60%) after drying under vacuum overnight.
OLED fabrication and characterization. Glass substrates covered with indium tin oxide (ITO) pattern were washed ultrasonically in distilled water, acetone, and isopropyl alcohol. On the substrates, a water solution (2 mg/mL) of P2 was spincoated using 1400 rpm rotation, and subsequently the substrates were annealed at 100 °C for 10 minutes inside nitrogen-filled glovebox. Emissive layer was prepared from poly(9,9-dioctylfluorene-alt-benzothiadiazole) (F8BT, American Dye Source) dissolved in toluene at a concentration of 10 mg/mL and deposited with a rotation of 1000 rpm. Then, the devices were transferred to an evaporator where 7 nm of barium and 100 nm of aluminum were deposited through a shadow mask at 10−6 mbar pressure.
Electroluminescence (EL) spectra were recorded with liquid nitrogen-cooled charge-coupled device (CCD) combined with a monochromator (Spex 270M) and applying a constant bias. Keithley 2602 source meter was employed during current density–voltage–luminance measurements. Light emitted in forward direction from the devices was collected by a photodiode with known spectral response. In efficiency calculations, Lambert distribution was assumed. All characterization was performed in a nitrogen atmosphere.

3. Results and Discussion

Chemical characterization. The syntheses of compounds 1 and 2 are shown in Figure 2 and confirmed by 1H NMR spectra (Figure 3). In particular, the monosubstitution of monomer 2 was confirmed by the presence of a triplet at 3.74, attributable to H in 4 position in cylopenta-[2,1-b:3,4-b’]dithiophene [20].
Polymer P1 and P2 were synthesized through Suzuki cross-coupling of a dibromide- of the donor unit and the bis-boronic ester of benzothiadiazole, with Pd(PPh3)4 as a catalyst and K2CO3 as a base in degassed DMF as solvent. The polymers were purified via dialysis membrane with a cutoff of 3500–5000 Da.
Differently from what was reported earlier [15], where the pendant anionic moieties ensured CPEs solubility in polar organic solvents (e.g., ethanol) or water, both P1 and P2 polymers showed poor solubility both in traditional organic solvents and in polar solvents, probably due to the amphiphilic character given by the combination of the apolar backbone and the polar side chains.
The NMR spectra of the polymers showed only non-informative broad peaks, probably due to their paramagnetic characteristic, but the formation of the polymers was assured by the strong change of the color of the solution from colorless to blue/violet.
Cyclic voltammetry (Figure 4) was applied to investigate the electrochemical properties. The HOMO levels of the polymer P1 and P2 (−4.8 and −4.92 eV, respectively) were in agreement with that reported by Bazan et al. [14] for a similar polymer and were comparable to that of PEDOT:PSS (≈−5 eV) reported in the literature.
Polymer P2 showed an interesting behavior—the first oxidation peak was at high potential, while the second and third, perfectly reproducible, were at lower potential. Most probably this behavior indicated the formation of solid-state aggregates that disaggregate after the first potential application [21].
Self-doping behavior. In order to understand the self-doping properties of the polymers, we recorded absorption spectra in water solution at different pH values. In fact, it has been shown for CPDT-based CPE that the acidic environment favors the doping of the polymers and that the formation of a polaron contributes in the absorption spectrum.
In UV–VIS–NIR absorption characterization (Figure 5), the peak at 410 nm corresponded to p–p* absorption from localized p orbitals, while the higher intensity peak around 700 nm corresponded to the absorption from the intramolecular charge transfer (ICT) band [14].
As can be observed in Figure 5, the absorption spectrum of polymer P1 seemed to be unaffected by the addition of an acid or base to change the pH, pointing out that the polymer cannot be doped in acidic conditions. On the other hand, polymer P2 showed the uprising of a polaron band in the near-infrared region with the increasing of the acidity of the solution, confirming the self-doping behavior of the material [14].
Further evidence of self-doping behavior was obtained by electron paramagnetic resonance (EPR) measurements in aqueous solution (Figure 6)—strong EPR symmetric signal consistent with the presence of unpaired electrons was observed for both PEDOT:PSS and P2, while no signal was observed for polymer P1, in agreement with the absorption spectra.
Since both polymers had a low and very similar energy gap, the explanation of the different behavior was probably to be found in the different polar groups. The sulfonate group appeared to have a crucial impact in the formation of self-doped systems.
Preliminary OLED device. The polymer P2, showing self-doping behavior, was tested as a hole-transporting layer (HTL) in a PEDOT:PSS-free OLED architecture with F8BT as emitter. In the left part of Figure 4, the EL spectrum of the diode is shown together with Commission Internationale de l’Eclairage (CIE) chromaticity coordinates (0.41, 0.58). The EL spectrum is typical for F8BT, and no contribution from P2 polymer was observed.
The performance of the device is presented in the right panel of Figure 7. The J(U) and L(U) curves had forms typically observed for diodes. The device switched on at 6 V and the maximal recorded luminance approached 40 cd/m2. Current density values were quite high and in combination with modest luminance led to a weak external quantum efficiency (EQE), shown in the inset of the right panel of Figure 7. It should be underlined that performance was achieved for the device without any optimization.

4. Conclusions

Two new low bandgap-conjugated polyelectrolytes, polymer P1 and P2, were synthesized and characterized.
The DTP moiety, reported to be a stronger donor with respect to CPDT monomers, was used for the first time in the synthesis of a low bandgap CPE, giving a material with suitable HOMO–LUMO levels. Polymer P1 did not show a self-doping behavior, as confirmed by EPR measurements. On the contrary, polymer P2 seemed to be a promising material because it showed a self-doping behavior, hence highlighting the importance of the proper combination between low bandgap-conjugated backbone and suitable anionic substituent.
A very preliminary OLED device using polymer P2 as HTL was fabricated. The performance, even if modest due to the non-optimized architecture and parameters, showed the potential application of P2 compound.
This work will help guide the synthetic design of new, highly conductive CPEs. Specifically, we showed the importance of incorporating an electron-rich structural unit into the polymer backbone, which can be oxidized easily, as well as the necessity of having a pendant anionic group to stabilize the resulting positively charged backbone for self-doped polymer systems.
However, a more comprehensive understanding of the importance of sulfonate for the self-doping mechanism of CPE and its effect on charge transport and mobility is required to improve the design and the use of this class of polymers.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article.

Acknowledgments

The authors thanks Barbara Vercelli for cyclic voltammetry measurements and Maristella Gussoni and Simona Mrakic-Sposta for the EPR measurements.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fahlman, M.; Fabiano, S.; Gueskine, V.; Simon, D.; Berggren, M.; Crispin, X. Interfaces in organic electronics. Nat. Rev. Mater. 2019, 4, 627–650. [Google Scholar] [CrossRef]
  2. Steim, R.; Kogler, F.R.; Brabec, C.J. Interface materials for organic solar cells. J. Mater. Chem. 2010, 20, 2499–2512. [Google Scholar] [CrossRef]
  3. Chueh, C.-C.; Li, C.-Z.; Jen, A.K.-Y. Recent progress and perspective in solution-processed Interfacial materials for efficient and stable polymer and organometal perovskite solar cells. Energy Environ. Sci. 2015, 8, 1160–1189. [Google Scholar] [CrossRef]
  4. Carulli, F.; Scavia, G.; Lassi, E.; Pasini, M.; Galeotti, F.; Brovelli, S.; Giovanella, U.; Luzzati, S. A bifunctional conjugated polyelectrolyte for the interfacial engineering of polymer solar cells. J. Colloid Interface Sci. 2019, 538, 611–619. [Google Scholar] [CrossRef]
  5. Lagonegro, P.; Martella, C.; Squeo, B.M.; Carulli, F.; Scavia, G.; Lamperti, A.; Galeotti, F.; Dubertret, B.; Pasini, M.; Brovelli, S.; et al. Prolonged Lifetime in Nanocrystal Light-Emitting Diodes Incorporating MoS2-Based Conjugated Polyelectrolyte Interfacial Layer as an Alternative to PEDOT:PSS. ACS Appl. Electron. Mater. 2020, 2, 1186–1192. [Google Scholar] [CrossRef]
  6. Castelli, A.; Meinardi, F.; Pasini, M.; Galeotti, F.; Pinchetti, V.; Lorenzon, M.; Manna, L.; Moreels, I.; Giovanella, U.; Brovelli, S. High-Efficiency All-Solution-Processed Light-Emitting Diodes Based on Anisotropic Colloidal Heterostructures with Polar Polymer Injecting Layers. Nano Lett. 2015, 15, 5455–5464. [Google Scholar] [CrossRef] [PubMed]
  7. Choi, H.; Mai, C.-K.; Kim, H.-B.; Jeong, J.; Song, S.; Bazan, G.C.; Kim, J.Y.; Heeger, A.J. Conjugated polyelectrolyte hole transport layer for inverted-type perovskite solar cells. Nat. Commun. 2015, 6, 7348. [Google Scholar] [CrossRef] [PubMed]
  8. Zhang, L.; Zhou, X.; Xie, J.; Chen, S.; Bae, S.; Kim, J.; Xu, B. Conjugated polyelectrolyte with potassium cations enables inverted perovskite solar cells with an efficiency over 20%. J. Mater. Chem. A 2020, 8, 8238–8243. [Google Scholar] [CrossRef]
  9. Carulli, F.; Mróz, W.; Lassi, E.; Sandionigi, C.; Squeo, B.; Meazza, L.; Scavia, G.; Luzzati, S.; Pasini, M.; Giovanella, U.; et al. Effect of the introduction of an alcohol-soluble conjugated polyelectrolyte as cathode interlayer in solution-processed organic light-emitting diodes and photovoltaic devices. Chem. Pap. 2018, 72, 1753–1759. [Google Scholar] [CrossRef]
  10. Prosa, M.; Benvenuti, E.; Pasini, M.; Giovanella, U.; Bolognesi, M.; Meazza, L.; Galeotti, F.; Muccini, M.; Toffanin, S. Organic Light-Emitting Transistors with Simultaneous Enhancement of Optical Power and External Quantum Efficiency via Conjugated Polar Polymer Interlayers. ACS Appl. Mater. Interfaces 2018, 10, 25580–25588. [Google Scholar] [CrossRef] [PubMed]
  11. Squeo, B.M.; Carulli, F.; Lassi, E.; Galeotti, F.; Giovanella, U.; Luzzati, S.; Pasini, M. Benzothiadiazole-based conjugated polyelectrolytes for interfacial engineering in optoelectronic devices. Pure Appl. Chem. 2019, 91, 477–488. [Google Scholar] [CrossRef]
  12. Zhou, H.; Zhang, Y.; Mai, C.-K.; Seifter, J.; Nguyen, T.-Q.; Bazan, G.C.; Heeger, A.J. Solution-Processed pH-Neutral Conjugated Polyelectrolyte Improves Interfacial Contact in Organic Solar Cells. ACS Nano 2015, 9, 371–377. [Google Scholar] [CrossRef] [PubMed]
  13. Cui, Y.; Xu, B.; Yang, B.; Yao, H.; Li, S.; Hou, J. A Novel pH Neutral Self-Doped Polymer for Anode Interfacial Layer in Efficient Polymer Solar Cells. Macromolecules 2016, 49, 8126–8133. [Google Scholar] [CrossRef]
  14. Mai, C.-K.; Zhou, H.; Zhang, Y.; Henson, Z.B.; Nguyen, T.-Q.; Heeger, A.J.; Bazan, G.C. Facile Doping of Anionic Narrow-Band-Gap Conjugated Polyelectrolytes During Dialysis. Angew. Chem. Int. Ed. 2013, 52, 12874–12878. [Google Scholar] [CrossRef]
  15. Lee, B.R.; Yu, J.C.; Park, J.H.; Lee, S.; Mai, C.-K.; Zhao, B.; Wong, M.S.; Jung, E.D.; Nam, Y.S.; Park, S.Y.; et al. Conjugated Polyelectrolytes as Efficient Hole Transport Layers in Perovskite Light-Emitting Diodes. ACS Nano 2018, 12, 5826–5833. [Google Scholar] [CrossRef]
  16. Zhou, H.; Zhang, Y.; Mai, C.-K.; Collins, S.D.; Bazan, G.C.; Nguyen, T.-Q.; Heeger, A.J. Polymer Homo-Tandem Solar Cells with Best Efficiency of 11.3%. Adv. Mater. 2015, 27, 1767–1773. [Google Scholar] [CrossRef]
  17. Zhou, H.; Zhang, Y.; Mai, C.-K.; Collins, S.D.; Nguyen, T.-Q.; Bazan, G.C.; Heeger, A.J. Conductive Conjugated Polyelectrolyte as Hole-Transporting Layer for Organic Bulk Heterojunction Solar Cells. Adv. Mater. 2014, 26, 780–785. [Google Scholar] [CrossRef]
  18. Popere, B.C.; Pelle, A.M.D.; Poe, A.; Balaji, G.; Thayumanavan, S. Predictably tuning the frontier molecular orbital energy levels of panchromatic low band gap BODIPY-based conjugated polymers. Chem. Sci. 2012, 3, 3093–3102. [Google Scholar] [CrossRef]
  19. Berlin, A.; Zotti, G.; Schiavon, G.; Zecchin, S. Adsorption of Carboxyl-Terminated Dithiophene and Terthiophene Molecules on ITO Electrodes and Their Electrochemical Coupling to Polymer Layers. The Influence of Molecular Geometry. J. Am. Chem. Soc. 1998, 120, 13453–13460. [Google Scholar] [CrossRef]
  20. Pasini, M.; Destri, S.; Botta, C.; Porzio, W. Synthesis, optical and electrochemical characterization of lnter-ring bridged tetramers based on thiophene. Tetrahedron 1999, 55, 14985–14994. [Google Scholar] [CrossRef]
  21. Pasini, M.; Vercelli, B.; Zotti, G.; Berlin, A. Solid-state Effects in the Cyclovoltammetric HOMO-LUMO Determination: The case of Dinitrophenyl-hydrazone α,Ͽ-substituted Oligothiophenes. Electrochim. Acta 2016, 193, 261–267. [Google Scholar] [CrossRef]
Figure 1. General structure of conjugated polyelectrolytes and their main characteristics.
Figure 1. General structure of conjugated polyelectrolytes and their main characteristics.
Chemproc 03 00018 g001
Figure 2. Synthesis of polymers P1 and P2.
Figure 2. Synthesis of polymers P1 and P2.
Chemproc 03 00018 g002
Figure 3. 1H-NMR spectra of compounds 2 and 4.
Figure 3. 1H-NMR spectra of compounds 2 and 4.
Chemproc 03 00018 g003
Figure 4. Voltammograms of polymers P1 and P2.
Figure 4. Voltammograms of polymers P1 and P2.
Chemproc 03 00018 g004
Figure 5. Absorption spectra in water solution of polymers P1 and P2.
Figure 5. Absorption spectra in water solution of polymers P1 and P2.
Chemproc 03 00018 g005
Figure 6. Electron paramagnetic resonance (EPR) spectra of poly(3,4-ethylenedioxythiophene):poly(styrenesulfonate) (PEDOT:PSS), P1 and P2.
Figure 6. Electron paramagnetic resonance (EPR) spectra of poly(3,4-ethylenedioxythiophene):poly(styrenesulfonate) (PEDOT:PSS), P1 and P2.
Chemproc 03 00018 g006
Figure 7. (Left) EL spectrum of poly(9,9-dioctylfluorene-alt-benzothiadiazole) (F8BT)-based diode with P2 as a hole-transporting layer (HTL), registered at 10 V. In the inset, Commission Internationale de l’Eclairage (CIE) coordinates of the spectrum are shown. (Right) Current density (J)–voltage (U)–luminance (L) characteristics of the organic light-emitting diode (OLED). In the inset, external quantum efficiency (EQE) is presented as a function of J.
Figure 7. (Left) EL spectrum of poly(9,9-dioctylfluorene-alt-benzothiadiazole) (F8BT)-based diode with P2 as a hole-transporting layer (HTL), registered at 10 V. In the inset, Commission Internationale de l’Eclairage (CIE) coordinates of the spectrum are shown. (Right) Current density (J)–voltage (U)–luminance (L) characteristics of the organic light-emitting diode (OLED). In the inset, external quantum efficiency (EQE) is presented as a function of J.
Chemproc 03 00018 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Squeo, B.M.; Mróz, W.; Giovanella, U.; Pasini, M. Anionic Low Band Gap-Conjugated Polyelectrolytes as Hole-Transporting Layer in Optoelectronics Devices. Chem. Proc. 2021, 3, 18. https://doi.org/10.3390/ecsoc-24-08406

AMA Style

Squeo BM, Mróz W, Giovanella U, Pasini M. Anionic Low Band Gap-Conjugated Polyelectrolytes as Hole-Transporting Layer in Optoelectronics Devices. Chemistry Proceedings. 2021; 3(1):18. https://doi.org/10.3390/ecsoc-24-08406

Chicago/Turabian Style

Squeo, Benedetta Maria, Wojciech Mróz, Umberto Giovanella, and Mariacecilia Pasini. 2021. "Anionic Low Band Gap-Conjugated Polyelectrolytes as Hole-Transporting Layer in Optoelectronics Devices" Chemistry Proceedings 3, no. 1: 18. https://doi.org/10.3390/ecsoc-24-08406

Article Metrics

Back to TopTop