Next Article in Journal
Investigating the Crystallization Process of Boron-Bearing Silicate-Phosphate Glasses by Thermal and Spectroscopic Methods
Next Article in Special Issue
Influence of Hexagonal Boron Nitride on Electronic Structure of Graphene
Previous Article in Journal
Expression, Purification, and Comparative Inhibition of Helicobacter pylori Urease by Regio-Selectively Alkylated Benzimidazole 2-Thione Derivatives
Previous Article in Special Issue
Mineralized Polyvinyl Alcohol/Sodium Alginate Hydrogels Incorporating Cellulose Nanofibrils for Bone and Wound Healing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis, Characterization, and Thin-Film Transistor Response of Benzo[i]pentahelicene-3,6-dione

1
Department of Chemical Engineering Materials and Environment, University of Rome “Sapienza”, Via Eudossiana 18, 00184 Rome, Italy
2
Department of Chemical and Biomolecular Engineering, Sogang University, Seoul 04107, Korea
3
Department of Chemistry, Biology and Biotechnology, University of Perugia, Via Elce di Sotto 8, 06123 Perugia, Italy
*
Authors to whom correspondence should be addressed.
Submission received: 30 December 2021 / Revised: 24 January 2022 / Accepted: 25 January 2022 / Published: 27 January 2022

Abstract

:
Organic semiconductors hold the promise of simple, large area solution deposition, low thermal budgets as well as compatibility with flexible substrates, thus emerging as viable alternatives for cost-effective (opto)-electronic devices. In this study, we report the optimized synthesis and characterization of a helically shaped polycyclic aromatic compound, namely benzo[i]pentahelicene-3,6-dione, and explored its use in the fabrication of organic field effect transistors. In addition, we investigated its thermal, optical absorption, and electrochemical properties. Finally, the single crystal X-ray characterization is reported.

1. Introduction

The field of organic electronics has been pushed strongly forward over recent decades, due to the synergy between different disciplines, including materials science, physics, chemistry, and engineering, as well as the interplay between academic and industrial actors [1,2,3,4,5].
From a synthetic chemistry point of view, significant efforts have been made to develop organic semiconducting materials, which could offer a viable alternative to their traditional inorganic counterparts. A wide range of small molecules and polymers have been successfully designed and synthesized for applications in many different devices (e.g., organic thin-film transistors (OTFTs), electrooptic modulators, organic (OPVs) or hybrid solar cells, organic light-emitting diodes (OLEDs), laser diodes, optical waveguides, and sensors), including both p-type and n-type semiconductors [6,7,8]. The structural diversity in terms of the core type, side chains, substituents, bridging units, conjugation extent, and shape of organic semiconductors [9,10,11,12,13,14] enables the fine-tuning of the material’s properties depending on the specific device requirements.
Among the molecular materials, the possibility to create devices from organic semiconductors with a helically shaped conjugated electronic system, namely helicenes, has been gaining attention recently.
Helicenes [15,16,17,18] are inherently dissymmetric structures, made up of ortho-fused aromatic or heteroaromatic rings, in which the repulsive steric overlap of the (terminal) aromatic units and, eventually, the presence of functional groups in them determines the non-planarity of the conjugated structure, imparting a helical twist. Hence, helicenes exist in both left-handed (M, minus) and right-handed (P, plus) chiral (i.e., enantiomeric) forms.
This helically chiral, fully conjugated architecture endows these molecules with unique chiroptical properties, e.g., high optical rotatory power, strong circularly polarized luminescence, and circular dichroism [19,20,21].
It has been demonstrated that helicenes may be able to self-assemble into long corkscrew-shaped columns, which subsequently aggregate to form large crystalline domains in both concentrated solutions and thin films [22,23,24,25]. Besides dramatically enhancing the helicene chiroptical properties, this supramolecular aggregation may provide unique charge transport properties, as observed for, e.g., DNA [26].
Despite their peculiar properties, however, helicenes still remain rather unexplored in solid-state (opto-)electronic devices. Racemic mixtures of tetrathia-[7]-helicene [27], naphthalene double [6]helicenes [28], pyrene double [4]heterohelicene [29], 5,14-diaryldiindeno[2,1-f:1′,2′-j]picene [30,31], perylene diimide double-[5]heterohelicene [32,33], and a multiple helicene containing two [5]helicene and four [4]helicene moieties [34] have been used in organic thin-film transistors (OTFTs). Some authors exploited the chirality of the helicenes by using the enantiomerically pure 1-aza[6]helicene [35] and perylene diimide double-[7]heterohelicene [36] in circularly polarized detecting OTFTs.
A handful of studies are available for other types of devices, such as organic light emitting diodes [37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56], organic solar cells [57,58,59], and hybrid DSSC [60] and perovskite solar cells [61,62,63,64,65,66,67,68,69,70,71,72].
In this paper, we report the optical, electrochemical, and structural characteristics of racemic benzo[i]pentahelicene-3,6-dione as well as its thin film, and OTFTs’ semiconductor properties.

2. Results and Discussion

2.1. Synthesis

Benzo[i]pentahelicene-3,6-dione (1, Scheme 1) was synthesized by optimizing a previously published procedure [73]. The simplicity in this design makes this system worthy of investigation. Indeed, many advanced semiconducting helical structures are extremely challenging to access and, therefore, small quantities of material are typically available for study. Basically, the synthesis of compound 1 involves the use of α-tetralone (2) as a starting material, which is subjected to self-coupling in the presence of a Zn-chlorotrimethylsilane system to afford 3,3′,4,4′-tetrahydro-l,l′-binaphthalene (3). The synthetic process was completed by the solvent-free Diels-Alder reaction of the diene 3 with ten-fold excess of 1,4-benzoquinone (4), followed by the in situ dehydrogenation of the resulting cycloadduct 5. Notably, 1,4-benzoquinone, under these conditions, behaves simultaneously as a dienophile in the Diels-Alder reaction and as an oxidizing agent. Though the yield is just acceptable (~32%, see Materials and Methods), this one-step cycloaddition-dehydrogenation reaction is simple and may enable the production of the [5]helicenebenzoquinone 1 on a large scale, starting from readily available and reasonably cheap starting materials.
The Diels-Alder reaction represents a suitable and atom-efficient method for constructing annulated ring systems containing six-member rings [74], and it consequently offers an alternative synthetic approach to helicenes [75,76,77,78,79,80,81,82,83,84].
It is noteworthy to mention that, in our strategy, the use of excess 1,4-benzoquinone is not an issue, as the final optimized protocol also includes its recovery via steam distillation followed by filtration and reuse to minimize the waste associated with the process.

2.2. Thermal Properties

Thermal stability and transitions for [5]helicenebenzoquinone were investigated by simultaneous TGA/DSC. Representative TGA/DSC analysis curves are reported in Figure 1. As shown, the thermal decomposition takes place in two main steps until 650 °C. In the first step, the quinone loses 13.9 wt% between 30 and 341 °C. In this temperature range, the DSC curve shows an endothermic peak at 247 °C, which could be assigned to a melting point, and an exothermic peak at 295 °C due to thermal degradation or evaporation process. In addition, a continual mass loss up to 650 °C (~86.5%) associated with a broad exothermic peak, appears at 631 °C in the DSC curve. As shown, the synthetized material featured remarkable thermal stability. The onset point of the weight loss with 5% weight loss temperature (decomposition temperature Td) was 268 °C, and no weight loss was observed at lower temperatures.

2.3. Optical and Electrochemical Properties

The solution and thin-film photophysical properties of the synthetized quinone were assessed by optical absorption (ultraviolet-visible (UV-vis)) spectroscopy, and cyclic voltammetry (CV) was used to identify its electrochemical behavior.
Cyclic voltammogram and absorbance spectra of the compound are shown in Figure 2, and electrochemical and optical data are summarized in Table 1.
Optical absorption and emission maxima were performed in solution and as thin films on glass substrates (Figure 2a).
In chloroform (CHCl3), the studied quinone exhibits an absorption maximum (λmax-sol) located at 486 nm and a molar extinction coefficient (ε) of 8.33 × 103 Lmol−1cm−1.
A thin film of the same compound shows a sizeable broadening and exhibits an absorption maximum (λmax-film) at 490 nm. The broadening and the small bathochromic shift (<10 nm), observed going from the solution to the thin film, is clear evidence of solid-state J-aggregate formation [85,86]. Interestingly, a tail in the [5]helicene film absorption profile can be observed, spanning from 600 to 800 nm. This might have a component [87] from spatially proximal interactions between the dipole–dipole interactions of the quinone fragment, facilitated by the tendency of the helicene molecules to self-assemble, that can modulate the spectral transition above 600 nm and create a tail in the absorption spectrum.
The band gap for the [5]helicene 1 was estimated from the low-energy band edges (Table 1). The optical band gaps of the compound in solution and solid state are estimated to be 2.11 and 2.01 eV, respectively.
The CV plot (Figure 2b) show the presence of a reversible reduction process with half-wave potentials at −1.46 V. No oxidation peaks are observed in the investigated electrochemical window.
From the reduction potential, LUMO and HOMO energies were estimated knowing that the SCE energy level is −4.8 eV below the vacuum level and the band gap obtained by optical absorption data (Egsol) (Table 1), respectively. Using these relations, the LUMO value was estimated to lie at −3.14 eV, and HOMO was calculated as −5.25 eV.

2.4. Quantum Mechanical Calculations

To gain an insight into the electronic properties of [5]helicenebenzoquinone, quantum-chemical calculations were also performed. The optimized geometry, molecular orbitals’ contour plots (MO), and energy diagram are shown in Figure 3.
The molecular orbital (MO) energies and contours were calculated by using the DFT-optimized structure of the molecule [79]. Since charge is transported through the frontier molecular orbitals of organic semiconductors, their geometries, energies, and extents of delocalization directly affect carrier stability and intermolecular hopping rates. Ideally, the HOMO and LUMO are delocalized over the entire π-core of the semiconductor for p- and n-channel transport, respectively [86].
Here, the HOMO is delocalized over the ortho-anellated moiety, while the LUMO is more compact, since it is delocalized only on the benzoquinone subunit. This highly localized nature of the LUMO limits intermolecular π-overlap and may favor charge trapping at the quinone site. This may be a contributing factor to the limited mobility values of [5]helicenebenzoquinone (vide infra). This is not uncommon behavior [88] for quinone-containing semiconductors. The computed HOMO energy is −5.77 eV, while the LUMO energy is −3.20 eV.
The DFT-calculated HOMO and LUMO energies are in good agreement with the experimental values (Figure 3 and Table 1, respectively).

2.5. Crystallographic Characterization

The X-ray analysis was carried out on a recrystallized single crystal obtained from [5]helicenebenzoquinone solution. The crystal data and structure refinement are reported in Table 2 and Tables S1–S6 (Supplementary Information).

2.6. Thin-Film Transistor Characterization

Top-contact OTFTs (1000 µm channel widths (W) and 50 µm channel lengths (L)) were fabricated by spin coating or drop casting films on 300 nm SiO2 (Bare), poly-4-vinylphenol (PVP) and PS-brush (PS-OH)-treated SiO2 substrates. The current–voltage characteristics of the fabricated organic thin film transistor (OTFT) devices were measured at room temperature and the device performance (mobility, current on/off ratio, and threshold voltage) data and representative output and transfer curves are reported in Table 3 and Figure 4, respectively. The field-effect carrier mobility (µ) of the semiconductor was calculated in the saturation regime using the conventional transistors relationship: µsat = (2IDSL)/[WCox(Vsg − Vth)2], where Isd is the source-drain saturation current; Cox is the areal capacitance value of the insulator layer, Vsg is the gate voltage, and Vth is the threshold voltage. The latter can be estimated as the x intercept of the linear section of the plot of Vsg vs. (Isd)1/2.
[5]Helicenebenzoquinone showed a p-type behavior, with hole mobilities of 10−5–10−7 cm2/Vs and current on/off ratio of 103–104.
The device response was found to be dependent on both the deposition method (SC or DC) and the dielectric surface treatment. Both have a critical impact on the π–π stacking and intermolecular interaction between the semiconductor molecules and influence the neighbouring molecular orbital overlapping. In the case of SC, although the uniformity of the thin film is generally excellent, there is a disadvantage of low crystallinity. On the other hand, DC generally results in highly crystalline thin films, whereas the uniformity is poor. Overall, devices fabricated by DC deposition performed better than the spin-coated counterparts. By comparing the performance of the SC and DC devices in Table 3, it can be implied that the slow, static process of DC facilitates preferable ordering of the [5]helicenebenzoquinone molecules compared to the SC method. The majority of the non-working and poor SC devices may be at least in part be attributed to the poor molecular ordering due to the accelerated solvent evaporation from spin-coating.
Among the devices obtained by the DC method, those on the PS brush demonstrated the best results. PS brush treatment is a widely used surface modification method which optimizes the morphology, surface energy, and viscoelastic properties of the resulting dielectrics [89]. The modification of the semiconductor/dielectric interface is crucial to the growth of the semiconductor layer, which leads to enhanced OTFT characteristics.
Representative micrographies of the thin films are reported in Figure S1 (Supplementary Information). It can be observed that in all cases they typically show gaps between (inhomogeneous) crystallites, which likely contribute to compromising charge transport. No evident correlation between TD and device performance could be identified.

3. Materials and Methods

All chemicals were purchased from Merck KGaA, Darmstadt, Germany, and used without further purification unless otherwise noted. Column chromatography was performed on silica gel (230–400 mesh American Society for Testing and Materials, ASTM) as a stationary phase. 3,3’,4,4’-tetrahydro-l,1’-binaphthalene (2) was prepared according to a previously reported procedure [73].
NMR spectra were recorded on a Varian Associates (Palo Alto, CA, USA) VXR-400 multinuclear spectrometer using deuterated chloroform (CDCl3) as the solvent (internal Me4Si). Elemental microanalyses were performed using a Fisons’(Ipswich, London, UK) EA1106 CHN analyzer using atropine (C17H23NO3), 2,5-bis-2-(5-tertbutylbenzoxazol-yl)-thiophene (C26H26N2O2S; BBOT) and phenanthrene (C14H10) as a reference standard, with an accuracy of ≈ 2 μmol g−1. The melting point was measured on a Büchi 510 instrument (Büchi Labortechnik AG, Flawil, CH) and was not corrected. Thermogravimetric (TGA) and differential scanning calorimetric (DSC) analyses were performed on a TGA/DSC simultaneous apparatus (SDT Q600, TA Instruments, New Calstle, Delaware, UK) under a nitrogen flow rate of 100 mL min−1. Samples in an open platinum cell were heated from 30 up to 1000 °C with a rate of 10 °C min−1. Absorption spectra of the compound in chloroform (CHCl3) were recorded using a Varian Cary 100 UV-vis spectrophotometer (using a glass cuvette of 1 cm path length) over the spectral range of 350–800 nm. Electrochemical properties were determined using a Bass potentiostat, with a conventional three-electrode configuration (platinum-wire counter electrode, glassy carbon working electrode, and silver/silver chloride (Ag/AgCl) reference electrode). Cyclic voltammetry was performed in an electrolyte solution of 0.1 M tetrabutylammonium hexafluorophosphate (C16H36F6NP; TBAPF6) in dichloromethane (CH2Cl2; DCM) with scan rates between 50 and 100 mVs−1. The solutions were purged with nitrogen steam for about 15 min before every experiment. The glassy carbon working electrode was polished with a slurry of alumina and ultrapure water on a Buehler felt pad and rinsed thoroughly with ultrapure water. A ferrocene/ferrocenium redox (Fc/Fc+) couple was used as an internal standard. The HOMO and LUMO energy levels were calculated from the oxidation (Eox) and reduction (Ered) potentials data using the empirical relation ELUMO = [−(Ered − E1/2-ferrocene) + 4.8] eV or EHOMO = [−(Eox − E1/2-ferrocene) + 4.8] eV. Density functional theory (DFT) computations were carried out using the Spartan ’08 software package (Wavefunction, Inc., Irvine, CA, USA) at the B3LYP/6-31G* level. Single crystals of helicene were recrystallized by layering chloroform (CHCl3) solution of the synthetized material with hexane (C6H14). Suitable crystals were mounted in inert oil and transferred to the cold gas stream of a Bruker (Billerica, MA, USA) Kappa APEX CCD area detector equipped with a MoKα sealed tube with graphite. The crystal was kept at −173.16 °C during data collection. SADABS-2012/1 (Bruker) was used for absorption correction. Using Olex2 [90], the structure was solved with the XT structure solution program using direct methods and refined with the ShelXL [91] refinement package using least squares minimization.

3.1. Diels-Alder Reaction of 3,3’,4,4’-Tetrahydro-1,1’-binaphthalene (3) with 1,4-Benzoquinone (4)

The diene 3 (5.2 g, 19.9 mmol) and 1,4-benzoquinone (4) (22 g, 0.2 mol) were heated together at 150–170 °C for 18 h in a degassed flask. The excess benzoquinone was removed by steam distillation from the crude reaction mixture, and the residue was purified by column chromatography (silica gel, hexane/ethylacetate 3:1) to afford benzo[i]pentahelicene-3,6-dione (1). Yield: 32%; m.p. 246–247 °C. The spectra matched those in the literature. 1H-NMR (CDCl3) δ 6.97 (s, 1H), 7.21 (dd, 1H, J = 8.6, 7.0 Hz), 7.54 (dd, 1H, J = 8.1, 7.0 Hz), 7.92 (d, 1H, J = 8.1 Hz), 8.04 (d, 1H, J = 9.2 Hz), 8.13 (d, 1H, J = 8.6 Hz), 9.24 (d, 1H, J = 9.2 Hz). An. Calcd. for C26H14O2: C%, 87.13; H%, 3.94. Found: C%, 87.30; H%, 3.96.

3.2. OTFTs Fabrication and Characterization

Bottom-gate/top-contact organic thin-film transistors (OTFTs) were fabricated on heavily n-doped (100) silicon substrate (resistivity < 0.005 Ωcm) with a thermally grown silicon dioxide (300 nm SiO2) as the dielectric layer. Prior to deposition, the Si/SiO2 substrate were cleaned in an ultrasonic bath using acetone (C3H6O) for 15 min, dried using a nitrogen gun, and cleaned with air plasma for 5 min (Harrick Plasma, Ithaca, NY, USA18 W). In order to reduce charge trapping and thus enhance mobility, the SiO2 substrate was modified with poly-4-vinylphenol ((C8H8O)n; PVP) or PS-brush ((CH2CH(C6H5))nOH; PS-OH). Semiconducting layers were deposited via two different solution-processing methods: spin-coating (SC) and drop-casting (DC). For the SC process, the pentahelicenebenzoquinone was dissolved in chlorobenzene (C6H5Cl), dichlorobenzene (C6H4Cl2), or toluene (C7H8), and the solutions (15 mg/mL) were spin-coated at 3000 rpm for 30 s onto the substrates. The substrates were annealed at 150 °C under vacuum overnight. For the DC process, a 0.2 wt% solution of semiconductor in chlorobenzene (C6H5Cl) was deposited onto the substrates heated at 100 and 60–70 °C for bare, PS-brush- and PVP-treated substrates, respectively. The substrates were then annealed at 130–150 °C for 5 h under vacuum. Top contacts were fabricated by vapor deposition of gold electrodes (~10−6 Torr, 0.2 Ås−1, ~40 nm thick) onto the semiconductor thin films through a shadow mask to obtain devices with a channel width (W) and length (L) of 1000 and 50 μm, respectively. The capacitance of the 300 nm SiO2 gate insulator was 11.4 nFcm−2. Characterization of the devices was performed at room temperature in ambient atmosphere using a Keithley 4200 SCS (Keithley Instruments, Cleveland, OH, USA).

4. Conclusions

A helicene quinone, namely benzo[i]pentahelicene-3,6-dione, was prepared in one-pot reaction conditions via Diels–Alder cycloaddition-dehydrogenation steps. Its thermal, optical, and electrochemical behavior, along with its single-crystal structure and electronic structure, were also reported. [5]Helicenebenzoquinone exhibited excellent thermal stability, and near UV-vis electronic absorption.
An OTFT that incorporated the helical compound as a semiconductor with the configuration of bottom-gate/top-contact attained hole mobilities in the range of 10−5–10−7 cm2/Vs and current on/off ratio of 103–104 was formulated. The carrier mobilities were found to depend on the thin-film deposition method, and surface treatments.
This study contributes to gaining insights into the structure–properties–device response relationship in the class of helically shaped semiconducting materials. The strongly localized LUMO may be at least partially responsible for the limited mobility values of [5]helicenebenzoquinone. The tendency to form J-aggregates (i.e., less orbital overlap) and the presence of significant gaps between (heterogeneous) crystallites in thin-films also possibly compromise charge transport.
On the other hand, the LUMO localization and the J-aggregation, along with the excellent thermal stability, suggest the use of [5]helicenequinone in OLED devices with thermally activated delayed fluorescence (TADF) emitters [92].
Hence, future developments will be focused on expanding the range of device applications, including organic–inorganic hybrids. This is also worth pursuing in view of the fact that the optimized synthetic method enables us to prepare the [5]helicenebenzoquinone on a large scale, which is a key requirement for studying organics’ applications as materials for electronics. In addition, its structural simplicity, compared with other advanced double or multiple helicenes, is a design principle that can be applied in the development of semiconductors, to make the leap beyond the laboratory scale.
New helical architectures will be realized by introducing, e.g., substituents at the terminal aromatic rings, to tune the helical twisting of the structure, which can ultimately result in unique stacking structures.
Their optical resolution into enantiomerically pure forms will be also pursued.

Supplementary Materials

The following supporting information can be downloaded online, Figure S1: Microscopic view of thin films obtained by (a) SC from chlorobenzene onto bare SiO2 (not annealed); (b) SC from dichlorobenzene onto bare SiO2 (annealed at 150 °C); (c) SC from dichlorobenzene onto bare PVP (annealed at 150 °C); (d) DC from chlorobenzene onto bare SiO2 (annealed at 130 °C; TD = 100 °C) and (e) onto PS brush (annealed at 150 °C; TD = 70 °C); Table S1: Fractional Atomic Coordinates (×104) and Equivalent Isotropic Displacement Parameters (Å2 × 103) for compound 1.; Table S2: Anisotropic Displacement Parameters (Å2 × 103) for compound 1.; Table S3: Bond Lengths for compound 1.; Table S4: Bond Angles for compound 1.; Table S5: Torsion Angles for compound 1; Table S6: Hydrogen Atom Coordinates (Å × 104) and Isotropic Displacement Parameters (Å2 × 103) for compound 1.

Author Contributions

Conceptualization, M.P.B. and A.M.; methodology, M.P.B., C.K. and A.M.; literature survey, A.M., M.P.B., C.K., M.L.S. and A.M.; investigation, A.M., G.K. and D.H.; writing—original draft preparation, M.P.B., C.K. and A.M.; writing—review and editing, M.P.B., M.L.S. and A.M.; supervision, M.P.B., C.K. and A.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was partially funded by “Fondo Ricerca di Base” 2019, University of Perugia, Perugia, Italy.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Liao, C.-Y.; Chen, Y.; Lee, C.-C.; Wang, G.; Teng, N.-W.; Lee, C.-H.; Li, W.-L.; Chen, Y.-K.; Li, C.-H.; Ho, H.-L.; et al. Processing strategies for an organic photovoltaic module with over 10% efficiency. Joule 2020, 4, 189–206. [Google Scholar] [CrossRef]
  2. Cui, Y.; Yao, H.; Zhang, J.; Xian, K.; Zhang, T.; Hong, L.; Wang, Y.; Xu, Y.; Ma, K.; An, C.; et al. Single-junction organic photovoltaic cells with approaching 18% efficiency. Adv. Mater. 2020, 32, 1908205. [Google Scholar] [CrossRef] [PubMed]
  3. PV Magazine. An Organic Solar Cell with 25% Efficiency. 2020. Available online: https://www.pv-magazine.com/2020/03/24/an-organic-solar-cell-with-25-efficiency/ (accessed on 29 December 2021).
  4. Acharya, R.; Peng, B.; Chan, P.K.L.; Schmitz, G.; Klauk, H. Achieving ultralow turn-on voltages in organic thin-film transistors: Investigating fluoroalkylphosphonic acid self-assembled monolayer hybrid dielectrics. ACS Appl. Mater. Interfaces 2019, 11, 27104–27111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Hosono, H.; Kim, J.; Toda, Y.; Kamiya, T.; Watanabe, S. Transparent amorphous oxide semiconductors for organic electronics: Application to inverted OLEDs. Proc. Natl. Acad. Sci. USA 2016. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Mishra, A.; Bäuerle, P. Small molecule organic semiconductors on the move: Promises for future solar energy technology. Angew. Chem. Int. Ed. 2012, 51, 2020–2067. [Google Scholar] [CrossRef] [PubMed]
  7. Pankow, R.M.; Thompson, B.C. The development of conjugated polymers as the cornerstone of Organic electronics. Polymer 2020, 207, 122874. [Google Scholar] [CrossRef]
  8. Sun, H.; Guo, X.; Facchetti, A. High-performance n-type polymer semiconductors: Applications, recent development, and challenges. Chem 2020, 6, 1310–1326. [Google Scholar] [CrossRef]
  9. Cheng, P.; Yang, Y. Narrowing the band gap: The key to high-performance organic photovoltaics. Acc. Chem. Res. 2020, 53, 1218–1228. [Google Scholar] [CrossRef]
  10. Facchetti, A. π-Conjugated Polymers for Organic Electronics and Photovoltaic Cell Applications. Chem. Mater. 2011, 23, 733–758. [Google Scholar] [CrossRef]
  11. Seri, M.; Marrocchi, A. The carbon-carbon triple bond as a tool to design organic semiconductors for photovoltaic applications: An assessment of prospects and challenges. J. Mater. Chem C 2021, 9, 16164–16186. [Google Scholar] [CrossRef]
  12. Seri, M.; Marrocchi, A.; Bagnis, D.; Ponce, R.; Taticchi, A.; Marks, T.J.; Facchetti, A. Molecular shape-controlled photovoltaic performance probed via soluble pi-conjugated arylacetylenic semiconductors. Adv. Mater. 2011, 23, 3827–3831. [Google Scholar] [PubMed]
  13. Bagnis, D.; Beverina, L.; Huang, H.; Silvestri, F.; Yao, Y.; Yan, H.; Pagani, G.A.; Marks, T.J.; Facchetti, A. Marked alkyl-vs alkenyl-substitutent effects on squaraine dye solid-state structure, carrier mobility, and bulk-heterojunction solar cell efficiency. J. Am. Chem. Soc. 2010, 132, 4074–4075. [Google Scholar] [CrossRef]
  14. Broggi, A.; Tomasi, I.; Bianchi, L.; Marrocchi, A.; Vaccaro, L. Small molecular aryl acetylenes: Chemically tailoring high-efficiency organic semiconductors for solar cells and field-effect transistors. ChemPlusChem 2014, 79, 486–507. [Google Scholar] [CrossRef] [PubMed]
  15. Shen, Y.; Chen, C.F. Helicenes: Synthesis and applications. Chem. Rev. 2012, 112, 1463–1535. [Google Scholar] [CrossRef] [PubMed]
  16. Martin, R.H. The Helicenes. Angew. Chem. Int. Ed. Engl. 1974, 13, 649–660. [Google Scholar] [CrossRef]
  17. Laarhoven, W.H.; Prinsen, W.J.C. Carbohelicenes and Heterohelicenes. Top. Curr. Chem. 1984, 125, 63–130. [Google Scholar]
  18. Meurer, K.P.; Vögtle, F. Helical Molecules in Organic Chemistry. Top. Curr. Chem. 1985, 127, 1–76. [Google Scholar]
  19. Mori, T. Chiroptical Properties of Symmetric Double, Triple, and Multiple Helicenes. Chem. Rev. 2021, 121, 2373–2412. [Google Scholar] [CrossRef]
  20. Grimme, S.; Harren, J.; Sobanski, A.; Vögtle, F. Structure/Chiroptics Relationships of Planar Chiral and Helical Molecules. Eur. J. Org. Chem. 1998, 1998, 1491–1509. [Google Scholar] [CrossRef]
  21. Groen, M.B.; Wynberg, H. Optical Properties of Some Heterohelicenes. Absolute Configuration. J. Am. Chem. Soc. 1971, 93, 2968–2974. [Google Scholar] [CrossRef]
  22. Verbiest, T.; Elshocht, S.V.; Kauranen, M.; Hellemans, L.; Snauwaert, J.; Nuckolls, C.; Katz, T.J.; Persoons, A. Strong enhancement of nonlinear optical properties through supramolecular chirality. Science 1998, 282, 913–915. [Google Scholar] [CrossRef] [PubMed]
  23. Nuckolls, C.; Katz, T.J.; Castellanos, L. Aggregation of conjugated helical molecules. J. Am. Chem. Soc. 1996, 118, 3767–3768. [Google Scholar] [CrossRef]
  24. Rybáček, J.; Stará, I.G.; Starý, I.; Rahe, P.; Nimmrich, M.; Kühnle, A. Racemic and Optically Pure Heptahelicene-2-carboxylic Acid: Its Synthesis and Self-Assembly into Nanowire-Like Aggregates. Eur. J. Org. Chem. 2011, 853–860. [Google Scholar] [CrossRef] [Green Version]
  25. Katz, T.J. Syntheses of Functionalized and Aggregating Helical Conjugated Molecules. Angew. Chem. Int. Ed. 2000, 39, 1921–1923. [Google Scholar] [CrossRef]
  26. Lewis, F.D.; Liu, X.; Wu, Y.; Zuo, X. DNA-Mediated Exciton Coupling and Electron Transfer between Donor and Acceptor Stilbenes Separated by a Variable Number of Base Pairs. J. Am. Chem. Soc. 2003, 125, 12729. [Google Scholar] [CrossRef]
  27. Kim, C.; Marks, T.J.; Facchetti, A.; Schiavo, M.; Bossi, A.; Maiorana, S.; Licandro, E.; Todescato, F.; Toffanin, S.; Muccini, M.; et al. Synthesis, characterization, and transistor response of tetrathia-[7]-helicene precursors and derivatives. Org. Electron. 2009, 10, 1511–1520. [Google Scholar] [CrossRef]
  28. Fujikawa, T.; Mitoma, N.; Wakamiya, A.; Saeki, A.; Segawa, Y.; Itami, K. Synthesis, properties, and crystal structures of π-extended double [6]helicenes: Contorted multi-dimensional stacking lattice. Org. Biomol. Chem. 2017, 15, 4697–4703. [Google Scholar] [CrossRef]
  29. Moriguchi, T.; Higashi, M.; Yakeya, D.; Jalli, V.; Tsuge, A.; Okauchi, T.; Nagamatsu, S.; Takashima, W. Synthesis, characterization and air stable semiconductor properties of thiophene-condensed pyrene derivatives, J. Mol. Struct. 2017, 1127, 413–418. [Google Scholar] [CrossRef]
  30. Hsieh, Y.-C.; Wu, C.F.; Chen, Y.-T.; Fang, C.-T.; Wang, C.-S.; Li, C.-H.; Chen, L.-Y.; Cheng, M.-J.; Chueh, C.-C.; Chou, P.-T.; et al. 5,14-Diaryldiindeno[2,1-f:1′,2′-j]picene: A New Stable [7] Helicene with a Partial Biradical Character. J. Am. Chem. Soc. 2018, 140, 14357–14366. [Google Scholar] [CrossRef]
  31. Shen, S.-W.; Chen, D.G.; Chen, I.-T.; Chang, K.H.; Lee, C.-W.; Fang, C.T.; Chen, Y.T.; Chuang, W.T.; Lee, Y.H.; Wu, Y.T.; et al. Delayed Charge Recombination by Open-Shell Organics: Its Application in Achieving Superb Photodetectors with Broadband (400–1160 nm) Ultrahigh Sensitivity and Stability. Adv. Opt. Mater. 2020, 8, 1902179. [Google Scholar] [CrossRef]
  32. Schaack, C.; Evans, A.M.; Ng, F.; Steigerwald, M.L.; Nuckolls, C. High-Performance Organic Electronic Materials by Contorting Perylene Diimides. J. Am. Chem. Soc. 2021, 144, 42–51. [Google Scholar] [CrossRef] [PubMed]
  33. Zeng, C.; Xiao, C.; Feng, X.; Zhang, L.; Jiang, W.; Wang, Z. Electron-Transporting Bis (heterotetracenes) with Tunable Helical Packing. Angew. Chem. Int. Ed. 2018, 57, 10933. [Google Scholar] [CrossRef] [PubMed]
  34. Yang, Y.; Yuan, L.; Shan, B.; Liu, Z.; Miao, Q. Twisted Polycyclic Arenes from Tetranaphthyldiphenylbenzenes by Controlling the Scholl Reaction with Substituents. Chem. Eur. J. 2016, 22, 18620. [Google Scholar] [CrossRef] [PubMed]
  35. Yang, Y.; Correa da Costa, R.; Fuchter, M.J.; Campbell, A.J. Circularly polarized light detection by a chiral organic semiconductor transistor. Nat. Photonics 2013, 7, 634–638. [Google Scholar] [CrossRef]
  36. Zhang, L.; Song, I.; Ahn, J.; Han, M.; Linares, M.; Surin, M.; Zhang, H.-J.; Oh, J.H.; Lin, J. π-Extended perylene diimide double-heterohelicenes as ambipolar organic semiconductors for broadband circularly polarized light detection. Nat. Commun. 2021, 12, 142. [Google Scholar] [CrossRef]
  37. Oda, S.; Kawakami, B.; Yamasaki, Y.; Matsumoto, R.; Yoshioka, M.; Fukushima, D.; Nakatsuka, S.; Hatakeyama, T. One-Shot Synthesis of Expanded Heterohelicene Exhibiting Narrowband Thermally Activated Delayed Fluorescence. J. Am. Chem. Soc. 2021. [Google Scholar] [CrossRef]
  38. Shi, Y.; Yang, G.; Shen, B.; Yang, Y.; Yan, L.; Yang, F.; Liu, J.; Liao, X.; Yu, P.; Bin, Z.; et al. Insight into Regioselective Control in Aerobic Oxidative C-H/C-H Coupling for C3-Arylation of Benzothiophenes: Toward Structurally Nontraditional OLED Materials. J. Am. Chem. Soc. 2021, 143, 21066–21076. [Google Scholar] [CrossRef]
  39. Li, M.; Wang, Y.-F.; Zhang, D.-W.; Zhang, D.; Hu, Z.-Q.; Duan, L.; Chen, C.-F. Thermally activated delayed fluorescence material-sensitized helicene enantiomer-based OLEDs: A new strategy for improving the efficiency of circularly polarized electroluminescence. Sci. China Chem. 2021, 64, 899–908. [Google Scholar] [CrossRef]
  40. Dhbaibi, K.; Abella, L.; Meunier-Della-Gatta, S.; Roisnel, T.; Vanhuyne, N.; Jamoussi, B.; Pieters, G.; Racine, B.; Quesnel, E.; Autschbach, J.; et al. Achieving high circularly polarized luminescence with push-pull helicenic systems: From rationalized design to top-emission CP-OLED applications. Chem. Sci. 2021, 12, 5522–5533. [Google Scholar] [CrossRef]
  41. Ikari, Y.; Kaihara, T.; Goto, S.; Bovenkerk, M.; Grenz, D.C.; Esser, B.; Ferreira, M.; Stachelek, P.; Data, P.; Yoshida, T.; et al. Peripherally Donor-Installed 7,8-Diaza[5]helicenes as a Platform for Helical Luminophores. Synthesis 2021, 53, 1584–1596. [Google Scholar]
  42. Yan, Z.-P.; Luo, X.-F.; Liao, K.; Zheng, Y.-X.; Zuo, J.-L. Rational design of the platinahelicene enantiomers for deep-red circularly polarized organic light-emitting diodes. Front. Chem. 2020, 8, 501. [Google Scholar] [CrossRef] [PubMed]
  43. Zhao, Y.; Ding, J.; Han, X.; Geng, T.; Zhou, X.; Hu, C.; Wang, Y.; Xiao, G.; Zou, B.; Hou, H. Tuning the optical properties of N-aryl benzothiadiazole via Cu(II)-catalyzed intramolecular C-H amination: The impact of the molecular structure on aggregation and solid state luminescence. Org. Chem. Front. 2020, 7, 3853–3861. [Google Scholar] [CrossRef]
  44. Wan, L.; Wade, J.; Shi, X.; Xu, S.; Fuchter, M.J.; Campbell, A.J. Highly Efficient Inverted Circularly Polarized Organic Light-Emitting Diodes. ACS Appl. Mater. Interfaces 2020, 12, 39471–39478. [Google Scholar] [CrossRef] [PubMed]
  45. Yan, Z.-P.; Luo, X.-F.; Liu, W.-Q.; Wu, Z.-G.; Liang, X.; Liao, K.; Wang, Y.; Zheng, Y.-X.; Zhou, L.; Zuo, J.-L.; et al. Configurationally Stable Platinahelicene Enantiomers for Efficient Circularly Polarized Phosphorescent Organic Light-Emitting Diodes. Chem. Eur. J. 2019, 25, 5672–5676. [Google Scholar] [CrossRef]
  46. Yavari, K.; Delaunay, W.; De Rycke, N.; Reynaldo, T.; Aillard, P.; Srebro-Hooper, M.; Chang, V.Y.; Muller, G.; Tondelier, D.; Geffroy, B.; et al. Phosphahelicenes: From Chiroptical and Photophysical Properties to OLED Applications. Chem. Eur. J. 2019, 25, 5303–5310. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Klimash, A.; Pander, P.; Klooster, W.T.; Coles, S.J.; Data, P.; Dias, F.B.; Skabara, P.J. Intermolecular interactions in molecular crystals and their effect on thermally activated delayed fluorescence of helicene-based emitters. J. Mater. Chem. C 2018, 6, 10557–10568. [Google Scholar] [CrossRef] [Green Version]
  48. Jhulki, S.; Mishra, A.K.; Chow, T.J.; Moorthy, J.N. Carbo[5]helicene versus planar phenanthrene as a scaffold for organic materials in OLEDs: The electroluminescence of anthracene-functionalized emissive materials. New J. Chem. 2017, 41, 14730–14737. [Google Scholar] [CrossRef]
  49. Chen, T.; Zhang, B.; Liu, Z.; Duan, L.; Dong, G.; Feng, Y.; Luo, X.; Cui, D. Synthesis and properties of thiophene-substituted diaza[7]helicene for application as a blue emitter in organic light-emitting diodes. Tetrahedron Lett. 2017, 58, 531–535. [Google Scholar] [CrossRef]
  50. Sahasithiwat, S.; Sooksimuang, T.; Kangkaew, L.; Panchan, W. 3,12-Dimethoxy-5,6,9,10-tetrahydro-7,8-dicyano-[5]helicene as a new emitter for blue and white organic light-emitting diodes. Dyes Pigm. 2017, 136, 754–760. [Google Scholar] [CrossRef]
  51. Brandt, J.R.; Wang, X.; Yang, Y.; Campbell, A.J.; Fuchter, M.J. Circularly Polarized Phosphorescent Electroluminescence with a High Dissymmetry Factor from PHOLEDs Based on a Platinahelicene. J. Am. Chem. Soc. 2016, 138, 9743–9746. [Google Scholar] [CrossRef] [Green Version]
  52. Hirai, H.; Nakajima, K.; Nakatsuka, S.; Shiren, K.; Ni, J.; Nomura, S.; Ikuta, T.; Hatakeyama, T. One-Step Borylation of 1,3-Diaryloxybenzenes Towards Efficient Materials for Organic Light-Emitting Diodes. Angew. Chem. Int. Ed. 2015, 54, 13581–13585. [Google Scholar] [CrossRef] [PubMed]
  53. Hua, W.; Liu, Z.; Duan, L.; Dong, G.; Qiu, Y.; Zhang, B.; Cui, D.; Tao, X.; Cheng, N.; Liu, Y. Deep-blue electroluminescence from nondoped and doped organic light-emitting diodes (OLEDs) based on a new monoaza[6]helicene. RSC Adv. 2015, 5, 75–84. [Google Scholar] [CrossRef]
  54. Shi, L.; Liu, Z.; Dong, G.; Duan, L.; Qiu, Y.; Jia, J.; Guo, W.; Zhao, D.; Cui, D.; Tao, X. Synthesis, Structure, Properties, and Application of a Carbazole-Based Diaza [7] helicene in a Deep-Blue-Emitting OLED. Chem. Eur. J. 2012, 18, 8092–8099. [Google Scholar] [CrossRef]
  55. Sahasithiwat, S.; Mophuang, T.; Menbangpung, L.; Kamtonwong, S.; Sooksimuang, T. 3,12-Dimethoxy-7,8-dicyano-[5] helicene as a novel emissive material for organic light-emitting diode. Synth. Met. 2010, 160, 1148–1152. [Google Scholar] [CrossRef]
  56. Chen, J.P. [5]-Helicene and dibenzofluorene materials for use in organic light-emitting devices. Patent Application Number PCT/US2002/039350, 10 December 2002. [Google Scholar]
  57. Ma, Z.; Winands, T.; Liang, N.; Meng, D.; Jiang, W.; Doltsinis, N.L.; Wang, Z. A C2-symmetric triple [5]helicene based on N-annulated triperylene hexaimide for chiroptical electronics. Sci. China Chem. 2020, 63, 208–214. [Google Scholar] [CrossRef]
  58. Josse, P.; Favereau, L.; Shen, C.; Dabos-Seignon, S.; Blanchard, P.; Cabanetos, C.; Crassous, J. Enantiopure versus Racemic Naphthalimide End-Capped Helicenic Non-fullerene Electron Acceptors: Impact on Organic Photovoltaics Performance. Chem. Eur. J. 2017, 23, 6277–6281. [Google Scholar] [CrossRef] [Green Version]
  59. Lewinska, G.; Danel, K.S.; Sanetra, J. The bulk heterojunction cells based on new helicenes—Preparation, implementation and surface examination. Sol. Energy 2016, 135, 848–853. [Google Scholar] [CrossRef]
  60. Dova, D.; Cauteruccio, S.; Manfredi, N.; Prager, S.; Dreuw, A.; Arnaboldi, S.; Mussini, P.R.; Licandro, E.; Abbotto, A. An unconventional helical push-pull system for solar cells. Dye. Pigm. 2019, 161, 382–388. [Google Scholar] [CrossRef]
  61. Xu, N.; Zhang, Y.; Fang, L.; Li, T.; Xie, X.; Zhang, J.; Wang, P. A Helical Polycycle Molecular Semiconductor for Durable and Efficient Perovskite Solar Cells. ACS Mater. Lett. 2022, 4, 11–20. [Google Scholar] [CrossRef]
  62. Tang, Z.; Li, T.; Cao, Y.; Zhang, Y.; He, L.; Zheng, A.; Lei, M. Chrysene-Based Azahelicene π-Linker of D-π-D-Type Hole-Transporting Materials for Perovskite Solar Cells. ChemSusChem 2021, 14, 4923–4928. [Google Scholar] [CrossRef]
  63. Wei, Y.; Zheng, A.; Xie, X.; Zhang, J.; He, L.; Wang, P. A Pyrrole-Bridged Bis(oxa [5] helicene)-Based Molecular Semiconductor for Efficient and Durable Perovskite Solar Cells: Microscopic Insights. ACS Mater. Lett. 2021, 3, 947–955. [Google Scholar] [CrossRef]
  64. Lin, Y.-S.; Abate, S.Y.; Wang, C.-I.; Wen, Y.-S.; Chen, C.-I.; Hsu, C.-P.; Chueh, C.-C.; Tao, Y.-T.; Sun, S.-S. Low-Cost Hole-Transporting Materials Based on Carbohelicene for High-Performance Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2021, 13, 20051–20059. [Google Scholar] [CrossRef] [PubMed]
  65. Wang, J.; Wang, Y.; Xie, X.; Ren, Y.; Zhang, B.; He, L.; Zhang, J.; Wang, L.-D.; Wang, P. A Helicene-Based Molecular Semiconductor Enables 85 °C Stable Perovskite Solar Cells. ACS Energy Lett. 2021, 6, 1764–1772. [Google Scholar] [CrossRef]
  66. Zheng, A.; Xie, X.; Wang, Y.; Xu, N.; Zhang, J.; Yuan, Y.; Wang, P. A Triple Axial Chirality, Racemic Molecular Semiconductor Based on Thiahelicene and Ethylenedioxythiophene for Perovskite Solar Cells: Microscopic Insights on Performance Enhancement. Adv. Funct. Mater. 2021, 31, 2009854. [Google Scholar] [CrossRef]
  67. Wang, J.; Shi, H.; Xu, N.; Zhang, J.; Yuan, Y.; Lei, M.; Wang, L.; Wang, P. Aza [5] helicene Rivals N-Annulated Perylene as π-Linker of D-π-D Typed Hole-Transporters for Perovskite Solar Cells. Adv. Funct. Mater. 2020, 30, 2002114. [Google Scholar] [CrossRef]
  68. Xu, N.; Zheng, A.; Wei, Y.; Yuan, Y.; Zhang, J.; Lei, M.; Wang, P. D-p-D molecular semiconductors for perovskite solar cells: The superior role of helical versus planar p-linkers. Chem. Sci. 2020, 11, 3418–3426. [Google Scholar] [CrossRef] [Green Version]
  69. Ren, M.; Wang, J.; Xie, X.; Zhang, J.; Wang, P. Double-Helicene-Based Hole-Transporter for Perovskite Solar Cells with 22% Efficiency and Operation Durability. ACS Energy Lett. 2019, 4, 2683–2688. [Google Scholar] [CrossRef]
  70. Xu, N.; Li, Y.; Ricciarelli, D.; Wang, J.; Mosconi, E.; Yuan, Y.; De Angelis, F.; Zakeeruddin, S.M.; Gratzel, M.; Wang, P. An Oxa[5]helicene-Based Racemic Semiconducting Glassy Film for Photothermally Stable Perovskite Solar Cells. iScience 2019, 15, 234–242. [Google Scholar] [CrossRef] [Green Version]
  71. Lee, C.-C.; Chen, C.-I.; Fang, C.-T.; Huang, P.-Y.; Wu, Y.-T.; Chueh, C.-C. Improving Performance of Perovskite Solar Cells Using [7]Helicenes with Stable Partial Biradical Characters as the Hole-Extraction Layers. Adv. Funct. Mater. 2019, 29, 1808625. [Google Scholar] [CrossRef]
  72. Lin, Y.-S.; Abate, S.Y.; Lai, K.-W.; Chu, C.-W.; Lin, Y.-D.; Tao, Y.-T.; Sun, S.-S. New Helicene-Type Hole-Transporting Molecules for High-Performance and Durable Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2018, 10, 41439–41449. [Google Scholar] [CrossRef]
  73. Minuti, L.; Taticchi, A.; Marrocchi, A.; Gacs-Baitz, E. Diels-Alder reaction of 3,3’,4,4’-tetrahydro-1,1’-binaphthalene. One pot synthesis of a pentahelicenebenzoquinone. Tetrahedron 1997, 53, 6873–6878. [Google Scholar] [CrossRef]
  74. Fringuelli, F.; Taticchi, A. Diels Alder reaction. In Selected Practical Methods; John Wiley and Sons: New York, NY, USA, 2002. [Google Scholar]
  75. Willmore, N.D.; Liu, L.; Katz, T.J. A Diels–Alder Route to [5]- and [6]-Helicenes. Angew. Chem. Int. Ed. 1992, 31, 1093–1095. [Google Scholar] [CrossRef]
  76. Katz, T.J.; Liu, L.; Willmore, N.D.; Fox, J.M.; Rheingold, A.L.; Shi, S.; Nuckolls, C.; Rickman, B.H. An Efficient Synthesis of Functionalized Helicenes. J. Am. Chem. Soc. 1997, 119, 10054–10063. [Google Scholar] [CrossRef]
  77. Carreño, M.C.; Hernández-Sánchez, R.; Mahugo, J.; Urbano, A. Enantioselective Approach to Both Enantiomers of Helical Bisquinones. J. Org. Chem. 1999, 64, 1387–1390. [Google Scholar] [CrossRef]
  78. Minuti, L.; Taticchi, A.; Marrocchi, A.; Gacs-Baitz, E. A new short route to hexahelicenes. Synth. Commun. 1998, 28, 2181–2190. [Google Scholar] [CrossRef]
  79. Minuti, L.; Taticchi, A.; Marrocchi, A.; Gacs-Baitz, E.; Galeazzi, R. An efficient synthetic approach to substituted penta- and hexahelicenes. Eur. J. Org. Chem. 1999, 11, 3155–3163. [Google Scholar] [CrossRef]
  80. Marrocchi, A.; Minuti, L.; Taticchi, A.; Dix, I.; Hopf, H.; Jones, P.G.; Gacs-Baitz, E. The preparation of helical cyclophanes containing five-membered rings. Eur. J. Org. Chem. 2001, 22, 4259–4268. [Google Scholar] [CrossRef]
  81. Minuti, L.; Taticchi, A.; Lanari, D.; Marrocchi, A.; Gacs-Baitz, E. Synthesis of enantiopure helical cyclophanes containing five-membered heterocyclic rings. Tetrahedron Asymm. 2003, 14, 2775–2779. [Google Scholar] [CrossRef]
  82. Minuti, L.; Taticchi, A.; Marrocchi, A.; Lanari, D.; Tesei, I.; Gacs-Baitz, E. Synthesis of helical [2.2]paracyclophanes containing carbocyclic and heterocyclic five-membered rings. Tetrahedron 2004, 60, 11759–11764. [Google Scholar]
  83. Minuti, L.; Taticchi, A.; Marrocchi, A.; Gacs-Baitz, E. Synthesis of helicenophanes containing two carbocyclic five-membered rings. Polycycl. Aromat. Compd. 2005, 25, 13–22. [Google Scholar] [CrossRef]
  84. Minuti, L.; Taticchi, A.; Gacs-Baitz, E.; Marrocchi, A. High pressure Diels-Alder reaction of 2-vinyl-3,4-dihydronaphthalene. Synthesis of ciclopenta[c]- and indeno[c]-phenanthrenones. Tetrahedron 1995, 51, 8953–8958. [Google Scholar] [CrossRef]
  85. Yoon, M.-H.; Facchetti, A.; Stern, C.E.; Marks, T.J. Fluorocarbon-Modified Organic Semiconductors: Molecular Architecture, Electronic, and Crystal Structure Tuning of Arene- versus Fluoroarene-Thiophene Oligomer Thin-Film Properties. J. Am. Chem. Soc. 2006, 128, 5792–5801. [Google Scholar] [CrossRef] [PubMed]
  86. Facchetti, A.; Mushrush, M.; Yoon, M.-H.; Hutchison, G.R.; Ratner, M.A.; Marks, T.J. Building Blocks for n-Type Molecular and Polymeric Electronics. Perfluoroalkyl- versus Alkyl-Functionalized Oligothiophenes (nT; n = 2−6). Systematics of Thin Film Microstructure, Semiconductor Performance, and Modeling of Majority Charge Injection in Field-Effect Transistors. J. Am. Chem. Soc. 2004, 126, 13859–13874. [Google Scholar] [PubMed]
  87. Chen, L.X.; Shaw, G.B.; Tiede, D.M.; Zuo, X.; Zapol, P.; Redfern, P.C.; Curtiss, L.A.; Sooksimuang, T.; Mandal, B.K. Excited State Dynamics and Structures of Functionalized Phthalocyanines. 1. Self-Regulated Assembly of Zinc Helicenocyanine. J. Phys. Chem. B 2005, 109, 16598–16609. [Google Scholar] [CrossRef]
  88. Letizia, J.A.; Cronin, S.; Ponce Ortiz, R.; Facchetti, A.; Ratner, M.A.; Marks, T.J. Phenacyl–Thiophene and Quinone Semiconductors Designed for SolutionProcessability and Air-Stability in High Mobility n-Channel Field-Effect Transistors. Chem. Eur. J. 2010, 16, 1911–1928. [Google Scholar] [CrossRef]
  89. Park, S.H.; Lee, H.S.; Kim, J.-D.; Breiby, D.W.; Kim, E.; Park, Y.D.; Ryu, D.Y.; Lee, D.Y.; Cho, J.H. A polymer brush organic interlayer improves the overlying pentacene nanostructure and organic field-effect transistor performance. J. Mater. Chem. 2011, 21, 15580–15586. [Google Scholar] [CrossRef]
  90. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Cryst. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  91. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr. A 2008, A64, 112–122. [Google Scholar] [CrossRef] [Green Version]
  92. Skaisgiris, R.; Serevičius, T.; Dodonova, J.; Banevičius, D.; Kazlauskas, K.; Tumkevičius, S.; Juršėnas, S. Tuning of HOMO-LUMO localization for achieving thermally activated delayed fluorescence. J. Lumin. 2022, 241, 118473. [Google Scholar] [CrossRef]
Scheme 1. Retrosynthesis of [5]helicenebenzoquinone 1.
Scheme 1. Retrosynthesis of [5]helicenebenzoquinone 1.
Molecules 27 00863 sch001
Figure 1. TGA (solid line) and DSC (dotted line) thermograms of the [5]helicenebenzoquinone (exotherm up).
Figure 1. TGA (solid line) and DSC (dotted line) thermograms of the [5]helicenebenzoquinone (exotherm up).
Molecules 27 00863 g001
Figure 2. (a) UV-vis normalized absorptions of [5]helicenebenzoquinone in solution (CHCl3) and as a thin film; (b) cyclic voltammetry; 10−3 M CH2Cl2, 0.1 M NBu4PF6, T = 25 °C, scan rate: 100 mVs−1, reference electrode: Ag/AgCl, working electrode: glassy carbon. Fc/Fc+ was used as internal reference (E1/2-ferrocene = 0.21 V).
Figure 2. (a) UV-vis normalized absorptions of [5]helicenebenzoquinone in solution (CHCl3) and as a thin film; (b) cyclic voltammetry; 10−3 M CH2Cl2, 0.1 M NBu4PF6, T = 25 °C, scan rate: 100 mVs−1, reference electrode: Ag/AgCl, working electrode: glassy carbon. Fc/Fc+ was used as internal reference (E1/2-ferrocene = 0.21 V).
Molecules 27 00863 g002
Figure 3. Computational (DFT) values of HOMO/LUMO/gap energies, MO electron density contours and structure of [5]helicenebenzoquinone.
Figure 3. Computational (DFT) values of HOMO/LUMO/gap energies, MO electron density contours and structure of [5]helicenebenzoquinone.
Molecules 27 00863 g003
Figure 4. Transfer plot of the OTFT device obtained via (a) the SC method onto bare SiO2 (not annealed), and (b) the DC method onto the PS brush with annealing treatment (150 °C).
Figure 4. Transfer plot of the OTFT device obtained via (a) the SC method onto bare SiO2 (not annealed), and (b) the DC method onto the PS brush with annealing treatment (150 °C).
Molecules 27 00863 g004
Table 1. Thermal, electrochemical and spectroscopic characterization of [5]helicenebenzoquinone 1.
Table 1. Thermal, electrochemical and spectroscopic characterization of [5]helicenebenzoquinone 1.
m.p
(°C)
E1/2-ox (V)E1/2-red
(V)
λmax-sol
(nm) a
ε
(Lmol−1cm1)
Egsol (eV) aλmax-film
(nm) b
Egfilm (eV) bLUMO (eV) cHOMO (eV) d
246-−1.674868.33 · 1032.114902.01−3.14−5.25
a Optical band gap estimated from the low-energy band edge in the optical spectrum. Concentrations are 10−4–10−5 M in CHCl3; b Optical bandgap estimated from absorption onset (Eg = 1240/λonset) of spin-cast film from 5 mg/mL of CHCl3 solution on glass substrates; c ELUMO = [−(Ered − E1/2-ferrocene) + 4.8]; d EHOMO = ELUMO − Eopt.gap.
Table 2. Summary of crystal data and structure refinement.
Table 2. Summary of crystal data and structure refinement.
ParameterData
Empirical FormulaC26H14O2
Formula weight358.37
Temperature/°C−173.16
Crystal systemOrthorhombic
Space groupPbca (no. 61)
a /Å, b/Å, c/Å12.1651 (3), 7.4140 (2), 36.6996 (7)
α/°, β/°, γ/°90, 90, 90
Volume / Å33310.01 (14)
Z8
ρcalc/mg mm−31.438
μ(MoKα)/mm−10.090
F (000)1488
Crystal size/mm30.24 × 0.15 × 0.069
2Θ range for data collection5.562 to 56.558°
Index ranges−15 ≤ h ≤ 16, −9 ≤ k ≤ 9, −48 ≤ l ≤ 48
Reflections collected55,869
Independent reflections4094 [R(int) = 0.1036]
Data/restraints/parameters4094/0/253
Goodness-of-fit on F21.035
Final R indexes [I > 2σ (I)]R1 = 0.0459, wR2 = 0.1030
Final R indexes [all data]R1 = 0.0698, wR2 = 0.1151
Largest diff. peak/hole/e Å−30.301/−0.228
Table 3. Hole mobilities (cm2/V−1s−1), current on/off ratios (Ion/Ioff), and threshold voltages (VT, V) for films deposited at various substrates, substrate temperatures (TD, °C) and annealing temperatures (TA, °C); nw: not working.
Table 3. Hole mobilities (cm2/V−1s−1), current on/off ratios (Ion/Ioff), and threshold voltages (VT, V) for films deposited at various substrates, substrate temperatures (TD, °C) and annealing temperatures (TA, °C); nw: not working.
SubstrateSolventTDTAµhIon/IoffVT
Spin-Coated (SC)
Barechlorobenzene--1.6 × 10−51.5 × 1031.5
Baredichlorobenzene-150nw
BareToluene-150nw
PVPdichlorobenzene-150nw
PVPToluene-150nw
PS brushdichlorobenzene-150nw
PS brushToluene-150nw
Drop-casted (DC)
Barechlorobenzene1001301.5 × 10−73.0 × 104−23
PVPchlorobenzene1001305.3 × 10−56.0 × 103−10
PS brushdichlorobenzene601506.0 × 10−53.0 × 103−67
PS brushchlorobenzene70-2.5 × 10−65.0 × 103−35
PS brushchlorobenzene701504.8 × 10−52.5 × 104−23
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bracciale, M.P.; Kwon, G.; Ho, D.; Kim, C.; Santarelli, M.L.; Marrocchi, A. Synthesis, Characterization, and Thin-Film Transistor Response of Benzo[i]pentahelicene-3,6-dione. Molecules 2022, 27, 863. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27030863

AMA Style

Bracciale MP, Kwon G, Ho D, Kim C, Santarelli ML, Marrocchi A. Synthesis, Characterization, and Thin-Film Transistor Response of Benzo[i]pentahelicene-3,6-dione. Molecules. 2022; 27(3):863. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27030863

Chicago/Turabian Style

Bracciale, Maria Paola, Guhyun Kwon, Dongil Ho, Choongik Kim, Maria Laura Santarelli, and Assunta Marrocchi. 2022. "Synthesis, Characterization, and Thin-Film Transistor Response of Benzo[i]pentahelicene-3,6-dione" Molecules 27, no. 3: 863. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27030863

Article Metrics

Back to TopTop