Next Article in Journal
28-[1-(3-(Propionyloxy)propyl)-1H-1,2,3-triazol-4-yl]carbonylbetulin
Previous Article in Journal
8,13-Dimethylicosa-9,11-diyne-8,13-diol
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Short Note

4,4-Bis(2-ethylhexyl)-6-(9-(2-ethylhexyl)-2,3,4,4a,9,9a-hexahydro-1H-carbazol-6-yl)-4H-cyclopenta[2,1-b:3,4-b′]dithiophene-2-carbaldehyde

by
Maxim S. Mikhailov
1,2,
Olga O. Ustimenko
1,3,
Ekaterina A. Knyazeva
1,* and
Oleg A. Rakitin
1
1
N. D. Zelinsky Institute of Organic Chemistry Russian Academy of Sciences, 47 Leninsky Prospekt, Moscow 119991, Russia
2
Nanotechnology Education and Research Center, South Ural State University, 76 Lenina Avenue, Chelyabinsk 454080, Russia
3
Chemistry Department, Lomonosov Moscow State University, GSP-1, Leninskie Gory, Moscow 111033, Russia
*
Author to whom correspondence should be addressed.
Molbank 2022, 2022(4), M1486; https://doi.org/10.3390/M1486
Submission received: 20 October 2022 / Revised: 3 November 2022 / Accepted: 4 November 2022 / Published: 8 November 2022
(This article belongs to the Section Organic Synthesis)

Abstract

:
Dyes with a donor–π–spacer–acceptor (D-π-A) structure containing a dicyanovinyl group as an acceptor have recently been of interest for the production of single-component organic solar cells. The most convenient precursors for their synthesis are the corresponding aldehydes. In this communication, 4,4-bis(2-ethylhexyl)-6-(9-(2-ethylhexyl)-2,3,4,4a,9,9a-hexahydro-1H-carbazol-6-yl)-4H-cyclopenta[2,1-b:3,4-b′]dithiophene-2-carbaldehyde was synthesized by the Suzuki cross-coupling reaction between 4,4-bis(2-ethylhexyl)-4H-cyclopenta[2,1-b:3,4-b′]dithiophene-2,6-dicarbaldehyde and 9-(2-ethylhexyl)-6-(4,4,5,5-tetramethyl-1,3,2-dioxaborolan-2-yl)-2,3,4,4a,9,9a-hexahydro-1H-carbazole in the presence of tetrakis(triphenylphosphine)palladium(0). The structure of the newly synthesized compound was established by means of high-resolution mass spectrometry, 1H, 13C NMR, IR, and UV spectroscopy.

1. Introduction

Organic solar cells (OSCs) have received unrelenting interest over the past 50 years [1]. The development of various solar cell components, such as narrow-gap π-conjugated polymers [2,3], molecular donors [4,5], non-fullerene acceptors [6,7], has made it possible to increase power conversion efficiency (PCE) in donor–acceptor bulk heterojunction OSCs (BHJ) from 1% to 20% over the past two decades [8]. However, the industrial development of OSCs is constrained by issues of cost, synthesis scalability, and stability of organic dyes [1,9,10]. To solve the problem of reducing the cost of OSC production, the efforts of chemists in recent years have been focused on simplifying the structure of active materials and devices, including single-component organic solar cells (SMOSCs) [11,12]. As a rule, such simple molecules have a donor–π–spacer–acceptor (D-π-A) structure, where the dicyanovinyl group acts as an acceptor [11,12,13,14]. The main method for introducing this group is the Knoevenagel reaction of malononitrile with the corresponding aldehydes. Therefore, the synthesis of the aldehyde precursor is the most important step in this short reaction scheme. Herein, we report the synthesis of 4,4-bis(2-ethylhexyl)-6-(9-(2-ethylhexyl)-2,3,4,4a,9,9a-hexahydro-1H-carbazol-6-yl)-4H-cyclopenta[2,1-b:3,4-b′]dithiophene-2-carbaldehyde 1 as a key precursor for the preparation of the organic solar cells component.

2. Results and Discussion

π–Spacers containing a thiophene ring have demonstrated their perfect suitability in organic dyes for DSSCs [15]. Among them, one of the most striking examples is the 4H-cyclopenta[2,1-b:3,4-b′]dithiophene fragment, the use of which in organic solar cells leads to the highest PCE values [16]. On the other hand, among the donor building blocks, one of the most interesting is 2,3,4,4a,9,9a-hexahydro-1H-carbazole, the introduction of which into the dye molecule also gives an increase in the photovoltaic efficiency [17,18]. We hypothesized that the combination of these two building blocks in a sensitizer molecule could lead to interesting results. It was found that Suzuki cross-coupling reaction between 4,4-bis(2-ethylhexyl)-4H-cyclopenta[2,1-b:3,4-b′]dithiophene-2,6-dicarbaldehyde 2 [19] and 9-(2-ethylhexyl)-6-(4,4,5,5-tetramethyl-1,3,2-dioxaborolan-2-yl)-2,3,4,4a,9,9a-hexahydro-1H-carbazole 3 [17] upon prolonged reflux in anhydrous THF and catalysis with tetrakis(triphenylphosphine)palladium (Pd(PPh3)4) led to 4,4-bis(2-ethylhexyl)-6-(9-(2-ethylhexyl)-2,3,4,4a,9,9a-hexahydro-1H-carbazol-6-yl)-4H-cyclopenta[2,1-b:3,4-b′]dithiophene-2-carbaldehyde 1 in a high yield (Scheme 1).
We have measured the optical absorption spectra of compound 1 in dichloromethane and hexane solutions. It was found that, regardless of the polarity of the solvent used, two broad absorption bands were observed in the absorption spectra, one of which, located in the short-wavelength region at 250–310 nm, refers to π–π* transitions of the conjugated system, and the long-wavelength band in the range 400–520 nm due to the presence of intramolecular charge transfer (ICT) [20,21]. As expected, the ICT absorption maximum depended on the polarity of the solvent used: in n-hexane solution, it had a red shift relative to this value recorded in a dichloromethane solution (460 nm for CH2Cl2 and 430 nm for hexane), which indicates the presence of a solvatochromic effect.
The structure of 4,4-bis(2-ethylhexyl)-6-(9-(2-ethylhexyl)-2,3,4,4a,9,9a-hexahydro-1H-carbazol-6-yl)-4H-cyclopenta[2,1-b:3,4-b′]dithiophene-2-carbaldehyde 1 was confirmed by means of high-resolution mass-spectrometry, 1H, 13C NMR, IR, and UV spectroscopy.

3. Materials and Methods

4,4-Bis(2-ethylhexyl)-4H-cyclopenta[2,1-b:3,4-b′]dithiophene-2,6-dicarbaldehyde 2 [19] and 9-(2-ethylhexyl)-6-(4,4,5,5-tetramethyl-1,3,2-dioxaborolan-2-yl)-2,3,4,4a,9,9a-hexahydro-1H-carbazole 3 [17] were prepared according to the published methods. The solvents and reagents were purchased from commercial sources and used as received. Melting point was determined on a Kofler hot-stage apparatus and is uncorrected. 1H and 13C NMR spectra were taken with a Bruker AM-300 machine (Bruker AXS Handheld Inc., Kennewick, WA, USA) (at frequencies of 300 and 75 MHz) in CDCl3 solution, with TMS as the standard. J values are given in Hz. IR spectrum was measured with a Bruker “Alpha-T” instrument in a KBr pellet. High-resolution MS spectrum was measured on a Bruker micrOTOF II instrument (Bruker Daltonik Gmbh, Bremen, Germany) using electrospray ionization (ESI). Solution UV-visible absorption spectra were recorded using a OKB Spektr SF-2000 UV/Vis/NIR spectrophotometer controlled with SF-2000 software. The sample was measured in the standard 10 mm photometric quartz cells in HPLC grade CH2Cl2 and C6H14 in a concentration of 5 × 10−6 M.
Synthesis of 4,4-bis(2-ethylhexyl)-6-(9-(2-ethylhexyl)-2,3,4,4a,9,9a-hexahydro-1H-carbazol-6-yl)-4H-cyclopenta[2,1-b:3,4-b′]dithiophene-2-carbaldehyde 1 (Supplementary Materials).
Aldehyde 2 (390 mg, 0.76 mmol) and boronic ester 3 (377 mg, 0.92 mmol) were dissolved in anhydrous THF (20 mL), and 2 M K2CO3 (10 mL) was added. The mixture was degassed for 20 min with a stream of argon, and Pd(PPh3)4 (50 mg, 40 μmol, 5%) was added in one portion. After refluxing for 10 h in inert atmosphere, the mixture was poured into water (50 mL) and extracted with EtOAc (3 × 20 mL). The combined organic layers were washed with brine, dried over MgSO4, filtered, and concentrated under reduced pressure. The crude product was purified by column chromatography on silica gel (Silica gel Merck 60, eluent ethyl acetate /hexane, 1:25, v/v). Yield 400 mg (73%), orange oil, Rf = 0.6 (hexane/ethyl acetate, 10:1, v/v). IR spectrum, ν, cm–1: 2957, 2927, 2855, 1654, 1612, 1481, 1385, 1315, 1224, 1138. 1H NMR (ppm): δ 9.72 (s, 1H), 7.45 (s 1H), 7.29–7.23 (m, 1H), 7.18 (s, 1H), 6.95 (s, 1H), 6.35 (d, J = 8.1, 1H), 3.48 (m, 1H), 3.07 (m, 1H), 2.94–2.75 (m, 2H), 1.92–1.81 (m, 4H), 1.77–1.68 (m, 2H), 1.66–1.58 (m, 2H), 1.53–1.44 (m, 2H), 1.43–1.34 (m, 3H), 1.27 (s, 8H), 0.98–0.80 (m, 22H), 0.72–0.62 (m, 8H), 0.52–0.57 (m, 6H). 13C NMR (ppm): δ 182.1, 163.4, 156.4, 152.9, 151.6, 149.2, 141.9, 134.2, 132.6, 130.6, 125.4, 123.6, 120.3, 116.0, 106.6, 64.9, 64.6, 53.9, 50.3, 49.9, 43.2, 40.4, 38.9, 38.7, 35.3, 34.2, 34.1, 31.4, 31.1, 29.1, 28.9, 28.6, 27.5, 27.3, 27.1, 25.4, 25.3, 24.7, 24.3, 23.2, 22.8, 21.6, 14.0, 11.1, 10.8, 10.6. HRMS (ESI-TOF), m/z: calcd for C46H67OS2N [M]+, 713.4659, found, 713.4669. UV-Vis spectrum, λmax: 460 nm (ε = 29,160 M−1 cm−1, CH2Cl2), 430 nm (ε = 29,020 M−1 cm−1, C6H14).

Supplementary Materials

The following are available online: copies of 1H, 13C NMR, IR, UV/Vis, and HRMS-spectra for the compound 1.

Author Contributions

Conceptualization, E.A.K.; methodology, O.A.R.; software, E.A.K.; validation, O.A.R.; formal analysis, investigation, M.S.M. and O.O.U.; resources, O.A.R.; data curation, M.S.M.; writing—original draft preparation, M.S.M.; writing—review and editing, E.A.K.; visualization, O.O.U.; supervision, O.A.R.; project administration, O.A.R.; funding acquisition, O.A.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Russian Science Foundation, grant number 22-73-00102.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Riede, M.; Spoltore, D.; Leo, K. Organic Solar Cells—The Path to Commercial Success. Adv. Energy Mater. 2021, 11, 2002653. [Google Scholar] [CrossRef]
  2. Roncali, J. Molecular Engineering of the Band Gap of π-Conjugated Systems: Facing Technological Applications. Macromol. Rapid Commun. 2007, 28, 1761–1775. [Google Scholar] [CrossRef]
  3. Chen, J.; Cao, Y. Development of Novel Conjugated Donor Polymers for High-Efficiency Bulk-Heterojunction Photovoltaic Devices. Acc. Chem. Res. 2009, 42, 1709–1718. [Google Scholar] [CrossRef] [PubMed]
  4. Roncali, J.; Leriche, P.; Blanchard, P. Molecular Materials for Organic Photovoltaics: Small is Beautiful. Adv. Mater. 2014, 26, 3821–3838. [Google Scholar] [CrossRef] [Green Version]
  5. Collins, S.D.; Ran, N.A.; Heiber, M.C.; Nguyen, T.-Q. Small is Powerful: Recent Progress in Solution-Processed Small Molecule Solar Cells. Adv. Energy Mater. 2017, 7, 1602242. [Google Scholar] [CrossRef]
  6. Lin, Y.; Wang, J.; Zhang, Z.-G.; Bai, H.; Li, Y.; Zhu, D.; Zhan, X. An Electron Acceptor Challenging Fullerenes for Efficient Polymer Solar Cells. Adv. Mater. 2015, 27, 1170–1174. [Google Scholar] [CrossRef]
  7. Zhang, G.; Zhao, J.; Chow, P.C.Y.; Jiang, K.; Zhang, J.; Zhu, Z.; Zhang, J.; Huang, F.; Yan, H. Nonfullerene Acceptor Molecules for Bulk Heterojunction Organic Solar Cells. Chem. Rev. 2018, 118, 3447–3507. [Google Scholar] [CrossRef]
  8. Cui, Y.; Xu, Y.; Yao, H.; Bi, P.; Hong, L.; Zhang, J.; Zu, Y.; Zhang, T.; Qin, J.; Ren, J.; et al. Single-Junction Organic Photovoltaic Cell with 19% Efficiency. Adv. Mater. 2021, 33, 2102420. [Google Scholar] [CrossRef]
  9. Po, R.; Bianchi, G.; Carbonera, C.; Pellegrino, A. “All That Glisters Is Not Gold”: An Analysis of the Synthetic Complexity of Efficient Polymer Donors for Polymer Solar Cells. Macromolecules 2015, 48, 453–461. [Google Scholar] [CrossRef]
  10. Knyazeva, E.A.; Rakitin, O.A. Influence of structural factors on the photovoltaic properties of dye-sensitized solar cells. Russ. Chem. Rev. 2016, 85, 1146–1183. [Google Scholar] [CrossRef]
  11. Nakayama, K.; Okura, T.; Okuda, Y.; Matsui, J.; Masuhara, A.; Yoshida, T.; White, M.S.; Yumusak, C.; Stadler, P.; Scharber, M.; et al. Single-Component Organic Solar Cells Based on Intramolecular Charge Transfer Photoabsorption. Materials 2021, 14, 1200. [Google Scholar] [CrossRef] [PubMed]
  12. Terenti, N.; Giurgi, G.-I.; Crişan, A.P.; Anghel, C.; Bogdan, A.; Pop, A.; Stroia, I.; Terec, A.; Szolga, L.; Grosu, I.; et al. Structure–properties of small donor–acceptor molecules for homojunction single-material organic solar cells. J. Mater. Chem. C 2022, 10, 5716–5726. [Google Scholar] [CrossRef]
  13. Slodek, A.; Zych, D.; Kotowicz, S.; Szafraniec-Gorol, G.; Zimosz, S.; Schab-Balcerzak, E.; Siwy, M.; Grzelak, J.; Maćkowski, S. “Small in size but mighty in force”—The first principle study of the impact of A/D units in A/D-phenyl-π-phenothiazine-π-dicyanovinyl systems on photophysical and optoelectronic properties. Dyes Pigments 2021, 189, 109248. [Google Scholar] [CrossRef]
  14. Bu, L.; Rémond, M.; Colinet, P.; Jeanneau, E.; Le Bahers, T.; Chaput, F.; Andraud, C.; Bretonnière, Y. Sensitive 1,1-dicyanovinyl push-pull dye for primary amine sensing in solution by fluorescence. Dyes Pigments 2022, 202, 110258. [Google Scholar] [CrossRef]
  15. Almenningen, D.M.; Hansen, H.E.; Vold, M.F.; Buene, A.F.; Venkatraman, V.; Sunde, S.; Hoff, B.H.; Gautun, O.R. Effect of thiophene-based π-spacers on N-arylphenothiazine dyes for dye-sensitized solar cells. Dyes Pigments 2021, 185, 108951. [Google Scholar] [CrossRef]
  16. Kakiage, K.; Aoyama, Y.; Yano, T.; Oya, K.; Fujisawa, J.; Hanaya, M. Highly-efficient dye-sensitized solar cells with collaborative sensitization by silyl-anchor and carboxy-anchor dyes. Chem. Commun. 2015, 51, 15894–15897. [Google Scholar] [CrossRef] [PubMed]
  17. Tanaka, E.; Mikhailov, M.S.; Gudim, N.S.; Knyazeva, E.A.; Mikhalchenko, L.V.; Robertson, N.; Rakitin, O.A. Structural features of indoline donors in D–A-π-A type organic sensitizers for dye-sensitized solar cells. Mol. Syst. Des. Eng. 2021, 6, 730–738. [Google Scholar] [CrossRef]
  18. Gudim, N.S.; Knyazeva, E.A.; Mikhalchenko, L.V.; Mikhailov, M.S.; Zhang, L.; Robertson, N.; Rakitin, O.A. Novel D-A-π-A1 Type Organic Sensitizers from 4,7-Dibromobenzo[d][1,2,3]thiadiazole and Indoline Donors for Dye-Sensitized Solar Cells. Molecules 2022, 27, 4197. [Google Scholar] [CrossRef]
  19. Mone, M.; Yang, K.; Murto, P.; Zhang, F.; Wang, E. Low-gap zinc porphyrin as an efficient dopant for photomultiplication type photodetectors. Chem. Commun. 2020, 56, 12769–12772. [Google Scholar] [CrossRef]
  20. Budy, S.M.; Suresh, S.; Spraul, B.K.; Smith, D.W. High-Temperature Chromophores and Perfluorocyclobutyl Copolymers for Electro-optic Applications. J. Phys. Chem. C 2008, 112, 8099–8104. [Google Scholar] [CrossRef]
  21. Xia, P.F.; Feng, X.J.; Lu, J.; Tsang, S.; Movileanu, R.; Tao, Y.; Wong, M.S. Donor-Acceptor Oligothiophenes as Low Optical Gap Chromophores for Photovoltaic Applications. Adv. Mater. 2008, 20, 4810–4815. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of 4,4-bis(2-ethylhexyl)-6-(9-(2-ethylhexyl)-2,3,4,4a,9,9a-hexahydro-1H-carbazol-6-yl)-4H-cyclopenta[2,1-b:3,4-b′]dithiophene-2-carbaldehyde 1.
Scheme 1. Synthesis of 4,4-bis(2-ethylhexyl)-6-(9-(2-ethylhexyl)-2,3,4,4a,9,9a-hexahydro-1H-carbazol-6-yl)-4H-cyclopenta[2,1-b:3,4-b′]dithiophene-2-carbaldehyde 1.
Molbank 2022 m1486 sch001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mikhailov, M.S.; Ustimenko, O.O.; Knyazeva, E.A.; Rakitin, O.A. 4,4-Bis(2-ethylhexyl)-6-(9-(2-ethylhexyl)-2,3,4,4a,9,9a-hexahydro-1H-carbazol-6-yl)-4H-cyclopenta[2,1-b:3,4-b′]dithiophene-2-carbaldehyde. Molbank 2022, 2022, M1486. https://doi.org/10.3390/M1486

AMA Style

Mikhailov MS, Ustimenko OO, Knyazeva EA, Rakitin OA. 4,4-Bis(2-ethylhexyl)-6-(9-(2-ethylhexyl)-2,3,4,4a,9,9a-hexahydro-1H-carbazol-6-yl)-4H-cyclopenta[2,1-b:3,4-b′]dithiophene-2-carbaldehyde. Molbank. 2022; 2022(4):M1486. https://doi.org/10.3390/M1486

Chicago/Turabian Style

Mikhailov, Maxim S., Olga O. Ustimenko, Ekaterina A. Knyazeva, and Oleg A. Rakitin. 2022. "4,4-Bis(2-ethylhexyl)-6-(9-(2-ethylhexyl)-2,3,4,4a,9,9a-hexahydro-1H-carbazol-6-yl)-4H-cyclopenta[2,1-b:3,4-b′]dithiophene-2-carbaldehyde" Molbank 2022, no. 4: M1486. https://doi.org/10.3390/M1486

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop