Next Article in Journal
Preparation of Few-Layered Wide Bandgap MoS2 with Nanometer Lateral Dimensions by Applying Laser Irradiation
Next Article in Special Issue
Polyethylenimine-Ethoxylated Interfacial Layer for Efficient Electron Collection in SnO2-Based Inverted Organic Solar Cells
Previous Article in Journal
Ethylammonium Lead Iodide Formation in MAPbI3 Precursor Solutions by DMF Decomposition and Organic Cation Exchange Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Theoretical Study of a Class of Organic D-π-A Dyes for Polymer Solar Cells: Influence of Various π-Spacers

1
Institute for Tropical Technology, Vietnam Academy of Science and Technology (VAST), 18 Hoang Quoc Viet, Hanoi 100000, Vietnam
2
Department of Chemistry and Center for Computational Science, Hanoi National University of Education, 136 Xuan Thuy, Hanoi 100000, Vietnam
3
Department of Chemistry, KU Leuven, B-3001 Leuven, Belgium
4
Institute for Computational Science and Technology (ICST), Ho Chi Minh City 700000, Vietnam
*
Author to whom correspondence should be addressed.
Crystals 2020, 10(3), 163; https://doi.org/10.3390/cryst10030163
Submission received: 14 January 2020 / Revised: 23 February 2020 / Accepted: 27 February 2020 / Published: 2 March 2020
(This article belongs to the Special Issue Organic Photovoltaic)

Abstract

:
A class of D-π-A compounds that can be used as dyes for applications in polymer solar cells has theoretically been designed and studied, on the basis of the dyes recently shown by experiment to have the highest power conversion efficiency (PCE), namely the poly[4,8-bis(5-(2-butylhexylthio)thiophen-2-yl)benzo[1,2-b:4,5-b’]dithiophene-2,6-diyl-alt-TZNT] (PBDTS-TZNT) and poly[4,8-bis(4-fluoro-5-(2-butylhexylthio)thiophen-2-yl)benzo[1,2-b:4,5-b’]dithiophene-2,6-diyl-alt-TZNT] (PBDTSF-TZNT) substances. Electronic structure theory computations were carried out with density functional theory and time-dependent density functional theory methods in conjunction with the 6−311G (d, p) basis set. The PBDTS donor and the TZNT (naphtho[1,2-c:5,6-c]bis(2-octyl-[1,2,3]triazole)) acceptor components were established from the original substances upon replacement of long alkyl groups within the thiophene and azole rings with methyl groups. In particular, the effects of several π-spacers were investigated. The calculated results confirmed that dithieno[3,2-b:2′,3′-d] silole (DTS) acts as an excellent π-linker, even better than the thiophene bridge in the original substances in terms of well-known criteria. Indeed, a PBDTS-DTS-TZNT combination forms a D-π-A substance that has a flatter structure, more rigidity in going from the neutral to the cationic form, and a better conjugation than the original compounds. The highest occupied molecular orbital (HOMO)-lowest unoccupied molecular orbital (LUMO) energy gap of such a D-π-A substance becomes smaller and its absorption spectrum is more intense and red-shifted, which enhances the intramolecular charge transfer and makes it a promising candidate to attain higher PCEs.

1. Introduction

For several decades, polymer solar cells (PSCs) have been the subject of intensive research due to a number of reasons including their easy fabrication, high flexibility and light weight, when compared to other photovoltaic technologies [1,2,3,4,5,6,7,8,9,10,11,12]. Along with their practical applications, great work has been accomplished in developing new active layer materials featuring broad absorption bands, appropriate molecular orbital energy levels and high charge mobilities. Many strategies have been devoted to the optimization of conjugated photovoltaic polymers in order to grasp more efficient PSCs. These include the interface engineering, morphology control of the materials and innovative device architectures, thus fostering their power conversion efficiencies (PCEs) to higher standards [13,14,15,16,17,18,19,20,21,22,23,24,25,26,27]. Clearly, the strong demand for the upgrading of the photovoltaic efficiency of PSCs is one of the main driving forces for the preparation of novel polymers that have excellent photovoltaic properties. Let us briefly describe the recent advances in molecular design strategies of polymeric photovoltaic donor materials used in PSC devices.
In order to meet the criteria for highly efficient polymer donors, research work has been carried out to enhance the intrinsic variables of a conjugated polymer and to design many novel molecular structural motifs. Besides the degree of polymerization and the molecular weight of the polymers, the two major molecular design strategies involve the optimization of parameters related either to their backbones or to their side chains [28,29,30,31,32,33,34,35,36,37,38,39]. Planar and rigid backbones are usually preferred because they exhibit small reorganization energies and a tendency to pack closely in solid films through strong intermolecular interactions and high charge-carrier mobility. Accordingly, promising candidates for fabricating high performance PSCs include low band gap conjugated D-A polymers that contain planar and rigid backbones. An overview of the strategies for backbone design reported in recent literature points out that the introduction of novel building blocks, functional substituents and, especially, π-spacers are the most frequent approaches. A literature analysis also emphasizes that the dithieno[3,2-b:2′,3′-d] thiophene (DTT), dithieno[3,2-b:2′,3′-d] silole (DTS), cyclopenta[2,1-b:3,4-b0] dithiophene (CPDT), dithieno[3,2-b:2′,3′-d]pyrrole (DTP) components have been used as bridges in D-π-A compounds and they lead to relatively good efficiency [40,41,42,43,44,45,46,47,48,49,50,51].
Benzo[1,2-b:4,5-b′] dithiophene (BDT), first introduced into photovoltaic polymers by Hou and co-workers in 2008 [52], has extensively been used as a building block as well as an electron donor unit for conjugated copolymers over the past five years [53]. Optimizations of BDT-based polymers thus provide us with a good strategy for the development of backbones. To alter the band gap and orbital energy levels in BDT-based polymers, different electron acceptor units, such as thiophene, benzothiadiazole (BT), thieno[3,4-b] pyrazine (TPZ), etc., have been explored [52]. The band gaps of these BDT-based polymers are located in the range of 1.1–2.0 eV and their highest occupied molecular orbital (HOMO) (−4.6 to −5.2 eV) and lowest unoccupied molecular orbital (LUMO) (−2.7 to −3.5 eV) energy levels could also be tuned effectively. The absorption edges were also tuned up from 600 to 1100 nm. Previous work provided valuable insights into the band gap and molecular orbital energy level modulation via change of backbone structure in conjugated polymers.
Recently, two novel TZNT-containing wide band gap (WBG) polymers, including the combined motifs PBDTS-TZNT and PBDTSF-TZNT, were successfully designed and synthesized for their use in highly efficient non-fullerence polymer solar cells (NF-PSCs) with low energy loss. The rigid planar backbones of both BDT and TZNT units provided these copolymers with high crystallinity and good molecular packing. Homo-tandem devices based on PBDTSF-TZNT:IT−4F subcells further enhanced the light-harvesting ability and boosted their PCE up to 14.52%, which is currently the best value for homo-tandem NF-PSCs [54]. Overall, any further improvement of the performance appears to depend more on the regulation of the π-conjugation than on the donor and acceptor components. It is also valuable to note that the selection of atoms, such as C, Si and N, can notably impact both the electronic traits of semiconducting polymers and tuning of the performance of organic optoelectronic devices.
In this context, we set out to obtain more insights into the correlation between the electronic properties of the D-π-A material that links closely to the device performance and the effects of their structural aspects such as the π-linkers. The two PBDTS and TZNT units were chosen to be the donor (D) and acceptor (A) components of the D-π-A compounds due to their orbital energy levels, absorption domain, crystallinity, charge carrier mobility and blend morphology that can feasibly be tailored by modifying the two-dimensional (2D) conjugated side chains of the PBDTS and TZNT components [40−52,54]. This would lead to a rational guidance for molecular design and fine-tuning of novel photovoltaic polymers. For this purpose, we employed electronic structure theory computations to predict relevant optoelectronic parameters. The calculated results allow us to propose a design strategy for novel materials with the aim being to achieve a better device performance. We also give some future directions and approaches to develop higher performance donor polymers for photovoltaic applications.

2. Computational Details

All the electronic structure theory computations are based on density functional theory (DFT) and time-dependent DFT (TD-DFT) [55,56]. All calculations are carried out with the Gaussian 09 package [57]. The popular hybrid B3LYP functional is used in combination with the split-valence polarization 6−311G (d, p) basis set [58] to optimize geometrical structures and to calculate the UV-VIS absorption spectra. Geometry optimizations are carried out for the singlet ground state (S0) of all compounds considered in their neutral forms. The open-shell doublet state is considered for the anionic and cationic counterparts. Harmonic vibrational frequencies are subsequently calculated at the same level to confirm that each ground state does not have any imaginary frequency and to evaluate their zero-point energies. In order to take the solvent effect into account, the polarizable continuum model (IEF-PCM) is used. Accordingly, TD-DFT calculations are used to obtain the absorption wavelengths and their oscillator strengths of the studied compounds both in the gas phase and in the solvent.
Based on the electronic structure of neutral geometries at their ground states, their band gaps (Egap) are simply determined from the difference in the energies of the HOMO and LUMO. To probe the electron accepting and donating abilities, the electron affinity (EA) and ionization energy (IE) are evaluated for both vertical and adiabatic processes. The charge transfer, one of the most important properties of semiconductor materials, can be determined by two different theories, namely the Band theory and the Hopping model [59]. In this work, the charge transports are obtained according to the semi-classical Marcus theory and can be written as shown in the following Equation (1) [60,61,62,63,64,65]:
k E T = 4 π 2 h 1 4 π λ k B T V 2 exp { λ 4 k B T }
where kET, h, kB, T, λ and V are the hopping rate, Planck constant, Boltzmann constant, temperature, the reorganization energy and transfer integral, respectively.
According to Equation (1), a lowering of the reorganization energy enhances the charge transfer property by an increase of the hopping rate. For a neutral molecule, the reorganization energy for hole (λh) and electron (λe) are described by Equations (2) and (3) [60,61,62,63,64,65]:
λh = (EC(N) − EC) + (EN(C) − EN)
λe = (EA(N) − EA) + (EN(A) − EN).
in these equations, EC(N) and EA(N) are the energy of cations/anions in the optimized geometry of neutral form, respectively. EN(C)/EN(A) is the energy of neutral molecules computed with the optimized cation/anionic states. EN/EC/EA could be viewed as the energy of neutral molecules/cations/anions in their corresponding optimized geometries.

3. Results and Discussion

3.1. Structural and Optoelectronic Properties

The structural shapes of the compounds considered are shown in Figure 1. The Cartesian coordinates of studied compounds can be found in Table S1. They are the analogs of the PBDTS-TZNT and PBDTSF-TZNT compounds that were found to be the most efficient polymer solar cells to date [54]. In this study, we theoretically designed the new compounds by replacing the thiophene bridge S1 in the original compound by the DTT, DTS, CPDT and DTP bridges in order to obtain the new compounds. To simplify the notations, the designed compounds are denoted as DTT, DTS, CPDT and DTP, respectively, which correspond to the D-π-A compounds having these bridges, as displayed in Figure 1. For computational ease, in this work, we replaced the long alkyl groups attached to the thiophene ring and the N atom in the azole ring by methyl groups. We thus assume that the electronic properties are not significantly affected by the length of alkyl groups.
The r1, r2 parameters are defined as the bond lengths between the bridge and the donor (PBDTS) and acceptor (TZNT) components, whereas the φ1, φ2 parameters are the dihedral angles between the π-linker plane and the planes of PBDTS and TZNT components, respectively. The main structural parameters, including both the bond length r and dihedral angle φ of the original and the modified compounds, are summarized in Table 1.
The computed r1 and r2 values do not change significantly when we compare the gas phase results with the solvated ones (see Table 1). It is interesting to note that the values of r1 and r2for the modified compounds with the bridge being DTS, CPDT, DTP, DTT are almost the same as compared to the original compound S1. These values are in the range of 1.44–1.45 Å, values well in between the lengths of the C–C single bond (1.53 Å) and the C=C double bond (1.34 Å). This indicates that these molecules display an excellent π-delocalization throughout their backbone chains, from the donor (BDT) through the π-spacer to the acceptor (BT). This is expected to enhance the intramolecular charge transfer (ICT), which is related to a red-shift of their absorption spectra [66]. This shall be discussed in Section 3.4.
In most of the compounds designed (Figure 1) the lengths of the C–C bridge bonds r1 are generally shorter than those of the corresponding bonds r2. This can be explained by the additional electron density at the r1 bond gained from the donor component, whereas in the opposite direction, there is a withdrawal of electron density of the r2 bond by the acceptor component BT.
The rather large dihedral angles φ of the original compound S1, which has just a single thiophene ring as a π-linker, suggest that there is a strong steric hindrance between the π-linker and the groups on both sides. When the DTS, CPDT, DTP or DTT units are used as the π-spacer, the dihedral angles φ1 and φ2 of the resulting compounds are changed non-monotonously. However, the φ2 dihedral angle of the investigated compounds changes insignificantly. The φ1 of the compounds in both the gas phase and chlorobenzene solvent increases slightly from 11° and 12° in S1 to 16°and 14° in CPDT; to 20° and 18° in DTT, and decreases slightly to 11° and 11°. Compared to the original compound S1, the φ1 of the DTS component decreases significantly to 3° and 1° both in the gas phase and in chlorobenzene solvent, respectively.
We also see that, for compounds S1, DTT, DTP, CPDT, the values of φ1 are significantly larger than those of φ2, indicating that the above bridges enjoy a better conjugation with the acceptor rather than with the donor component. Meanwhile, for DTS component, these values are not significantly different from each other, indicating that DTS bridge possesses the best conjugation with both donor and acceptor components.
These results emphasize that the DTS is the most co-planar π-spacer with the components in both sides, involving a strong conjugation between the spacer and the other D and A components. We would expect that the more co-planar structure in the ground electronic state is, the faster the photo-induced electron transfer from the ground electronic state to the first singlet excited state is. Therefore, a compound containing a DTS π-linker connecting the components on both sides is thus expected to behave as better transport materials.
It can be seen in Table 2 that, in going from the neutral state to the ionic state, all the modified compounds have smaller variations in the values of the bridge lengths r1 and r2 than in the original one. In detail, the bridge bond lengths in the parent compound change by 0.052 Å and 0.044 Å when switching from the neutral molecule to the anion and cation, respectively, whereas these values for the modified compounds vary between 0.045–0.050 Å and 0.037–0.040 Å, respectively. For the values of φ1 and φ2, when converting from the neutral to the ionic state, the DTS component has the smallest variations for these values, two and three degrees respectively, being the smallest changes as compared to other compounds.
The lengthening of the inter-ring bond can be understood by an anti-bonding interaction between the π orbitals in both rings. Hence, the loss of an electron from the HOMO upon ionization leads to a shortening of the bridge in the cationic state in comparison to that in the neutral state. A weak relaxation in geometrical parameters upon removal/addition of electrons is also expected to reduce the reorganization energy, which, in turn, would increase the hopping rate [67].
The lengthening of the inter-ring bond can be understood by an anti-bonding interaction between the π orbitals in both rings. Hence, the loss of an electron from the HOMO leads to a shortening of the bridge in the cationic state in comparison to that in the neutral state.

3.2. Frontier Molecular Orbitals

Calculated results show that the rigidification of dithiophene linkers can noticeably tune the orbital energy level alignments and orbital distribution.
Frontier molecular orbitals (FMOs) play an important role in the electrical properties since they usually govern the charge carrier transport nature of molecular systems [9]. The HOMOs and LUMOs are plotted in Figure 2. Both orbitals are distributed over the whole skeleton of the conjugated molecules, indicating that there is a strong spatial overlap between the π electrons, and this likely results in a stronger optical absorption corresponding to an electronic transition from the HOMO to the LUMO to generate the first excited state S1 [14]. A good delocalization in both frontier orbitals is favorable for enhancing the hole and electron transfer integrals of a transport material. In each system, the parent compound always has the larger negative value of HOMO and smaller negative values of LUMO as compared to its derivatives. As a consequence, the substituents tend to improve the transport properties.
The calculated HOMO and LUMO energy level of the studied compounds in the gas phase and in the solvent are listed in Table 3. The HOMOs and LUMOs levels of the modified compounds are found to be in between −5.0 eV and −5.2 eV and −2.5 eV and −2.6 eV, respectively, in the gas phase, and in between −5.1 eV and −5.3 eV and −2.6 eV and −2.7 eV, respectively in the chlorobenzene solvent. The HOMO levels of all modified compounds are lower than those of the original compound S1 (being −5.3 eV in the gaseous state and −5.4 eV in the solvent), which was proven to be a good hole transport material (HTM). This finding suggests that the designed compounds could behave even better as HTMs than the original compound S1.
It is helpful to state again that the energies of HOMO and LUMO can be used to characterize the hole and electron injection of material. A hole transport material having small negative HOMO energy can lose its electrons more easily (low IE). On the contrary, an electron transport material (ETM) needs to possess a large LUMO energy, which accepts electrons more easily (large EA). The HOMO and LUMO levels themselves illustrate the charge transfer interactions occurring within a molecule in which a small Egap could facilitate the interaction. The Egap values of the molecules studied spread out over a wide range from 2.5 eV for CPDT to 2.7 eV for the original compound S1. A comparison of Egap among the studied compounds points out an increasing ordering of Egap as follows: CPDT < DTP < DTS < DTT < S1.
The calculated results show that while the LUMO energy levels of the compounds remain almost unchanged, the corresponding HOMO energy levels become less negative (lower IEa), which reduces the HOMO-LUMO energy gap and thereby facilitates their electron releases.

3.3. Electronic Properties

As mentioned in Section 2 above, in order to probe the electron accepting and donating abilities, we calculate the vertical electron affinity (EAv) and vertical ionization energy (IEv) which can widely be measured by the photoelectron spectroscopy (PES) method, along with the adiabatic ones. In addition, we also calculate these indexes in adiabatic processes (IEa and EAa, respectively).
The calculated results listed in Table 4 show that the EAv has no significant change due to the minor fluctuation of their LUMOs, as stated above, whereas the IEv significantly changes due to the significant shift in their HOMOs. All modified compounds not only have smaller IEv values but also larger EAv values, indicating that the substituted molecules can be considered good candidates for better transport materials in comparison with compound S1. Moreover, the reorganization energies for both the electrons and holes of the modified compounds are small and nearly equal to each other, reflecting that these compounds exhibit improved characteristics of a balanced transformation of both the holes and electrons within each molecule.

3.4. Absorption Spectra

The simulated absorption spectra are illustrated in Figure 3. The calculated wavelengths (in nm) and oscillator strengths f and the corresponding transition assignments of the compounds listed in Table 5 show that the modified compounds are all red-shifted as compared with the original S1 by 63 nm, 59 nm, 53 nm and 20 nm for CPDT, DTS, DTP and DTT, respectively. This is also consistent with the analysis in Section 3.1 above. The absorption wavelengths of the modified compounds are significantly higher than that of the parent compound. The main electronic transition arises in all cases from HOMO →LUMO. In addition, the computed absorption wavelength of compound S1 (525 nm) is in good agreement with the experimental result (537 nm).

4. Concluding Remarks

We have designed D-π-A compounds with the aim of their application in polymer solar cells attaining a high power conversion efficiency. Based on the compounds recently shown to have the highest PCE to date, the new class was designed to achieve better performance. Relevant optoelectronic properties were determined with density functional theory and time-dependent density functional theory calculations. An outstanding aim of the present work was an investigation of the effects of different π-spacers to find the bridges better than the thiophene ring connecting both donor D and acceptor A components. Indeed, calculated results suggest that the DTS unit acts as the best π-linker and much better than the thiophene bridge in the original substance, in terms of the planarity and optoelectronic parameters. Its presence in the designed PBDTS-DTS-TZNT compound induces a flatter structure, shorter r1 and r2 bond distances and much better electron conjugation than the original compounds. Its HOMO-LUMO band gap becomes smaller and its absorption spectrum is more intense and red-shifted. We propose the use of this compound for an actual experimental preparation, followed by a test of its application.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4352/10/3/163/s1. Table S1: Cartesian coordinates (Å) for studied compound, as obtained using B3LYP/6-311G (d, p).

Author Contributions

All authors contributed significantly to the conceptualization and realisation of the manuscript. Methodology, T.N.D. and L.T.H.H.; The writing of the first draft was mainly done by N.V.T. and N.T.C.; Writing—review and editing, M.T.N. and D.E.; Supervision, H.M.T.N. and revision was done by her with help of the other authors. All authors have read and agree to the published version of the manuscript.

Funding

This research was funded by the program ‘Fostering Innovation through Research, Science and Technology (FIRST) of Vietnam under grant number 12/FIRST/1.a/HNUE and N.V.T. thanks the project VAST.CTG.01/17-19 from Vietnam Academy of Science and Technology (VAST).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liu, F.; Gu, Y.; Shen, X.; Ferdous, S.; Wang, H.-W.; Russell, T.P. Characterization of the Morphology of Solution-Processed Bulk Heterojunction Organic Photovoltaics. Prog. Polym. Sci. 2013, 38, 1990. [Google Scholar] [CrossRef]
  2. Chen, L.-M.; Hong, Z.; Li, G.; Yang, Y. Recent Progress in Polymer Solar Cells: Manipulation of Polymer: Fullerene Morphology and the Formation of Efficient Inverted Polymer Solar Cells. Adv. Mater. 2009, 21, 1434. [Google Scholar] [CrossRef]
  3. Ye, L.; Jing, Y.; Guo, X.; Sun, H.; Zhang, S.; Huo, L.; Hou, J. Remove the Residual Additives toward Enhanced Efficiency with Higher Reproducibility in Polymer Solar Cells. J. Phys. Chem. C 2013, 117, 14920. [Google Scholar] [CrossRef]
  4. Ye, L.; Zhang, S.; Ma, W.; Fan, B.; Guo, X.; Huang, Y.; Ade, H.; Hou, J. From binary to ternary solvent: Morphology fine-tuning of D/A blends in PDPP3T-based polymer solar cells. Adv. Mater. 2012, 24, 6335. [Google Scholar] [CrossRef] [PubMed]
  5. Yu, G.; Gao, J.; Hummelen, J.C.; Wudl, F.; Heeger, A.J. Polymer Photovoltaic Cells: Enhanced Efficiencies via a Network of Internal Donor-Acceptor Heterojunctions. Science 1995, 270, 1789. [Google Scholar] [CrossRef] [Green Version]
  6. Heeger, A.J. 25th anniversary article: Bulk heterojunction solar cells: Understanding the mechanism of operation. Adv. Mater. 2014, 26, 10. [Google Scholar] [CrossRef]
  7. Cheng, Y.J.; Yang, S.H.; Hsu, C.S. Synthesis of Conjugated Polymers for Organic Solar Cell Applications. Chem. Rev. 2009, 109, 5868. [Google Scholar] [CrossRef]
  8. Chen, J.; Cao, Y. Development of Novel Conjugated Donor Polymers for High-Efficiency Bulk-Heterojunction Photovoltaic Devices. Acc. Chem. Res. 2009, 42, 1709. [Google Scholar] [CrossRef]
  9. Zhan, X.; Zhu, D. Conjugated polymers for high-efficiency organic photovoltaics. Polym. Chem. 2010, 1, 409. [Google Scholar] [CrossRef]
  10. Beaujuge, P.M.; Fréchet, J.M.J. Molecular Design and Ordering Effects in π-Functional Materials for Transistor and Solar Cell Applications. J. Am. Chem. Soc. 2011, 133, 20009. [Google Scholar] [CrossRef]
  11. He, F.; Yu, L. How Far Can Polymer Solar Cells Go? In Need of a Synergistic Approach. J. Phys. Chem. Lett. 2011, 2, 3102. [Google Scholar] [CrossRef]
  12. Boudreault, P.-L.T.; Najari, A.; Leclere, M. Processable Low-Bandgap Polymers for Photovoltaic Applications. Chem. Mater. 2011, 23, 456. [Google Scholar] [CrossRef]
  13. Spanggaard, H.; Krebs, F.C. A brief history of the development of organic and polymeric photovoltaics. Sol. Energy Mater. Sol. Cells 2004, 83, 125. [Google Scholar] [CrossRef]
  14. Li, Y.; Zou, Y. Conjugated Polymer Photovoltaic Materials with Broad Absorption Band and High Charge Carrier Mobility. Adv. Mater. 2008, 20, 2952. [Google Scholar] [CrossRef]
  15. Shaheen, S.E.; Brabec, C.J.; Sariciftci, N.S.; Padinger, F.; Fromherz, T.; Hummelen, J.C. 2.5% efficient organic plastic solar cells. Appl. Phys. Lett. 2001, 78, 841. [Google Scholar] [CrossRef] [Green Version]
  16. Zhou, Q.; Zheng, L.; Sun, D.; Deng, X.; Yu, G.; Cao, Y. Efficient polymer photovoltaic devices based on blend of MEH-PPV and C60 derivatives. Synth. Met. 2003, 135–136, 825–826. [Google Scholar] [CrossRef]
  17. Dang, M.T.; Hirsch, L.; Wantz, G. P3HT:PCBM, Best Seller in Polymer Photovoltaic Research. Adv. Mater. 2011, 23, 3597. [Google Scholar] [CrossRef]
  18. Li, G.; Shrotriya, V.; Huang, J.; Yao, Y.; Moriarty, T.; Emery, K.; Yang, Y. High-efficiency solution processable polymer photovoltaic cells by self-organization of polymer blends. Nat. Mater. 2005, 4, 864–868. [Google Scholar] [CrossRef]
  19. Yamaguchi, S.; Goto, T.; Tamao, K. Silole-Thiophene Alternating Copolymers with Narrow Band Gaps. Angew. Chem. Int. Ed. 2000, 39, 1695–1697. [Google Scholar] [CrossRef]
  20. Murata, H.; Kafafi, Z.H.; Uchida, M. Efficient organic light-emitting diodes with undoped active layers based on silole derivative. Appl. Phys. Lett. 2002, 80, 189–191. [Google Scholar] [CrossRef] [Green Version]
  21. Chen, H.Y.; Lam, W.Y.; Luo, J.D.; Ho, Y.L.; Tang, B.Z.; Zhu, D.B.; Wong, M.; Kwok, H.S. Highly efficient organic light-emitting diodes with a silole-based compound. Appl. Phys. Lett. 2002, 81, 574–575. [Google Scholar] [CrossRef] [Green Version]
  22. Liu, J.; Wang, K.; Xu, F.; Tang, Z.; Zheng, W.; Zhang, J.; Li, C.; Yu, T.; You, X. Synthesis and photovoltaic performances of donor– p–acceptor dyes utilizing 1,3,5-triazine as π-spacers. Tetrahedron Lett. 2011, 52, 6492–6496. [Google Scholar] [CrossRef]
  23. Liu, T.; Meng, D.; Cai, Y.; Sun, X.; Li, Y.; Huo, L.; Liu, F.; Wang, Z.; Russell, T.P.; Sun, Y. High-Performance Non-Fullerene Organic Solar Cells Based on a Selenium-Containing Polymer Donor and a Twisted Perylene Bisimide Acceptor. Adv. Sci. 2016, 3, 1600117. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Sandberg, H.G.O.; Frey, G.L.; Shkunov, M.N.; Sirringhaus, H.; Friend, R.H.; Nielsen, M.M.; Kumpf, C. Ultrathin Regioregular Poly(3-hexyl thiophene) Field-Effect Transistors. Langmuir 2002, 18, 10176–10182. [Google Scholar] [CrossRef]
  25. Shuto, A.; Kushida, T.; Fukushima, T.; Kaji, H.; Yamaguchi, S. π Extended Planarized Triphenylboranes with Thiophene Spacers. Org. Lett. 2013, 15, 6234–6237. [Google Scholar] [CrossRef]
  26. Yang, H.; LeFevre, S.W.; Ryu, C.Y.; Bao, Z. Solubility-driven thin film structures of regioregular poly (3-hexyl thiophene) using volatile solvents. Appl. Phys. Lett. 2007, 90, 172116. [Google Scholar] [CrossRef]
  27. Yao, H.; Yu, R.; Shin, T.J.; Zhang, H.; Zhang, S.; Jang, B.; Uddin, M.A.; Woo, H.Y.; Hou, J. A Wide Bandgap Polymer with Strong π–π Interaction for Efficient Fullerene-Free Polymer Solar Cells. Adv. Energy Mater. 2016, 6, 1600742. [Google Scholar] [CrossRef]
  28. Li, Y. Molecular Design of Photovoltaic Materials for Polymer Solar Cells: Toward Suitable Electronic Energy Levels and Broad Absorption. Acc. Chem. Res. 2012, 45, 723. [Google Scholar] [CrossRef]
  29. Zhou, H.; Yang, L.; You, W. Rational Design of High Performance Conjugated Polymers for Organic Solar Cells. Macromolecules 2012, 45, 607. [Google Scholar] [CrossRef] [Green Version]
  30. Henson, Z.B.; Mullen, K.; Bazan, G.C. Design strategies for organic semiconductors beyond the molecular formula. Nat. Chem. 2012, 4, 699. [Google Scholar] [CrossRef]
  31. Yu, R.L.; Price, S.C.; You, W. Structure-property optimizations in donor polymers via electronics, substituents, and side chains toward high effiency solar cells. Macromol. Rapid Commun. 2012, 33, 1162. [Google Scholar]
  32. Dang, M.T.; Hirsch, L.; Wantz, G.; Wuest, J.D. Controlling the morphology and performance of bulk heterojunctions in solar cells. Lessons learned from the benchmark poly(3-hexylthiophene):[6,6]-phenyl-C61-butyric acid methyl ester system. Chem. Rev. 2013, 113, 3734–3765. [Google Scholar] [CrossRef] [PubMed]
  33. Ohshita, J.; Kai, H.; Takata, A.; Iida, T.; Kunai, A.; Ohta, N.; Komaguchi, K.; Shiotani, M.; Adachi, A.; Sakamaki, K.; et al. Effects of Conjugated Substituents on the Optical, Electrochemical, and Electron-Transporting Properties of Dithienosiloles. Organometallics 2001, 20, 4800−4805. [Google Scholar] [CrossRef]
  34. Bonnier, C.; Machin, D.D.; Abdi, O.; Koivisto, B.D. Manipulating non-innocent π -spacers: The challenges of using 2,6-disubstituted BODIPY cores within donor– acceptor light-harvesting motifs. Org. Biomol. Chem. 2013, 11, 3756. [Google Scholar] [CrossRef] [Green Version]
  35. Goh, C.; Kline, R.J.; McGehee, M.D.; Kadnikova, E.N.; Fréchet, J.M. Molecular-weight-dependent mobilities in regioregular poly(3-hexyl-thiophene) diodes. J. Appl. Phys. Lett. 2005, 86, 122110. [Google Scholar] [CrossRef] [Green Version]
  36. Hua, Y.; He, J.; Zhang, C.; Qin, C.; Han, L.; Zhao, J.; Chen, T.; Wong, W.-Y.; Wong, W.-K.; Zhu, X. Effects of Various π-Conjugated Spacers in Thiadiazole[3,4-c]pyridine-Cored Panchromatic Organic Dyes for Dye-Sensitized Solar Cells. J. Mater. Chem. A 2015, 3, 3103–3112. [Google Scholar] [CrossRef]
  37. Tan, D.; Wan, J.; Xu, X.; Lee, Y.W.; Woo, H.Y.; Fenga, K.; Peng, Q. Naphthobistriazole-Based Wide Bandgap Donor Polymers for Efficient Non-Fullerene Organic Solar Cells: Significant Fine-Tuning Absorption and Energy Level by Backbone Fluorination. Nano Energy 2018, 53, 258–269. [Google Scholar]
  38. Xu, X.; Zhang, G.; Li, Y.; Peng, Q. The recent progress of wide bandgap donor polymers towards non-fullerene organic solar cells. Chin. Chem. Lett. 2019, 30, 809–825. [Google Scholar] [CrossRef]
  39. Huang, F.; Yip, H.-L.; Cao, Y. Polymer Photovoltaics: Materials, Physics, and Device Engineering, RSC Polymer Chemistry Series No. 17; Royal Society of Chemistry: London, UK, 2016. [Google Scholar]
  40. Bin, H.J.; Gao, L.; Zhang, Z.G.; Yang, Y.K.; Zhang, Y.D.; Zhang, C.F.; Chen, S.S.; Xue, L.W.; Yang, C.; Xiao, M.; et al. 11.4% Efficiency non-fullerene polymer solar cells with trialkylsilyl substituted 2D-conjugated polymer as donor. Nat. Commun. 2016, 7, 13651. [Google Scholar] [CrossRef] [Green Version]
  41. Bin, H.J.; Zhang, Z.G.; Gao, L.; Chen, S.S.; Zhong, L.; Xue, L.W.; Yang, C.; Li, Y.F. Nonfullerene polymer solar cells based on alkylthio and fluorine substituted 2D-conjugated polymers reach 9.5% efficiency. J. Am. Chem. Soc. 2016, 138, 4657–4664. [Google Scholar] [CrossRef]
  42. Fan, Q.P.; Su, W.Y.; Meng, X.Y.; Guo, X.; Li, G.D.; Ma, W.; Zhang, M.J.; Li, Y.F. High performance non-fullerene polymer solar cells based on fluorine substituted wide bandgap copolymers without extra treatments. Solar RRL 2017, 1, 1700020. [Google Scholar] [CrossRef]
  43. Fan, Q.P.; Su, W.Y.; Wang, Y.; Guo, B.; Jiang, Y.F.; Guo, X.; Liu, F.; Russell, T.P.; Zhang, M.J.; Li, Y.F. Synergistic effect of fluorination on both donor and acceptor materials for high performance non-fullerene polymer solar cells with 13.5% efficiency. Sci. Chin. Chem. 2018, 61, 531–537. [Google Scholar] [CrossRef]
  44. Fei, Z.P.; Eisner, F.D.; Jiao, X.C.; Azzouzi, M.; Rohr, J.A.; Han, Y.; Shahid, M.; Chesman, A.S.R.; Easton, C.D.; McNeill, C.R.; et al. An alkylated indacenodithieno[3,2-b]thiophene based nonfullerene acceptor with high crystallinity exhibiting single junction solar cell efficiencies greater than 13% with low voltage losses. Adv. Mater. 2018, 30, 1800728. [Google Scholar] [CrossRef] [PubMed]
  45. Tang, A.L.; Xiao, B.; Chen, F.; Zhang, J.Q.; Wei, Z.X.; Zhou, E.J. The introduction of fluorine and sulfur atoms into benzotriazole based p-type polymers to match with a benzotriazole-containing n-type small molecule: ‘‘The same-acceptor-strategy’’ to realize high open-circuit voltage. Adv. Energy Mater. 2018, 8, 1801582. [Google Scholar] [CrossRef]
  46. Xu, X.P.; Bi, Z.Z.; Ma, W.; Wang, Z.S.; Choy, W.C.H.; Wu, W.L.; Zhang, G.J.; Li, Y.; Peng, Q. Highly efficient ternary-blend polymer solar cells enabled by a nonfullerene acceptor and two polymer donors with a broad composition tolerance. Adv. Mater. 2017, 29, 1704271. [Google Scholar] [CrossRef]
  47. Xu, X.P.; Li, Z.J.; Bi, Z.Z.; Yu, T.; Ma, W.; Feng, K.; Li, Y.; Peng, Q. Highly efficient nonfullerene polymer solar cells enabled by a copper(I) coordination strategy employing a 1,3,4-oxadiazole-containing wide bandgap copolymer donor. Adv. Mater. 2018, 30, 1800737. [Google Scholar] [CrossRef]
  48. Ye, L.; Zhang, S.Q.; Huo, L.J.; Zhang, M.J.; Hou, J.H. Realizing over 10% efficiency in polymer solar cell by device optimization. Acc. Chem. Res. 2014, 47, 1595–1603. [Google Scholar] [CrossRef]
  49. Zhang, G.J.; Xu, X.P.; Bi, Z.Z.; Ma, W.; Tang, D.S.; Li, Y.; Peng, Q. Fluorinated and Alkylthiolated Polymeric Donors Enable both Efficient Fullerene and Nonfullerene Polymer Solar Cells. Adv. Funct. Mater. 2018, 28, 1706404. [Google Scholar] [CrossRef]
  50. Zhang, H.; Yao, H.F.; Hou, J.X.; Zhu, J.; Zhang, J.Q.; Li, W.N.; Yu, R.N.; Gao, B.W.; Zhang, S.Q.; Hou, J.H. Over 14% Efficiency in Organic Solar Cells Enabled by Chlorinated Nonfullerene Small-Molecule Acceptors. Adv. Mater. 2018, 30, 1800613. [Google Scholar] [CrossRef]
  51. Zhao, W.C.; Li, S.S.; Yao, H.F.; Zhang, S.Q.; Zhang, Y.; Yang, B.; Hou, J.H. Molecular optimization enables over 13% efficiency in organic solar cells. J. Am. Chem. Soc. 2017, 139, 7148–7151. [Google Scholar] [CrossRef]
  52. Hou, J.; Park, M.-H.; Zhang, S.; Yao, Y.; Chen, L.-M.; Li, J.-H.; Yang, Y. Bandgap and Molecular Energy Level Control of Conjugated Polymer Photovoltaic Materials Based on Benzo[1,2-b:4,5-b′]dithiophene. Macromoleculaes 2008, 41, 6012–6018. [Google Scholar] [CrossRef] [Green Version]
  53. Hou, L.; Hou, J. Benzo[1,2-b:4,5-b′]dithiophene-based conjugated polymers: Band gap and energy level control and their application in polymer solar cells. Polym. Chem. 2011, 2, 2453. [Google Scholar]
  54. Feng, K.; Yuan, J.; Bi, Z.; Ma, W.; Xu, X.; Zhang, G.; Peng, Q. Low-Energy-Loss Polymer Solar Cells with 14.52% Efficiency Enabled by Wide-Band-Gap Copolymers. iScience 2019, 12, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Hohenberg, P.; Kohn, W. Inhomogeneous Electron Gas. Phys. Rev. 1964, 136, B864–B871. [Google Scholar] [CrossRef] [Green Version]
  56. Bauernschmitt, R.; Ahlrichs, R. Treatment of electronic excitations within the adiabatic approximation of time dependent density functional theory. Chem. Phys. Lett. 1996, 256, 454–464. [Google Scholar] [CrossRef]
  57. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 09, Revision A.02; Gaussian Inc.: Wallingford, UK, 2016. [Google Scholar]
  58. Ditchfield, R.; Hehre, W.J.; Pople, J.A. Self-Consistent Molecular Orbital Methods. 9. Extended Gaussian-type Basis for Molecular Orbital Studies of Organic Molecules. J. Chem. Phys. 1971, 54, 724–730. [Google Scholar] [CrossRef]
  59. Tang, C.W.; VanSlyke, S.A. Organic electroluminescent diodes. Appl. Phys. Lett. 1987, 51, 913–920. [Google Scholar] [CrossRef]
  60. Tang, C.W.; VanSlyke, S.A.; Chen, C.H. Electroluminescence of doped organic thin films. J. Appl. Phys. 1989, 65, 3610–3620. [Google Scholar] [CrossRef]
  61. Tamao, K.; Yamaguchi, S.; Shiozaki, M.; Nakagawa, Y.; Ito, Y. Thiophene-silole cooligomers and copolimers. J. Am. Chem. Soc. 1992, 114, 5867–5869. [Google Scholar] [CrossRef]
  62. Tamao, K.; Yamaguchi, S.; Ito, Y.; Matsuzaki, Y.; Yamabe, T.; Fukushima, M.; Mori, S. Silole-Containing π-Conjugated Systems. 3.1 A Series of Silole-Thiophene Cooligomers and Copolymers: Synthesis, Properties, and Electronic Structures. Macromolecules 1995, 28, 8668–8675. [Google Scholar] [CrossRef]
  63. Tamao, K.; Ohno, S.; Yamaguchi, S. Silole–pyrrole co-oligomers: Their synthesis, structure and UV-VIS absorption spectra. Chem. Commun. 1996, 1873–1874. [Google Scholar] [CrossRef]
  64. Tamao, K.; Yamaguchi, S. Regio-controlled intramolecular reductive cyclization of diynes. Pure Appl. Chem. 1996, 8, 139–144. [Google Scholar] [CrossRef]
  65. Yamaguchi, S.; Tamao, K. Theoretical Study of the Electronic Structure of 2,2′-Bisilole in Comparison with 1,1′-Bi-1,3-cyclopentadiene: σ*–π* Conjugation and a Low-Lying LUMO as the Origin of the Unusual Optical Properties of 3,3′,4,4′-Tetraphenyl-2,2′-bisilole. Bull. Chem. Soc. Jpn. 1996, 69, 2327–2336. [Google Scholar] [CrossRef]
  66. Kim, D.-H.; Ohshita, J.; Lee, K.-H.; Kunugi, Y.; Kunai, A. Synthesis of p-conjugated oligomers containing dithienosilole units. Organometallics 2006, 25, 1511–1516. [Google Scholar] [CrossRef]
  67. Ohshita, J. Conjugated Oligomers and Polymers Containing Dithienosilole Units. J. Macromol. Chem. Phys. 2009, 210, 1360–1370. [Google Scholar] [CrossRef]
Figure 1. Molecular structures of the compounds designed with rigid π-spacers. S1 is the original compound with the thiophene bridge. DTS, CPDT, DTP and DTT denote the compounds having the corresponding bridges.
Figure 1. Molecular structures of the compounds designed with rigid π-spacers. S1 is the original compound with the thiophene bridge. DTS, CPDT, DTP and DTT denote the compounds having the corresponding bridges.
Crystals 10 00163 g001
Figure 2. Spatial distributions of the frontier molecular orbitals of the compounds designed.
Figure 2. Spatial distributions of the frontier molecular orbitals of the compounds designed.
Crystals 10 00163 g002aCrystals 10 00163 g002b
Figure 3. Simulated UV-Vis absorption spectrum in chlorobenzene solvent of the compounds considered (TD-DFT/B3LYP/6−311G (d, p)).
Figure 3. Simulated UV-Vis absorption spectrum in chlorobenzene solvent of the compounds considered (TD-DFT/B3LYP/6−311G (d, p)).
Crystals 10 00163 g003
Table 1. Selected bond lengths (in Å) and dihedral angles (in degrees) of the optimized structures in gas phase and in solvent (chlorobenzene) using B3LYP/6−311G (d, p) calculations.
Table 1. Selected bond lengths (in Å) and dihedral angles (in degrees) of the optimized structures in gas phase and in solvent (chlorobenzene) using B3LYP/6−311G (d, p) calculations.
r1r2φ1φ2raverφaver
0S1 (1)Gas1.4431.4541131.4497
Solvent1.4441.4551221.4507
CPDT (2)Gas1.4401.4511641.44610
Solvent1.4411.4521451.44710
DTS (3)Gas1.4411.452321.4472
Solvent1.4421.4530.821.4482
DTP (4)Gas1.4401.451110.51.4466
Solvent1.4411.452110.41.4476
DTT (5)Gas1.4431.4532021.44811
Solvent1.4441.4541831.44911
Table 2. Distortion between the neutral and ionic states of designed compounds (B3LYP/6–311G (d, p) calculations, bond distances in angstrom and bond angles in degrees).
Table 2. Distortion between the neutral and ionic states of designed compounds (B3LYP/6–311G (d, p) calculations, bond distances in angstrom and bond angles in degrees).
AnionCation
Δr (Å)Δφ (°)Δr (Å)Δφ (°)
S10.052100.04413
CPDT0.050160.03817
DTS0.04820.0383
DTP0.04580.0378
DTT0.048180.04018
Table 3. HOMO and LUMO energy levels of compounds considered in the gas phase and solvent (chlorobenzene) (eV from B3LYP/6−311G (d, p) computations).
Table 3. HOMO and LUMO energy levels of compounds considered in the gas phase and solvent (chlorobenzene) (eV from B3LYP/6−311G (d, p) computations).
HOMOLUMOEgap
S1Gas Phase−5.28−2.562.7
Solvent−5.40−2.672.74
CPDTGas Phase−5.03−2.562.47
Solvent−5.11−2.642.46
DTSGas Phase−5.08−2.592.50
Solvent−5.17−2.672.50
DTPGas Phase−4.98−2.492.49
Solvent−5.08−2.592.49
DTTGas Phase−5.23−2.602.63
Solvent−5.31−2.672.64
Table 4. Calculated chemical reactivities, vertical electron affinity (EAv) and vertical ionization energy (IEv), along with the adiabatic ones; hole and electron reorganization energy (λh, λe) of the compounds studied (in eV, B3LYP/6−311G (d, p)).
Table 4. Calculated chemical reactivities, vertical electron affinity (EAv) and vertical ionization energy (IEv), along with the adiabatic ones; hole and electron reorganization energy (λh, λe) of the compounds studied (in eV, B3LYP/6−311G (d, p)).
DyeIEvIEaEAvEAaλhλe
S16.216.111.731.850.230.22
CPDT5.935.821.801.920.220.21
DTS5.975.861.821.940.230.21
DTP5.895.791.751.850.200.19
DTT6.126.011.841.960.220.20
Table 5. Calculated absorption wavelengths (λ in nm), oscillator strengths (f) and corresponding orbital transition assignments of the designed compounds (TD-B3LYP/6−311G (d, p)).
Table 5. Calculated absorption wavelengths (λ in nm), oscillator strengths (f) and corresponding orbital transition assignments of the designed compounds (TD-B3LYP/6−311G (d, p)).
Dyeλ (nm)FOrbital TransitionSinglet Electronic
State Transition
S1525
53754
1.7H → L (99%)So → S1
CPDT5882.2H → L (99%)
DTS5842.0H → L (99%)
DTP5782.4H → L (99%)
DTT5452.2H → L (98%)

Share and Cite

MDPI and ACS Style

Trang, N.V.; Dung, T.N.; Cuong, N.T.; Hai, L.T.H.; Escudero, D.; Nguyen, M.T.; Nguyen, H.M.T. Theoretical Study of a Class of Organic D-π-A Dyes for Polymer Solar Cells: Influence of Various π-Spacers. Crystals 2020, 10, 163. https://doi.org/10.3390/cryst10030163

AMA Style

Trang NV, Dung TN, Cuong NT, Hai LTH, Escudero D, Nguyen MT, Nguyen HMT. Theoretical Study of a Class of Organic D-π-A Dyes for Polymer Solar Cells: Influence of Various π-Spacers. Crystals. 2020; 10(3):163. https://doi.org/10.3390/cryst10030163

Chicago/Turabian Style

Trang, Nguyen Van, Tran Ngoc Dung, Ngo Tuan Cuong, Le Thi Hong Hai, Daniel Escudero, Minh Tho Nguyen, and Hue Minh Thi Nguyen. 2020. "Theoretical Study of a Class of Organic D-π-A Dyes for Polymer Solar Cells: Influence of Various π-Spacers" Crystals 10, no. 3: 163. https://doi.org/10.3390/cryst10030163

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