Next Article in Journal
A Gelatin Hydrogel-Containing Nano-Organic PEI–Ppy with a Photothermal Responsive Effect for Tissue Engineering Applications
Next Article in Special Issue
An Electron-Transporting Thiazole-Based Polymer Synthesized Through Direct (Hetero)Arylation Polymerization
Previous Article in Journal
Synthesis, Design, and Structure–Activity Relationship of the Pyrimidone Derivatives as Novel Selective Inhibitors of Plasmodium falciparum Dihydroorotate Dehydrogenase
Previous Article in Special Issue
Synthesis of a 1,2-Dithienylethene-Containing Donor-Acceptor Polymer via Palladium-Catalyzed Direct Arylation Polymerization (DArP)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Performance Comparisons of Polymer Semiconductors Synthesized by Direct (Hetero)Arylation Polymerization (DHAP) and Conventional Methods for Organic Thin Film Transistors and Organic Photovoltaics

Department of Chemical Engineering and Waterloo Institute for Nanotechnology (WIN), University of Waterloo, 200 University Ave West, Waterloo, ON N2L 3G1, Canada
*
Author to whom correspondence should be addressed.
Molecules 2018, 23(6), 1255; https://doi.org/10.3390/molecules23061255
Submission received: 6 May 2018 / Revised: 20 May 2018 / Accepted: 22 May 2018 / Published: 24 May 2018
(This article belongs to the Special Issue Direct (Hetero)Arylation: A New Tool for Organic Electronics)

Abstract

:
C-C bond forming reactions are central to the construction of π-conjugated polymers. Classical C-C bond forming reactions such as the Stille and Suzuki coupling reactions have been widely used in the past for this purpose. More recently, direct (hetero)arylation polymerization (DHAP) has earned a place in the spotlight with an increasing number of π-conjugated polymers being produced using this atom-economic and more sustainable chemistry. As semiconductors in organic electronics, the device performances of the polymers made by DHAP are of great interest and importance. This review compares the device performances of some representative π-conjugated polymers made using the DHAP method with those made using the conventional C-C bond forming reactions when they are used as semiconductors in organic thin film transistors (OTFTs) and organic photovoltaics (OPVs).

Graphical Abstract

1. Introduction

Organic electronics enabled by polymer semiconductors is an exciting field of research that promises to pave the way for low-cost, flexible electronic devices such as organic thin film transistors (OTFTs) [1,2] and organic photovoltaics (OPVs) [3,4]. The performances of these devices have improved greatly in the past few years and has now become commercially viable, i.e., the field effect mobilities and the power conversion efficiencies for the polymer based OTFTs and OPVs have exceeded 10 cm2 V−1 s−1 [5,6,7,8] and 10% [9,10,11], respectively. While studies in this field have been more performance driven, the scale-up synthesis of polymer semiconductors has been attracting increased attention. In particular, novel synthetic methodologies that can produce high performance polymer materials in a more environmentally friendly way, i.e., using “green chemistry” [12,13,14,15], with increased atom economy and reduced production costs are highly desirable.
Central to the construction of polymer semiconductors is the C-C bond forming reaction that links the monomeric units. Common methods for the C-C bond formation, such as Stille and Suzuki coupling reactions require an aryl halide and an aromatic compound with a reactive directing group, e.g., a boronic acid (or ester) for the Suzuki coupling and an organostannyl group for the Stille coupling. These reactions, while highly effective in C-C bond formation, require additional steps to install the directing groups, which increases the production cost and generates stoichiometric amounts of by-products that are potential health and environmental hazards. Specifically, the organotin compounds formed from the Stille coupling reactions are known to be highly toxic [16], while the boronic acid derivatives used in the Suzuki coupling reactions, which were previously assumed to be less harmful, have been recently found to be potential genotoxic hazards [17].
To address these issues associated with the common synthetic methods used to prepare polymer semiconductors, a novel C-C bond forming methodology, the so-called direct (hetero) arylation polymerization (DHAP) has been explored recently (Figure 1) [12]. The DHAP method eliminates the need for adding a directing group. Instead, the carbon atom with the most “active hydrogen” in the monomer is able to couple with the halogenated carbon atom in another (or the same) monomer. However, many monomer compounds have multiple C-H bonds with close dissociation energies, which can potentially be activated and react with a C-halogen bond. Furthermore, two Pd(II) complex intermediates bearing equal (hetero)aryl groups may undergo a disproportionation reaction, resulting in a homocoupling defect [18,19,20]. These side reactions may impede the formation of soluble (in the case of crosslinking side reaction) or high molecular weight (in the case of homocoupling side reaction) polymer products. Even for the polymers with good solubility and high molecular weights made by DHAP, a certain amount of branching, crosslinking, and/or homocoupling defects are frequently observed [21,22,23,24,25,26]. Figure 1 shows the formation of these defects in the DHAP of 3-alkyl-2-bromo-thiophene to poly(3-alkylthiophene). In the past few years, rigorous studies have been conducted to optimize the synthetic conditions to minimize or eliminate these side reactions. With a better understanding of the DHAP mechanism, a number of high-quality polymer semiconductors with fewer structural defects have been synthesized using the DHAP method [27].
An important question or concern from the organic electronics community is: are the performances of the polymers made by DHAP comparable to those of the polymers made by the conventional synthetic methods? In this review, we will provide a preliminary answer to this question by judiciously selecting some representative polymer semiconductors that were made by both the DHAP method and the conventional methods and compare their performances in OTFTs and OPVs. For the recent and significant progress made in clarifying the mechanism and optimizing reaction conditions of DHAP to improve the yield and molecular weight and reduce/eliminate structural defects of the resulting polymers, readers are directed to several recent review articles [12,28,29,30].

2. Polymer Semiconductors for Organic Thin Film Transistors

Poly(3-hexyl-thiophene-2,5-diyl) (P1, Figure 2), also known as P3HT, is one of the most widely studied polymer semiconductors to date. Due to its widespread use in the organic electronics community, it has become a standard, to which new materials can be compared [31]. The monomer unit, 3-hexyl-thiophene-diyl, is unsymmetrical, which means that three different types of dyads, heat-to-head (HH), head-to-tail (HT), and tail-to-tail (TT) can be formed depending on the positions of the hexyl substituents on the two thiophene units. In an HH dyad, the two closely positioned hexyl substituents would cause severe twisting of the two thiophene rings, which would disrupt the π-conjugation pathway along the polymer backbone. The TT dyad is highly coplanar, but the formation of each TT dyad is accompanied with the formation of one HH dyad during the polymerization. On the other hand, the HT dyad is highly coplanar and a P3HT comprising of only HT dyads is a ‘regioregular’ HT P3HT or commonly rr-P3HT. The first P3HT was made using an FeCl3-mediated oxidative coupling polymerization, which contained a significant amount of HH dyads with an HT mol % or regioregularity of ~70–80% [32,33,34,35] and thus had a rather twisted backbone. P3HT made in this way has poor crystallinity, which leads to low charge carrier mobilities of ~10−4 cm2 V−1 s−1 in OTFT devices [36].
In 1992, McCullough et al. reported the use of Grignard reagents to produce an rr-P3HT with a high regioregularity of up to 98% [37,38]. Also reported in 1992 was the discovery of a synthetic methodology based on zinc reagents by Rieke et al. [39,40], which could produce rr-P3HT with regioregularity of >98%. In subsequent studies, regioregular poly(3-alkyl-thiophene-2,5-diyl) samples were also made using Suzuki coupling (HT mol % = 96–97%) [41], Stille coupling polymerization (HT mol % > 96%) [42], or Grignard metathesis (GRIM, HT mol % > 98%) [43,44]. The synthetic schemes for these methods are depicted in Figure 2. P3HT with a high HT mol % of 96% made by the Rieke method demonstrated a high hole mobility of up to 0.10 cm2 V−1 s−1 in OTFTs, which is several orders of magnitude higher than that for the P3HT with lower regioregularity prepared using FeCl3 and Yamamoto couplings [36]. Thus, it would be no exaggeration to say that the birth of rr-P3HT has brought the study of organic electronics into a new era.
The aforementioned synthetic routes to rr-P3HT require additional synthetic steps to add a directing group and/or produce stoichiometric harmful by-products. Thus, it was of interest to researchers to explore alternative synthetic pathways for forming highly regioregular P3HT without the above drawbacks.
In 1998, Lemaire and co-workers reported a new method to synthesize polythiophenes using 2-iodothiophenes as the monomers and Pd(OAc)2 as the catalyst [46]. The authors initially thought the polymerization was a Heck-type reaction, but it was later found to be a direct (hetero) arylation reaction, which involves a direct coupling of an active aromatic C-H bond with an aromatic C-X bond (X is halide) to form a C-C bonding motif [47]. In 2010, Wang et al. conducted a DHAP of 2-bromo-3-hexylthiophene with Herrmann’s catalyst to make an rr-P3HT [45]. High number average molecular weights (Mn) of up to 31 kDa were achieved, which are much higher than those of the polythiophenes (~103 Da) through DHAP reported previously by Lemaire and coworkers [46]. A very high degree of regioregularity (>98%) of this rr-P3HT was confirmed by the 1H NMR spectroscopic analysis, indicating that electronic grade P3HT was attainable using DHAP synthetic methods.
Pouliot et al. [48] prepared a sample of P1 with a very high regioregularity of >99.5% and an Mn of 33 kDa by optimizing the synthesis of the 2-bromo-3-hexylthiophene monomer and DHAP conditions (P1DHAP1, Table 1). They compared the OTFT performance of this polymer with other rr-P3HT polymers made by GRIM and Rieke’s methods using the same device configuration and testing conditions (P1GRIM and P1Rieke, Table 1). It was found that high hole mobilities of up to 0.19 cm2 V−1 s−1 were achieved by the polymer made by DHAP, which are much higher than those of P1GRIM (0.11 cm2 V−1 s−1) and P1Rieke (up to 0.02 cm2 V−1 s−1). The mobility trend can be explained by their regioregularity order: P1DHAP1 (HT mol % > 99.5%) > P1GRIM (HT mol % = 98.0%) > P1Rieke (HT mol % = 95.5%). The results from this study show that the performance of rr-P3HT made by DHAP in OTFTs is improved compared to those of the rr-P3HT polymers made by the Rieke and GRIM methods.
A donor-acceptor (D-A) type polymer, poly {[N,N′-bis(2-octyldodecyl)-naphthalene-1,4,5,8-bis(dicarboximide)-2,6-diyl]-alt-5,5′-(2,2′-bithiophene)}, (P2, Figure 3) also known as N2200, is considered to be the most extensively investigated n-type polymer for OTFTs [49,50,51] and OPVs [52,53,54,55,56,57]. P2 has commonly been synthesized using the Stille coupling polymerization [58], so the ability to synthesize it using atom-economical methods is of interest. In 2015, the DHAP synthesis of N2200 was reported by Sommer et al. [50] (P2DHAP, Table 1). Optimized DHAP conditions gave the polymer with an Mn of up to 31 kDa and a ‘perfect’ NDI-bithiophene alternating structure, with no homocoupling, branching or cross-linking, according to 1H NMR spectroscopic analysis. Importantly, the molecular weight of the polymer was readily controlled via in situ solvent (toluene) end capping by adjusting the monomer concentration. Another important observation was that the optical and thermal properties of the DHAP polymers reach their maxima at Mn = ~20 kDa. This is in excellent agreement with the sample produced via the Stille coupling (P2Stille, Table 1). In OTFTs, the mobilities of P2DHAP (2.9 cm2 V−1 s−1) and P2Stille (3.2 cm2 V−1 s−1) with similar Mn were very close, which are among the highest values reported for n-type polymers to date (OTFT transfer curves for P2DHAP and P2Stille are shown in Figure 4).
Another naphthalene-1,4,5,8-bis(dicarboximide) (NDI) based polymer P3 (Figure 3), which is composed of NDI, thiophene and tetrafluorobenzene, was synthesized by DHAP (P3DHAP, Table 1) [59]. A rather low Mn of 7.8 kDa was obtained. A detailed 1H NMR spectroscopic analysis showed the absence of β-arylation, indicating that this polymer has a linear structure. Despite its rather low molecular weight, P3DHAP showed high electron mobilities of up to 1.3 cm2 V−1 s−1 when being used as an n-channel semiconductor in OTFTs. The same polymer synthesized by other methods has not been reported.
Diketopyrrolopyrrole (DPP) or pyrrolo[3,4-c]pyrrole-1,4(2H,5H)-dione is a strong acceptor building block, which has been used extensively for the construction of a large number of ultrahigh mobility donor-acceptor (D-A) type polymer semiconductors for OTFTs [60]. However, the most high-performing DPP-based polymer materials for OTFTs such as PDBT-co-TT [5,61], PDQT [62], PDPP-TVT [63], and P-29-DPPDTSE [64] have not been successfully synthesized by DHAP yet. For example, an attempt to prepare PDPP-TVT by DHAP of 3,6-bis(5-bromothiophen-2-yl)-2,5-bis(2-decyltetradecyl) pyrrolo[3,4-c]pyrrole-1,4-(2H,5H)-dione (DPP) and (E)-1,2-bis(thiophen-2-yl)ethane (TVT) under typical conditions failed to produce the target polymer [65]. Interestingly, when a fluorinated TVT, (E)-1,2-bis(3,4-difluorothiophen-2-yl)ethane, 4FTVT), was used, a polymer PDPP-4FTVT (P4) with a high Mn of up to 60 kDa was successfully obtained (P4DHAP, Table 1). This new polymer showed an ambipolar charge transport behavior with high hole and electron mobilities of up to 3.4 cm2 V−1 s−1 and 5.9 cm2 V−1 s−1, respectively, in OTFTs. This study suggests that fluorine substitution might be able to suppress the β-arylation side reactions and/or activate the α-H of the 4FTVT monomer.
Another copolymer of DPP with tetrafluorobenzene, PDPPTh2F4 (P5), was synthesized by DHAP [66], which has not been approached by other synthetic methods (P5DHAP, Table 1). Due to the strong electron withdrawing effect of the tetrafluorobenzene units, this polymer showed n-type electron transport semiconductor behavior, opposed to the majority of other DPP polymers that showed either p-type or ambipolar charge transport performance. High electron mobility of up to 0.60 cm2 V−1 s−1 was achieved in OTFTs.
A copolymer of DPP and bithiazole, PDBTz-24 (P6), was synthesized by Guo et al. using DHAP of dibrominated bisthienyl DPP and 2,2′-bithiazole monomers (P6DHAP, Table 1) [67]. 1H NMR spectra showed some unidentified structural defects possibly resulting from homocoupling and/or branching. Nonetheless, this polymer exhibited an electron transport dominant ambipolar performance with hole and electron mobilities of up to 0.06 cm2 V−1 s−1 and 0.53 cm2 V−1 s−1 in OTFTs, outperforming a similar polymer PDBTz-27 (P6Stille) with 5-decylheptadecyl side chains made by the Stille coupling previously reported by Reichmanis et al. [68], which showed electron mobilities of up to 0.31 cm2 V−1 s−1.
A DPP-DPP type polymer (P7) was synthesized via DHAP between 3,6-bis(thiophen-2-yl)-2,5-bis(dodecyl)pyrrolo[3,4-c]pyrrole-1,4(2H,5H)-dione and 2,5-bis(2-octyldodecyl)-3,6-bis(5-bromo-4-methylthiophene-2-yl)pyrrolo[3,4-c]pyrrolo-1,4(2H,5H)-dione by Pouliot et al. (P7DHAP, Table 1) [24]. A high Mn of 46 kDa along with a narrow dispersity (Mw/Mn) of 2.5 was obtained, which suggested that negligible side reactions occurred. P-type semiconductor performance with high hole mobilities of up to 1.2 cm2 V−1 s−1 was achieved in OTFTs. Interestingly, when the reactive C-H and C-Br exchanged between the two monomers, that is, 2,5-bis(2-octyldodecyl)-3,6-bis(4-methylthiophene-2-yl)-pyrrolo[3,4-c]pyrrolo-1,4(2H,5H)-dione and 3,6-bis-(5-bromothiophen-2-yl)-2,5-bis(2-dodecyl)pyrrolo[3,4-c]pyrrole-1,4(2H,5H)-dione were polymerized, the resulting polymer had a lower Mn (15 kDa) and showed lower hole mobility (0.26 cm2 V−1 s−1), which was ascribed to some ill-defined couplings such as homocoupling and branching.

3. Polymer Semiconductors for Organic Photovoltaics

The use of rr-P3HT (P1) synthesized by DHAP in solar cells was first reported by Thompson et al. [25]. A sample of rr-P3HT (P1DHAP2) with Mn = 19 kDa, Mw/Mn = 2.0, and HT mol % = 90% and a control polymer P1Stille1 with Mn = 19 kDa, Mw/Mn = 2.7, and HT mol % = 93% were synthesized by DHAP and Stille coupling methods, respectively (Table 2). Interestingly, it was shown that P1DHAP2 had ~0.75% β-defects, but still showed even better solar cell performance compared to P1Stille1, which had no β-defects (PCE: 2.70% for P1DHAP2 vs. 2.30% for P1Stille1). The authors made another sample using DHAP, having 1.41% β-defect concentration, and found that the performance was poorer due to formation of unfavorable morphology when blended with fullerene and the low space-charge-limited current (SCLC) mobility, which resulted in low JSC. The authors concluded that rr-P3HT synthesized by DHAP with β-defect concentrations below 1% were able to achieve similar or better performance in solar cell devices when compared with their counterparts made by the Stille coupling method.
Another recent study by Thompson et al. compared the properties of rr-P3HT synthesized using a “green solvent assisted” DHAP (P1DHAP3) with a sample prepared by the traditional Stille coupling method (P1Stille2, Table 2) [69]. P1DHAP3 had Mn = 20 kDa, Mw/Mn = 2.1, and HT mol % = 96.2%, while P1Stille2 had Mn = 18 kDa, Mw/Mn = 2.4, and HT mol % = 92.9%. These authors found that the use of a bulky carboxylic acid additive, neodecanoic acid, and a “green” solvent, 2-methyl-tetrahydrofuran, was critical in completely avoiding the formation of β-defects. In UV-Visible absorption spectra, the β-defect free P1DHAP3 displayed a higher extinction coefficient at the wavelength of maximum absorbance (λmax) and stronger vibrational fine structure, indicating more structural ordering in the solid state. P1DHAP3 also had larger crystalline coherence lengths, as confirmed by X-ray diffraction, which were correlated to the increased SCLC hole mobility for this polymer. When evaluated as a donor with PC61BM as the acceptor in solar cells, P1DHAP3 outperformed P1Stille2 (PCE: 3.28% vs. 2.86%, Table 2). This is in agreement with the previous results by the same group, implying that a rr-P3HT sample with negligible β-defects produced by DHAP can outperform its counterpart made by the Stille coupling [25].
Kanbara et al. synthesized copolymers of fluorene and EDOT (P8, Figure 5) using both DHAP and Suzuki coupling methods (P8DHAP and P8Suzuki, Table 2) [70]. P8DHAP made by DHAP with a microwave reactor had a much higher Mn of 150 kDa than that of P8Suzuki made by Suzuki coupling, which had an Mn of 17 kDa. As a donor, P8DHAP demonstrated a high PCE of 4.08%, while P8Suzuki exhibited a very low PCE of 0.480%. The large discrepancy in the OPV performance of these two polymers is ascribed mainly to their different hole mobilities. In OTFTs, P8DHAP showed an average hole mobility of 1.2 × 10−3 cm2 V−1 s−1, while the average hole mobility of P8Suzuki is about two orders of magnitude lower at 3.2 × 10−5 cm2 V−1 s−1. Another P8DHAP with a lower Mn of 48 kDa showed a lower PCE of 2.55%, which was believed to be due to the presence of terminal Br groups and residual Pd impurities in the polymers, causing a decrease in the hole mobility (7.7 × 10−4 cm2 V−1 s−1). In a later study by the same group, P8DHAP showed further improved PCE of up to 4.60% after the terminal Br groups were removed through a post-polymerization treatment of the polymer with sodium N,N-diethyldithiocarbamate [71].
D-A polymers are among the most popular materials for use in OPVs and synthesis of high-quality D-A polymers by DHAP was therefore explored. Russel et al. reported on two D-A polymers based on DPP, DPP-co-phenylene (PDPP-TPT, P9, Figure 5) and DPP-co-thiophene, (PDPP3T, P10), which were synthesized by DHAP (P9DHAP and P10DHAP, Table 2) [22]. Optimized DHAP conditions provided a sample of P9DHAP with an Mn of 14 kDa and an Mw/Mn of 1.8. P9DHAP produced via DHAP in this work was compared with one produced by Janssen’s group using a Suzuki coupling protocol (P9Suzuki1, Table 2) [72], which showed a bimodal GPC trace with two peak molecular weights of 65 kDa and 10 kDa due to the gelation and aggregation. As a donor, P9DHAP showed slightly lower PCE in OPVs than P9Suzuki1 when PC61BM was used as the acceptor (4.37 vs. 5.50%, Table 2). Later, optimization of the Suzuki coupling reaction gave a sample of P9Suzuki2 with a higher Mn reaching 72 kDa and an Mw/Mn of 1.98. Solar cells using P9Suzuki2 gave PCE of up to 7.40% [73]. For the second polymer reported in this work [22], PDPP3T, optimized DHAP conditions gave a sample of P10DHAP with Mn 29 kDa and an Mw/Mn of 3.8. A previous paper by Janssen et al. in 2009 [74] reported on the synthesis of PDPP3T using a Suzuki coupling protocol, which gave P10Suzuki with an Mn of 54 kDa, nearly twice that of P10DHAP reported in this work. In OPVs, P10DHAP was also outperformed by P10Suzuki (4.01% vs. 4.69%, Table 2). A later work by Janssen et al. improved the synthesis of P10 via Stille coupling [73] (P10Stille, Table 2), achieving an improved Mn of 150 kDa and an Mw/Mn of 2.72. Solar cells made using P10Stille gave much better PCE of up to 7.10% when used as the donor material in OPVs. 1H NMR spectroscopy was unable to determine the presence of homocoupling defects in P9DHAP and P10DHAP due to the tendency to aggregate in solution; however, the authors suggested that some homocouplings (<5%) may be present in these samples because a broadening of the low energy absorption shoulders was seen, a feature which is often attributed to minor homocoupling defects [75]. Both the presence of the homocoupling defects as well as the lower molecular weights might account for the lower mobilities observed for P9DHAP and P10DHAP compared with their counterparts made by the Suzuki and Stille couplings, P9Suzuki1, P9Suzuki2, P10Suzuki and P10Stille which contained lesser amounts of defects and had higher molecular weights.
Poly[4,4-bis(2-ethylhexyl)-4H-cyclopenta[2,1-b;3,4-b′]dithiophene-2,6-diyl-alt-2,1,3-benzothiadiazole-4,7-diyl] (PCPDTBT, P11) is a high performance polymer semiconductor, used commonly as a donor in OPVs [76,77,78], and a p-type semiconductor in OTFTs [79]. In 2012, Horie et al. synthesized PCPDTBT using both DHAP and Suzuki methods (P11DHAP, P11Suzuki, Table 2) for a direct comparison [80]. P11DHAP had an Mn of 72 kDa with an Mw/Mn of 4.52, while P11Suzuki had a much lower Mn of 15 kDa with an Mw/Mn of 2.1. By using MALDI-TOF and 1H NMR spectroscopic analysis, it was found that P11DHAP contained both homocoupling and branching defects, while P11Suzuki did not. The β-defect concentration was found to be 12–13% for P11DHAP, which is well above the “threshold” value of ~1% for rr-P3HT for achieving good OPV performance observed by Thompson et al. [25]. Despite such high β-defect concentration, P11DHAP exhibited higher PCE in solar cells using PC71BM as the acceptor under the same processing and testing conditions compared to P11Suzuki (PCE: 3.98% vs. 3.74%. respectively) [80]. P11Stille synthesized by the Stille coupling previously, which had an Mn of 28 kDa, achieved a lower PCE of 3.50% [77] (P11Stille, Table 2). These results suggest that the PCPDTBT system has a high tolerance for β-defects, which might be due to the beneficial effect of the branching structure on the morphology of the P11DHAP:PC71BM blend films [80] as well as the much higher intrinsic hole mobility of this polymer [79] compared to P3HT. The fact that P11DHAP showed the best PCE is most likely due to its highest molecular weight, which is a more dominant factor than the β-defect concentration in the OPV performance. Given that the rigorous optimization of device fabrication for commercial samples of P11 afforded an improved PCE of up to 6% when paired with PC71BM [76,78], it is possible that P11DHAP may also exhibit a higher PCE when the active layer formation is further optimized.
Thompson et al. synthesized benzothiadiazole-co-benzodithiophene polymers (P12) using DHAP and Stille methods (P12DHAP, P12Stille2-LMW, P12Stille1-HMW, Table 2) [81]. The optimized DHAP protocol afforded a ‘defect-free’ PPDTBT sample P12DHAP with an Mn of 15 kDa and an Mw/Mn of 2.1 (P12DHAP, Table 2). The use of THF as a solvent and the bulky neodecanoic acid as an additive was found to be essential in minimizing structural defects during DHAP based on the 1H NMR spectroscopic analysis [69]. Two PPDTBT samples having Mn of 59 and 16 kDa and Mw/Mn of 3.3 and 2.1, respectively, were synthesized by the Stille method for comparison (P12Stille1-HMW vs. P12Stille2-LMW, Table 2). P12DHAP showed a better PCE (3.40%) compared with the counterpart P12Stille2-LMW made by Stille (PCE = 2.90%), which has a similar Mn and Mw/Mn. On the other hand, polymer P12Stille1-HMW with a high molecular weight made by the Stille method was able to achieve a higher PCE of 3.80%. These results illustrate, once again, the importance of the polymer molecular weight on device performance. It can be concluded that the quality of the polymer P12 made by DHAP is similar to that made from Stille coupling if the molecular weight can be matched.
In 2015, Farinola et al. reported on the DHAP synthesis of a random copolymer (P13), which used benzo[c][1,2,5]thiadiazole or benzo[d][1,2,3]triazole as the accepting units and benzo[1,2-b:4,5-b′]dithiophene as the donor component [23]. Samples of P13 were prepared via DHAP and also by Stille coupling polymerization for direct comparison. DHAP yielded a sample (P13DHAP) with a low Mn of only 10 kDa compared to the Mn of 20 kDa obtained for P13Stille synthesized via Stille coupling. In addition, the Mw/Mn was much larger for the DHAP synthesized sample compared to the one made using a Stille coupling (Mw/Mn: 7.6 for P13DHAP vs. 3.1 for P13Stille, Table 2). A hypsochromic shift in the absorption spectra is noticed for the DHAP synthesized polymer, indicative of a lower effective conjugation length. 1H NMR analysis revealed the presence of β-defects for P13DHAP. When the two polymers were tested in solar cells with PC71BM as the acceptor, P13DHAP gave a lower PCE of 2.80% compared to a value of 4.80% for P13Stille. Such a large difference in performance between P13DHAP and P13Stille can be directly ascribed to the lower molecular weight and higher defect concentration in the former. Optimization of the catalytic systems for DHAP such as the use of bulky carboxylic acid additives and specialized catalysts [27,69,82] may provide a sample of P13 with a lower concentration of defects and thus a higher effective conjugation length, which would improve its performance to be on par with P13Stille.
In 2016, Marks et al. reported a comprehensive study to compare the physical and electronic properties of a series of D-A polymers made using both DHAP and Stille methods [83]. The first polymer is PBDTT-FTTE (P14), a well-known high performance p-type polymer that has been used in record setting OPVs [84,85]. It was found that increasing the steric bulkiness of the carboxylic acid additive, i.e., the use of 2,2-diethylhexanoic acid instead of pivalic acid, in the DHAP system was critical in forming high molecular weight, defect-free polymers, which is in agreement with other recent reports [69,81]. Optimization of DHAP conditions gave P14DHAP with Mn = 25 and Mw/Mn = 2.2, a very close match with P14Stille made by the Stille coupling (Mn = 25 kDa and Mw/Mn = 2.2). When evaluated in solar cell devices with PC71BM as the acceptor, P14DHAP gave an almost same PCE compared with P14Stille (PCE: 8.36% vs. 8.40%, Table 2 and Figure 6a). In addition, the two polymers showed very similar film morphologies when blended with PC71BM (Figure 6b). Two other pairs of D-A polymers based on thieno[3,4-c]pyrrole-4,6-dione (TPD), namely PBDTT-TPD (P15DHAP and P15Stille, Table 2) and PTPD3T (P16DHAP and P16Stille Table 2), were also made using both DHAP and Stille methods [83]. It was found that P15DHAP has a higher Mn than that of P15Stille (Mn: 30 kDa vs. 15 kDa), while P16DHAP had a lower Mn than that of P16Stille (Mn: 19 kDa vs. 30 kDa). 1H NMR spectroscopy confirmed that the structural regularity of both polymers made by DHAP was similar to that of the ones made by the Stille method. In addition, the absorption spectra of each pair of polymers are nearly identical. In OPVs, P15DHAP was found to outperform P15Stille (PCE: 5.84% vs. 5.20%), while P16DHAP performed slightly poorer than P16Stille (PCE: 7.20% vs. 7.38%). Since P15DHAP and P16DHAP did not contain noticeable structural defects, the differences in their OPV performance with their counterparts P15Stille and P16Stille are likely related to their different molecular weights, with the higher molecular weight ones showing higher OPV performance.
PBDTTPD (P17) is another high performance donor polymer for OPVs [86]. Farinola et al. conducted a study comparing the optoelectronic properties and device performance of PBDTTPD made by both DHAP and Stille methods (P17DHAP and P17Stille, Table 2) [87]. Optimized DHAP conditions produced P17DHAP with a low Mn of 12 kDa, while P17Stille synthesized via the Stille coupling protocol had an even lower Mn of 10 kDa. The slightly higher Mn of P17DHAP resulted in a red-shifted UV-Visible absorption profile as well a clearer vibronic structure in the low-energy region compared to P17Stille. In addition, the aggregation propensity of P17DHAP in solution was noted to be higher. Furthermore, I t was found that P17Stille allowed the formation of large micro-sized domains of PC71BM in the bulk-heterojunction blend, while P17DHAP did not. OPV devices using these two polymers as the donors and PC71BM as the acceptor were fabricated and compared. PCEs of up to 5.31% and 4.82% were obtained for P17DHAP and P17Stille, respectively, which is in good agreement with the disparity in their molecular weights.
A copolymer of TPD and cyclopenta[2,1-b:3,4-b′]dithiophene (P18) was synthesized by Kanbara et al. using an improved DHAP in toluene, which had an Mn of 25 kDa with an Mw/Mn of 1.9 (P18DHAP, Table 2) [19]. The molecular weight is much higher than that (Mn = 2.5 kDa) of the same polymer made by DHAP by Harris et al. using DMAc as the solvent, where severe homocoupling occurred, resulting in the lower Mn [80]. On the other hand, the 1H NMR and MALDI-TOF-MS analysis of P18DHAP made in toluene in this work showed no detectable homocoupling defects. A high PCE of up to 6.80% was achieved when P18DHAP was used as a donor in OPVs, which is higher than that (PCE = 5.20%) of P18Stille prepared by the Stille coupling, which had a lower Mn of 15 kDa [88].
Ahmed et al. reported on another series of polymers based on [3,4-c]pyrrole-4,6-dione derivatives with sulfur, oxygen or selenium heteroatom substitution, made by the DHAP method [89]. The champion polymer from their study, P19DHAP, was based on a seleno[3,4-c]pyrrole-4,6-dione acceptor and a dithienosilole donor and was obtained with an Mn of 29 kDa and an Mw/Mn of 1.6. When paired with PC61BM in inverted organic solar cells, a high PCE of 7.13% was achieved. No equivalent polymer to P19DHAP made by a conventional coupling method was reported for comparison.
A recent report by Li et al. compared the properties of indacenodithiophene (IDT) and thiophene-quinoxaline-thiophene (TQ) derived polymers (P20) made using DHAP and Stille coupling methods (P20DHAP and P20Stille, Table 2) [21]. P20DHAP had an Mn of 27 kDa with an Mw/Mn of 1.6, while P20Stille had an Mn of 23 kDa and an Mw/Mn of 1.5. The close number average molecular weights of the two samples allow for a fair comparison between the two polymers. 1H NMR spectroscopy, MALDI-TOF MS, and elemental analysis revealed that P20Stille contained a certain amount of TQ−TQ homocoupling defects, while P20DHAP had a very well-defined IDT-TQ alternating structure with no detectable structural defects. Interestingly, SCLC experiments showed a weak correlation between their hole mobility and structural regularity. Furthermore, grazing incidence X-ray diffraction (GIXRD) studies indicated no significant differences in the film crystallinity between two polymers. P20DHAP and P20Stille achieved PCE of 5.10% and 4.82%, respectively. It was found that the larger number of defects in the backbone of P20Stille had influenced the morphological properties and SCLC mobilities, which caused severe recombination and thus the lower PCE.

4. Conclusions

Recently, direct (hetero)arylation polymerization (DHAP) of a halogenated (hetero)aryl compound with another (hetero)aryl compound possessing activated C-H bonds has emerged as an alternative approach to the conventional C-C bond formation methods such as Stille and Suzuki coupling polymerization for the synthesis of linear π-conjugated polymers. Since DHAP does not require the use of a (hetero)aryl compound possessing a reactive directing group such as an organostannyl group in the case of a Stille coupling polymerization, fewer steps are needed for DHAP, which can improve the atom economy and reduce the cost for the synthesis of conjugated polymers. Furthermore, the by-product of a DHAP is a hydrogen halide (HX), which is much easier to treat compared with the toxic wastes generated from the Stille and Suzuki coupling polymerization. Therefore, DHAP has been considered a more cost-effective and “greener” technique for the mass production of conjugated polymers. However, since many (hetero)aryl compounds have more activated C-H bonds than required for the formation of a linear polymer, side reactions at the unwanted C-H bonds result in branching or cross-linking structural defects, often leading to degradation of the optoelectronic properties or even the formation of unprocessable insoluble products. In addition, homocoupling between the halogenated monomer molecules produces homocoupling defects in the polymer backbone and also results in the formation of lower molecular weight polymers. The presence of these structural defects in the polymers made by DHAP potentially have impacts on the performances of these polymers in organic electronic devices.
This review presented some exemplary conjugated polymer semiconductors made by DHAP and compared their performances in OTFTs and OPVs with their counterparts synthesized by conventional synthetic methods, such as the Stille and Suzuki coupling polymerization. It can be seen that by minimizing the structural defects and improving the molecular weight of the polymers through optimization of the DHAP conditions, similar or sometimes better OTFT and OPV performances could be achieved for the polymers made by DHAP in comparison to the ones prepared by the conventional methods. With the rapid progress made in the development of novel catalyst systems and optimization of reaction conditions (solvent, temperature, reaction time, etc.), it is expected that more and more high-performance polymer semiconductors, which have been synthesized and cannot be approached by conventional methods, will be developed. Further study of the atom-economic, cost-effective, and “greener” DHAP to enable the mass production of top-performing polymer semiconductors will propel the commercialization and wide-spread applications of organic electronics.

Acknowledgments

This work is supported by the Natural Sciences and Engineering Research Council (NSERC) of Canada with the Postdoctoral Fellowship (ADH) and Discovery Grants (#RGPIN-2016-04366 for YL). Article processing charge was sponsored by MDPI.

Conflicts of Interest

The authors declare no conflicting interests.

References

  1. Bao, Z.; Locklin, J. Organic Field-Effect Transistors; Optical Science and Engineering; Taylor & Francis: Abingdon, UK, 2007; ISBN 978-1-4200-0801-2. [Google Scholar]
  2. Arias, A.C.; MacKenzie, J.D.; McCulloch, I.; Rivnay, J.; Salleo, A. Materials and Applications for Large Area Electronics: Solution-Based Approaches. Chem. Rev. 2010, 110, 3–24. [Google Scholar] [CrossRef] [PubMed]
  3. Dennler, G.; Scharber, M.C.; Brabec, C.J. Polymer-Fullerene Bulk-Heterojunction Solar Cells. Adv. Mater. 2009, 21, 1323–1338. [Google Scholar] [CrossRef]
  4. Lu, L.; Zheng, T.; Wu, Q.; Schneider, A.M.; Zhao, D.; Yu, L. Recent Advances in Bulk Heterojunction Polymer Solar Cells. Chem. Rev. 2015, 115, 12666–12731. [Google Scholar] [CrossRef] [PubMed]
  5. Li, J.; Zhao, Y.; Tan, H.S.; Guo, Y.; Di, C.-A.; Yu, G.; Liu, Y.; Lin, M.; Lim, S.H.; Zhou, Y.; et al. A stable solution-processed polymer semiconductor with record high-mobility for printed transistors. Sci. Rep. 2012, 2, 754. [Google Scholar] [CrossRef] [PubMed]
  6. Bucella, S.G.; Luzio, A.; Gann, E.; Thomsen, L.; McNeill, C.R.; Pace, G.; Perinot, A.; Chen, Z.; Facchetti, A.; Caironi, M. Macroscopic and high-throughput printing of aligned nanostructured polymer semiconductors for MHz large-area electronics. Nat. Commun. 2015, 6, 8394. [Google Scholar] [CrossRef] [PubMed]
  7. Kim, G.; Kang, S.-J.; Dutta, G.K.; Han, Y.-K.; Shin, T.J.; Noh, Y.-Y.; Yang, C. A Thienoisoindigo-Naphthalene Polymer with Ultrahigh Mobility of 14.4 cm2/V·s That Substantially Exceeds Benchmark Values for Amorphous Silicon Semiconductors. J. Am. Chem. Soc. 2014, 136, 9477–9483. [Google Scholar] [CrossRef] [PubMed]
  8. Luo, C.; Kyaw, A.K.K.; Perez, L.A.; Patel, S.; Wang, M.; Grimm, B.; Bazan, G.C.; Kramer, E.J.; Heeger, A.J. General Strategy for Self-Assembly of Highly Oriented Nanocrystalline Semiconducting Polymers with High Mobility. Nano Lett. 2014, 14, 2764–2771. [Google Scholar] [CrossRef] [PubMed]
  9. Zhao, W.; Li, S.; Yao, H.; Zhang, S.; Zhang, Y.; Yang, B.; Hou, J. Molecular Optimization Enables over 13% Efficiency in Organic Solar Cells. J. Am. Chem. Soc. 2017, 139, 7148–7151. [Google Scholar] [CrossRef] [PubMed]
  10. Ye, L.; Xiong, Y.; Zhang, Q.; Li, S.; Wang, C.; Jiang, Z.; Hou, J.; You, W.; Ade, H. Solar Cells: Surpassing 10% Efficiency Benchmark for Nonfullerene Organic Solar Cells by Scalable Coating in Air from Single Nonhalogenated Solvent. Adv. Mater. 2018, 30, 1870054. [Google Scholar] [CrossRef]
  11. Lin, Y.; Zhao, F.; Prasad, S.K.K.; Chen, J.-D.; Cai, W.; Zhang, Q.; Chen, K.; Wu, Y.; Ma, W.; Gao, F.; et al. Balanced Partnership between Donor and Acceptor Components in Nonfullerene Organic Solar Cells with >12% Efficiency. Adv. Mater. 2018, 30, 1706363. [Google Scholar] [CrossRef] [PubMed]
  12. Bohra, H.; Wang, M. Direct C-H arylation: A “Greener” approach towards facile synthesis of organic semiconducting molecules and polymers. J. Mater. Chem. A 2017, 5, 11550–11571. [Google Scholar] [CrossRef]
  13. Burke, D.J.; Lipomi, D.J. Green chemistry for organic solar cells. Energy Environ. Sci. 2013, 6, 2053–2066. [Google Scholar] [CrossRef]
  14. Anastas, P.; Warner, J. Green Chemistry, Theory and Practice; Oxford University Press: New York, NY, USA, 2000. [Google Scholar]
  15. Organic Electronics for a Better Tomorrow: Innovation, Accessibility, Sustainability; Chemical Society and Sciences Summit (CS3): San Francisco, CA, USA, 2012.
  16. Kimbrough, R.D. Toxicity and health effects of selected organotin compounds: A review. Environ. Health Perspect. 1976, 14, 51–56. [Google Scholar] [CrossRef] [PubMed]
  17. Hansen, M.M.; Jolly, R.A.; Linder, R.J. Boronic Acids and Derivatives—Probing the Structure–Activity Relationships for Mutagenicity. Org. Process Res. Dev. 2015, 19, 1507–1516. [Google Scholar] [CrossRef]
  18. Lombeck, F.; Komber, H.; Gorelsky, S.I.; Sommer, M. Identifying Homocouplings as Critical Side Reactions in Direct Arylation Polycondensation. ACS Macro Lett. 2014, 3, 819–823. [Google Scholar] [CrossRef]
  19. Kuwabara, J.; Fujie, Y.; Maruyama, K.; Yasuda, T.; Kanbara, T. Suppression of Homocoupling Side Reactions in Direct Arylation Polycondensation for Producing High Performance OPV Materials. Macromolecules 2016, 49, 9388–9395. [Google Scholar] [CrossRef]
  20. Rudenko, A.E.; Thompson, B.C. Optimization of direct arylation polymerization (DArP) through the identification and control of defects in polymer structure. J. Polym. Sci. Part A Polym. Chem. 2014, 53, 135–147. [Google Scholar] [CrossRef]
  21. Chen, S.; Lee, K.C.; Zhang, Z.-G.; Kim, D.S.; Li, Y.; Yang, C. An Indacenodithiophene–Quinoxaline Polymer Prepared by Direct Arylation Polymerization for Organic Photovoltaics. Macromolecules 2016, 49, 527–536. [Google Scholar] [CrossRef]
  22. Homyak, P.; Liu, Y.; Liu, F.; Russel, T.P.; Coughlin, E.B. Systematic Variation of Fluorinated Diketopyrrolopyrrole Low Bandgap Conjugated Polymers: Synthesis by Direct Arylation Polymerization and Characterization and Performance in Organic Photovoltaics and Organic Field-Effect Transistors. Macromolecules 2015, 48, 6978–6986. [Google Scholar] [CrossRef]
  23. Marzano, G.; Kotowski, D.; Babudri, F.; Musio, R.; Pellegrino, A.; Luzzati, S.; Po, R.; Farinola, G.M. Tin-Free Synthesis of a Ternary Random Copolymer for BHJ Solar Cells: Direct (Hetero)arylation versus Stille Polymerization. Macromolecules 2015, 48, 7039–7048. [Google Scholar] [CrossRef]
  24. Pouliot, J.-R.; Sun, B.; Leduc, M.; Najari, A.; Li, Y.; Leclerc, M. A high mobility DPP-based polymer obtained via direct (hetero)arylation polymerization. Polym. Chem. 2015, 6, 278–282. [Google Scholar] [CrossRef]
  25. Rudenko, A.E.; Latif, A.A.; Thompson, B.C. Influence of β-linkages on the morphology and performance of DArP P3HT–PC 61 BM solar cells. Nanotechnology 2014, 25, 014005. [Google Scholar] [CrossRef] [PubMed]
  26. Lombeck, F.; Marx, F.; Strassel, K.; Kunz, S.; Lienert, C.; Komber, H.; Friend, R.; Sommer, M. To branch or not to branch: C-H selectivity of thiophene-based donor-acceptor-donor monomers in direct arylation polycondensation exemplified by PCDTBT. Polym. Chem. 2017, 8, 4738–4745. [Google Scholar] [CrossRef]
  27. Bura, T.; Beaupre, S.; Legare, M.-A.; Quinn, J.; Rochette, E.; Blaskovits, J.T.; Fontaine, F.-G.; Pron, A.; Li, Y.; Leclerc, M. Direct heteroarylation polymerization: Guidelines for defect-free conjugated polymers. Chem. Sci. 2017, 8, 3913–3925. [Google Scholar] [CrossRef] [PubMed]
  28. Pouliot, J.-R.; Grenier, F.; Blaskovits, J.T.; Beaupré, S.; Leclerc, M. Direct (Hetero)arylation Polymerization: Simplicity for Conjugated Polymer Synthesis. Chem. Rev. 2016, 116, 14225–14274. [Google Scholar] [CrossRef] [PubMed]
  29. Kowalski, S.; Allard, S.; Zilberberg, K.; Riedl, T.; Scherf, U. Direct arylation polycondensation as simplified alternative for the synthesis of conjugated (co)polymers. Top. Issue Conduct. Polym. 2013, 38, 1805–1814. [Google Scholar] [CrossRef]
  30. Yu, S.; Liu, F.; Yu, J.; Zhang, S.; Cabanetos, C.; Gao, Y.; Huang, W. Eco-friendly direct (hetero)-arylation polymerization: Scope and limitation. J. Mater. Chem. C 2017, 5, 29–40. [Google Scholar] [CrossRef]
  31. Kleinschmidt, A.T.; Root, S.E.; Lipomi, D.J. Poly(3-hexylthiophene) (P3HT): Fruit fly or outlier in organic solar cell research? J. Mater. Chem. A 2017, 5, 11396–11400. [Google Scholar] [CrossRef]
  32. Sugimoto, R.; Takeda, S.; Gu, H.B.; Yoshino, K. Preparation of Soluble Polythiophene Derivatives Utilizing Transition Metal Halides as Catalysts and Their Property. Chem. Express 1986, 1, 635–638. [Google Scholar]
  33. Mao, H.; Xu, B.; Holdcroft, S. Synthesis and structure-property relationships of regioirregular poly(3-hexylthiophenes). Macromolecules 1993, 26, 1163–1169. [Google Scholar] [CrossRef]
  34. Stein, P.C.; Botta, C.; Bolognesi, A.; Catellani, M. NMR study of the structural defects in poly(3-alkylthiophene)s: Influence of the polymerization method. Proc. Int. Conf. Sci. Technol. Synth. Met. 1995, 69, 305–306. [Google Scholar] [CrossRef]
  35. Leclerc, M.; Martinez, D.F.; Wegner, G. Structural analysis of poly(3-alkylthiophene)s. Makromol. Chem. 2003, 190, 3105–3116. [Google Scholar] [CrossRef]
  36. Sirringhaus, H.; Brown, P.J.; Friend, R.H.; Nielsen, M.M.; Bechgaard, K.; Langeveld-Voss, B.M.W.; Spiering, A.J.H.; Janssen, R.A.J.; Meijer, E.W.; Herwig, P.; et al. Two-dimensional charge transport in self-organized, high-mobility conjugated polymers. Nature 1999, 401, 685. [Google Scholar] [CrossRef]
  37. McCullough, R.D.; Lowe, R.D. Enhanced electrical conductivity in regioselectively synthesized poly(3-alkylthiophenes). J. Chem. Soc. Chem. Commun. 1992, 70–72. [Google Scholar] [CrossRef]
  38. McCullough, R.D.; Lowe, R.D.; Jayaraman, M.; Anderson, D.L. Design, synthesis, and control of conducting polymer architectures: Structurally homogeneous poly(3-alkylthiophenes). J. Org. Chem. 1993, 58, 904–912. [Google Scholar] [CrossRef]
  39. Chen, T.A.; Rieke, R.D. The first regioregular head-to-tail poly(3-hexylthiophene-2,5-diyl) and a regiorandom isopolymer: Nickel versus palladium catalysis of 2(5)-bromo-5(2)-(bromozincio)-3-hexylthiophene polymerization. J. Am. Chem. Soc. 1992, 114, 10087–10088. [Google Scholar] [CrossRef]
  40. Chen, T.-A.; Wu, X.; Rieke, R.D. Regiocontrolled Synthesis of Poly(3-alkylthiophenes) Mediated by Rieke Zinc: Their Characterization and Solid-State Properties. J. Am. Chem. Soc. 1995, 117, 233–244. [Google Scholar] [CrossRef]
  41. Guillerez, S.; Bidan, G. New convenient synthesis of highly regioregular poly(3-octylthiophene) based on the Suzuki coupling reaction. Synth. Met. 1998, 93, 123–126. [Google Scholar] [CrossRef]
  42. Iraqi, A.; Barker, W.G. Synthesis and characterisation of telechelic regioregular head-to-tail poly(3-alkylthiophenes). J. Mater. Chem. 1998, 8, 25–29. [Google Scholar] [CrossRef]
  43. Loewe, R.S.; Khersonsky, S.M.; McCullough, R.D. A Simple Method to Prepare Head-to-Tail Coupled, Regioregular Poly(3-alkylthiophenes) Using Grignard Metathesis. Adv. Mater. 1999, 11, 250–253. [Google Scholar] [CrossRef]
  44. Loewe, R.S.; Ewbank, P.C.; Liu, J.; Zhai, L.; McCullough, R.D. Regioregular, Head-to-Tail Coupled Poly(3-alkylthiophenes) Made Easy by the GRIM Method:  Investigation of the Reaction and the Origin of Regioselectivity. Macromolecules 2001, 34, 4324–4333. [Google Scholar] [CrossRef]
  45. Wang, Q.; Takita, R.; Kikuzaki, Y.; Ozawa, F. Palladium-Catalyzed Dehydrohalogenative Polycondensation of 2-Bromo-3-hexylthiophene: An Efficient Approach to Head-to-Tail Poly(3-hexylthiophene). J. Am. Chem. Soc. 2010, 132, 11420–11421. [Google Scholar] [CrossRef] [PubMed]
  46. Se’vignon, M.; Papillon, J.; Schulz, E.; Lemaire, M. New synthetic method for the polymerization of alkylthiophenes. Tetrahedron Lett. 1999, 40, 5873–5876. [Google Scholar] [CrossRef]
  47. Campeau, L.-C.; Fagnou, K. Palladium-catalyzed direct arylation of simple arenes in synthesis of biaryl molecules. Chem. Commun. 2006, 1253–1264. [Google Scholar] [CrossRef] [PubMed]
  48. Pouliot, J.-R.; Wakioka, M.; Ozawa, F.; Li, Y.; Leclerc, M. Structural Analysis of Poly(3-hexylthiophene) Prepared via Direct Heteroarylation Polymerization. Macromol. Chem. Phys. 2016, 217, 1493–1500. [Google Scholar] [CrossRef]
  49. Yan, H.; Chen, Z.; Zheng, Y.; Newman, C.; Quinn, J.R.; Dotz, F.; Kastler, M.; Facchetti, A. A high-mobility electron-transporting polymer for printed transistors. Nature 2009, 457, 679–686. [Google Scholar] [CrossRef] [PubMed]
  50. Matsidik, R.; Komber, H.; Luzio, A.; Caironi, M.; Sommer, M. Defect-free Naphthalene Diimide Bithiophene Copolymers with Controlled Molar Mass and High Performance via Direct Arylation Polycondensation. J. Am. Chem. Soc. 2015, 137, 6705–6711. [Google Scholar] [CrossRef] [PubMed]
  51. Schuettfort, T.; Thomsen, L.; McNeill, C.R. Observation of a Distinct Surface Molecular Orientation in Films of a High Mobility Conjugated Polymer. J. Am. Chem. Soc. 2013, 135, 1092–1101. [Google Scholar] [CrossRef] [PubMed]
  52. Zhou, N.; Lin, H.; Lou, S.J.; Yu, X.; Guo, P.; Manley, E.F.; Loser, S.; Hartnett, P.; Huang, H.; Wasielewski, M.R.; et al. Morphology-Performance Relationships in High-Efficiency All-Polymer Solar Cells. Adv. Energy Mater. 2013, 4, 1300785. [Google Scholar] [CrossRef]
  53. Deshmukh, K.D.; Qin, T.; Gallaher, J.K.; Liu, A.C.Y.; Gann, E.; O’Donnell, K.; Thomsen, L.; Hodgkiss, J.M.; Watkins, S.E.; McNeill, C.R. Performance, morphology and photophysics of high open-circuit voltage, low band gap all-polymer solar cells. Energy Environ. Sci. 2014, 8, 332–342. [Google Scholar] [CrossRef]
  54. Mu, C.; Liu, P.; Ma, W.; Jiang, K.; Zhao, J.; Zhang, K.; Chen, Z.; Wei, Z.; Yi, Y.; Wang, J.; et al. High-Efficiency All-Polymer Solar Cells Based on a Pair of Crystalline Low-Bandgap Polymers. Adv. Mater. 2014, 26, 7224–7230. [Google Scholar] [CrossRef] [PubMed]
  55. Kang, H.; Uddin, M.A.; Lee, C.; Kim, K.-H.; Nguyen, T.L.; Lee, W.; Li, Y.; Wang, C.; Woo, H.Y.; Kim, B.J. Determining the Role of Polymer Molecular Weight for High-Performance All-Polymer Solar Cells: Its Effect on Polymer Aggregation and Phase Separation. J. Am. Chem. Soc. 2015, 137, 2359–2365. [Google Scholar] [CrossRef] [PubMed]
  56. Fan, B.; Ying, L.; Zhu, P.; Pan, F.; Liu, F.; Chen, J.; Huang, F.; Cao, Y. All-Polymer Solar Cells Based on a Conjugated Polymer Containing Siloxane-Functionalized Side Chains with Efficiency over 10%. Adv. Mater. 2017, 29, 1703906. [Google Scholar] [CrossRef] [PubMed]
  57. Gao, L.; Zhang, Z.; Xue, L.; Min, J.; Zhang, J.; Wei, Z.; Li, Y. All-Polymer Solar Cells Based on Absorption-Complementary Polymer Donor and Acceptor with High Power Conversion Efficiency of 8.27%. Adv. Mater. 2015, 28, 1884–1890. [Google Scholar] [CrossRef] [PubMed]
  58. Guo, X.; Watson, M.D. Conjugated Polymers from Naphthalene Bisimide. Org. Lett. 2008, 10, 5333–5336. [Google Scholar] [CrossRef] [PubMed]
  59. Luzio, A.; Fazzi, D.; Nübling, F.; Matsidik, R.; Straub, A.; Komber, H.; Giussani, E.; Watkins, S.E.; Barbatti, M.; Thiel, W.; et al. Structure–Function Relationships of High-Electron Mobility Naphthalene Diimide Copolymers Prepared Via Direct Arylation. Chem. Mater. 2014, 26, 6233–6240. [Google Scholar] [CrossRef]
  60. Li, Y.; Sonar, P.; Murphy, L.; Hong, W. High mobility diketopyrrolopyrrole (DPP)-based organic semiconductor materials for organic thin film transistors and photovoltaics. Energy Environ. Sci. 2013, 6, 1684–1710. [Google Scholar] [CrossRef]
  61. Li, Y.; Singh, S.P.; Sonar, P. A High Mobility P-Type DPP-Thieno[3,2-b]thiophene Copolymer for Organic Thin-Film Transistors. Adv. Mater. 2010, 22, 4862–4866. [Google Scholar] [CrossRef] [PubMed]
  62. Li, Y.; Sonar, P.; Singh, S.P.; Soh, M.S.; van Meurs, M.; Tan, J. Annealing-Free High-Mobility Diketopyrrolopyrrole−Quaterthiophene Copolymer for Solution-Processed Organic Thin Film Transistors. J. Am. Chem. Soc. 2011, 133, 2198–2204. [Google Scholar] [CrossRef] [PubMed]
  63. Chen, H.; Guo, Y.; Yu, G.; Zhao, Y.; Zhang, J.; Gao, D.; Liu, H.; Liu, Y. Highly π-Extended Copolymers with Diketopyrrolopyrrole Moieties for High-Performance Field-Effect Transistors. Adv. Mater. 2012, 24, 4618–4622. [Google Scholar] [CrossRef] [PubMed]
  64. Kang, I.; Yun, H.-J.; Chung, D.S.; Kwon, S.-K.; Kim, Y.-H. Record High Hole Mobility in Polymer Semiconductors via Side-Chain Engineering. J. Am. Chem. Soc. 2013, 135, 14896–14899. [Google Scholar] [CrossRef] [PubMed]
  65. Gao, Y.; Zhang, X.; Tian, H.; Zhang, J.; Yan, D.; Geng, Y.; Wang, F. High Mobility Ambipolar Diketopyrrolopyrrole-Based Conjugated Polymer Synthesized Via Direct Arylation Polycondensation. Adv. Mater. 2015, 27, 6753–6759. [Google Scholar] [CrossRef] [PubMed]
  66. Broll, S.; Nübling, F.; Luzio, A.; Lentzas, D.; Komber, H.; Caironi, M.; Sommer, M. Defect Analysis of High Electron Mobility Diketopyrrolopyrrole Copolymers Made by Direct Arylation Polycondensation. Macromolecules 2015, 48, 7481–7488. [Google Scholar] [CrossRef]
  67. Guo, C.; Quinn, J.; Sun, B.; Li, Y. Dramatically different charge transport properties of bisthienyl diketopyrrolopyrrole-bithiazole copolymers synthesized via two direct (hetero)arylation polymerization routes. Polym. Chem. 2016, 7, 4515–4524. [Google Scholar] [CrossRef]
  68. Fu, B.; Wang, C.-Y.; Rose, B.D.; Jiang, Y.; Chang, M.; Chu, P.-H.; Yuan, Z.; Fuentes-Hernandez, C.; Kippelen, B.; Brédas, J.-L.; et al. Molecular Engineering of Nonhalogenated Solution-Processable Bithiazole-Based Electron-Transport Polymeric Semiconductors. Chem. Mater. 2015, 27, 2928–2937. [Google Scholar] [CrossRef]
  69. Gobalasingham, N.S.; Pankow, R.M.; Ekiz, S.; Thompson, B.C. Evaluating structure-function relationships toward three-component conjugated polymers via direct arylation polymerization (DArP) for Stille-convergent solar cell performance. J. Mater. Chem. A 2017, 5, 14101–14113. [Google Scholar] [CrossRef]
  70. Kuwabara, J.; Yasuda, T.; Choi, S.J.; Lu, W.; Yamazaki, K.; Kagaya, S.; Han, L.; Kanbara, T. Direct Arylation Polycondensation: A Promising Method for the Synthesis of Highly Pure, High-Molecular-Weight Conjugated Polymers Needed for Improving the Performance of Organic Photovoltaics. Adv. Funct. Mater. 2014, 24, 3226–3233. [Google Scholar] [CrossRef]
  71. Kuwabara, J.; Yasuda, T.; Takase, N.; Kanbara, T. Effects of the Terminal Structure, Purity, and Molecular Weight of an Amorphous Conjugated Polymer on Its Photovoltaic Characteristics. ACS Appl. Mater. Interfaces 2016, 8, 1752–1758. [Google Scholar] [CrossRef] [PubMed]
  72. Bijleveld, J.C.; Gevaerts, V.S.; Di Nuzzo, D.; Turbiez, M.; Mathijssen, S.G.J.; de Leeuw, D.M.; Wienk, M.M.; Janssen, R.A.J. Efficient Solar Cells Based on an Easily Accessible Diketopyrrolopyrrole Polymer. Adv. Mater. 2010, 22, E242–E246. [Google Scholar] [CrossRef] [PubMed]
  73. Hendriks, K.H.; Heintges, G.H.L.; Gevaerts, V.S.; Wienk, M.M.; Janssen, R.A.J. High-Molecular-Weight Regular Alternating Diketopyrrolopyrrole-based Terpolymers for Efficient Organic Solar Cells. Angew. Chem. Int. Ed. 2013, 52, 8341–8344. [Google Scholar] [CrossRef] [PubMed]
  74. Bijleveld, J.C.; Zoombelt, A.P.; Mathijssen, S.G.J.; Wienk, M.M.; Turbiez, M.; de Leeuw, D.M.; Janssen, R.A.J. Poly(diketopyrrolopyrrole−terthiophene) for Ambipolar Logic and Photovoltaics. J. Am. Chem. Soc. 2009, 131, 16616–16617. [Google Scholar] [CrossRef] [PubMed]
  75. Hendriks, K.H.; Li, W.; Heintges, G.H.L.; van Pruissen, G.W.P.; Wienk, M.M.; Janssen, R.A.J. Homocoupling Defects in Diketopyrrolopyrrole-Based Copolymers and Their Effect on Photovoltaic Performance. J. Am. Chem. Soc. 2014, 136, 11128–11133. [Google Scholar] [CrossRef] [PubMed]
  76. Boland, P.; Lee, K.; Namkoong, G. Device optimization in PCPDTBT:PCBM plastic solar cells. Sol. Energy Mater. Sol. Cells 2010, 94, 915–920. [Google Scholar] [CrossRef]
  77. Zhu, Z.; Waller, D.; Gaudiana, R.; Morana, M.; Mühlbacher, D.; Scharber, M.; Brabec, C. Panchromatic Conjugated Polymers Containing Alternating Donor/Acceptor Units for Photovoltaic Applications. Macromolecules 2007, 40, 1981–1986. [Google Scholar] [CrossRef]
  78. Peet, J.; Kim, J.Y.; Coates, N.E.; Ma, W.L.; Moses, D.; Heeger, A.J.; Bazan, G.C. Efficiency enhancement in low-bandgap polymer solar cells by processing with alkane dithiols. Nat. Mater. 2007, 6, 497. [Google Scholar] [CrossRef] [PubMed]
  79. Tsao, H.N.; Cho, D.M.; Park, I.; Hansen, M.R.; Mavrinskiy, A.; Yoon, D.Y.; Graf, R.; Pisula, W.; Spiess, H.W.; Müllen, K. Ultrahigh Mobility in Polymer Field-Effect Transistors by Design. J. Am. Chem. Soc. 2011, 133, 2605–2612. [Google Scholar] [CrossRef] [PubMed]
  80. Chang, S.-W.; Waters, H.; Kettle, J.; Kuo, Z.-R.; Li, C.-H.; Yu, C.-Y.; Horie, M. Pd-Catalysed Direct Arylation Polymerisation for Synthesis of Low-Bandgap Conjugated Polymers and Photovoltaic Performance. Macromol. Rapid Commun. 2012, 33, 1927–1932. [Google Scholar] [CrossRef] [PubMed]
  81. Livi, F.; Gobalasingham, N.S.; Thompson, B.C.; Bundgaard, E. Analysis of diverse direct arylation polymerization (DArP) conditions toward the efficient synthesis of polymers converging with stille polymers in organic solar cells. J. Polym. Sci. Part A Polym. Chem. 2016, 54, 2907–2918. [Google Scholar] [CrossRef]
  82. Iizuka, E.; Wakioka, M.; Ozawa, F. Mixed-Ligand Approach to Palladium-Catalyzed Direct Arylation Polymerization: Effective Prevention of Structural Defects Using Diamines. Macromolecules 2016, 49, 3310–3317. [Google Scholar] [CrossRef]
  83. Dudnik, A.S.; Aldrich, T.J.; Eastham, N.D.; Chang, R.P.H.; Facchetti, A.; Marks, T.J. Tin-Free Direct C–H Arylation Polymerization for High Photovoltaic Efficiency Conjugated Copolymers. J. Am. Chem. Soc. 2016, 138, 15699–15709. [Google Scholar] [CrossRef] [PubMed]
  84. Zhang, S.; Ye, L.; Zhao, W.; Liu, D.; Yao, H.; Hou, J. Side Chain Selection for Designing Highly Efficient Photovoltaic Polymers with 2D-Conjugated Structure. Macromolecules 2014, 47, 4653–4659. [Google Scholar] [CrossRef]
  85. Ye, L.; Zhang, S.; Zhao, W.; Yao, H.; Hou, J. Highly Efficient 2D-Conjugated Benzodithiophene-Based Photovoltaic Polymer with Linear Alkylthio Side Chain. Chem. Mater. 2014, 26, 3603–3605. [Google Scholar] [CrossRef]
  86. Mateker, W.R.; Douglas, J.D.; Cabanetos, C.; Sachs-Quintana, I.T.; Bartelt, J.A.; Hoke, E.T.; El Labban, A.; Beaujuge, P.M.; Frechet, J.M.J.; McGehee, M.D. Improving the long-term stability of PBDTTPD polymer solar cells through material purification aimed at removing organic impurities. Energy Environ. Sci. 2013, 6, 2529–2537. [Google Scholar] [CrossRef] [Green Version]
  87. Marzano, G.; Carulli, F.; Babudri, F.; Pellegrino, A.; Po, R.; Luzzati, S.; Farinola, G.M. PBDTTPD for plastic solar cells via Pd(PPh3)4-catalyzed direct (hetero)arylation polymerization. J. Mater. Chem. A 2016, 4, 17163–17170. [Google Scholar] [CrossRef]
  88. Kim, H.; Lee, B.H.; Lee, K.C.; Kim, G.; Yu, J.Y.; Kim, N.; Lee, S.H.; Lee, K. Role of the Side Chain in the Phase Segregation of Polymer:Fullerene Bulk Heterojunction Composites. Adv. Energy Mater. 2013, 3, 1575–1580. [Google Scholar] [CrossRef]
  89. Najari, A.; Beaupré, S.; Allard, N.; Ouattara, M.; Pouliot, J.-R.; Charest, P.; Besner, S.; Simoneau, M.; Leclerc, M. Thieno, Furo, and Selenopheno[3,4-c]pyrrole-4,6-dione Copolymers: Air-Processed Polymer Solar Cells with Power Conversion Efficiency up to 7.1%. Adv. Energy Mater. 2015, 5, 1501213. [Google Scholar] [CrossRef]
Figure 1. Direct (hetero)arylation polymerization (DHAP) of a 2-bromo-3-alkylthiophene, showing the potential for forming homocoupling and branching defects.
Figure 1. Direct (hetero)arylation polymerization (DHAP) of a 2-bromo-3-alkylthiophene, showing the potential for forming homocoupling and branching defects.
Molecules 23 01255 g001
Figure 2. Synthesis of rr-P3HT by (a) McCullough’s [38], (b) Rieke’s [40], (c) Grignard metathesis (GRIM) [43], (d) Stille coupling [42], (e) Suzuki coupling [41], and (f) DHAP [45] methods.
Figure 2. Synthesis of rr-P3HT by (a) McCullough’s [38], (b) Rieke’s [40], (c) Grignard metathesis (GRIM) [43], (d) Stille coupling [42], (e) Suzuki coupling [41], and (f) DHAP [45] methods.
Molecules 23 01255 g002
Figure 3. Polymers made using DHAP and conventional methods evaluated in organic thin film transistors (OTFTs).
Figure 3. Polymers made using DHAP and conventional methods evaluated in organic thin film transistors (OTFTs).
Molecules 23 01255 g003
Figure 4. Transfer curves for transistors made from N2200 (P2), comparing batches from Stille and DHAP methods. Reproduced with permission from (J. Am. Chem. Soc. 2015, 137, 6705–6711). Copyright (2015) American Chemical Society.
Figure 4. Transfer curves for transistors made from N2200 (P2), comparing batches from Stille and DHAP methods. Reproduced with permission from (J. Am. Chem. Soc. 2015, 137, 6705–6711). Copyright (2015) American Chemical Society.
Molecules 23 01255 g004
Figure 5. Materials made using DHAP and conventional methods, evaluated in organic photovoltaic (OPV) devices.
Figure 5. Materials made using DHAP and conventional methods, evaluated in organic photovoltaic (OPV) devices.
Molecules 23 01255 g005
Figure 6. (a) J-V curves for Stille and DHAP produced PBDTT-FTTE with PC71BM; (b) AFM Images comparing the morphology of Stille and DHAP samples of P6, blended with PC71BM. Reproduced with permission from (J. Am. Chem. Soc. 2016, 138, 15699–15709). Copyright (2016) American Chemical Society.
Figure 6. (a) J-V curves for Stille and DHAP produced PBDTT-FTTE with PC71BM; (b) AFM Images comparing the morphology of Stille and DHAP samples of P6, blended with PC71BM. Reproduced with permission from (J. Am. Chem. Soc. 2016, 138, 15699–15709). Copyright (2016) American Chemical Society.
Molecules 23 01255 g006
Table 1. Comparison of Polymeric Materials Used in Thin Film Transistors.
Table 1. Comparison of Polymeric Materials Used in Thin Film Transistors.
SynthesisOTFT Performance
IDNameMn, kDa (HT mol %)Mw/MnYield, %Deviceuh/ue, cm2 V−1 s−1Ion/IoffVth, VRef.
P1DHAP1P3HT33 (>99.5)1.896BGBC0.19/-1000/--/-2016 [48]
P1GRIMP3HT88 (98.0)1.5NABGBC0.11/-1000/--/-2016 [48]
P1RiekeP3HT25 (95.5)1.9NABGBC0.02/-1000/--/-2016 [48]
P2DHAPN2200312.999TGBC-/2.9-/>1000-/-2015 [50]
P2StilleN2200325.4100TGBC-/3.2-/>1000-/-2015 [50]
P3DHAPP(ThNDIThF4)7.81.7-TGBC-/1.3-/~105-/-2014 [59]
P4DHAPPDPP-4FTVT604.993BGTC3.4/5.9>105/>10−1~−15/40~552015 [65]
P5DHAPPDPPTh2F4302.475TGBC-/0.60-/~104-/24.52015 [66]
P6DHAPPDBTz-24183.866TGBC0.06/0.53~106/~105-/-2016 [67]
P6StillePDBTz-27643.690BGTC-/0.31-/105-/42015 [68]
P7DHAPPDPP462.584BGBC1.2/-~103/-0/-2015 [24]
Table 2. Comparison of Polymeric Materials Used in Solar Cells.
Table 2. Comparison of Polymeric Materials Used in Solar Cells.
SynthesisSolar Cells
IDNameMn, kDa (HT %)Mw/MnYield, %AcceptorPCE (Best Reported)JSC, mA cm−2VOC, VFF, %Ref.
P1DHAP2 aP3HT19 (90.0)2.0-PC61BM2.70 8.700.6250.02013 [25]
P1Stille1 bP3HT19 (93.0)2.7-PC61BM2.30 8.370.6145.02013 [25]
P1DHAP3 aP3HT20 (96.2)2.174PC61BM3.28 9.400.5959.12017 [69]
P1Stille2 bP3HT18 (92.9)2.468PC61BM2.86 8.340.5958.12017 [69]
P8DHAPPEDOTF1502.8989PC71BM4.089.410.8352.02014 [70]
P8SuzukiPEDOTF172.0885PC71BM0.482.580.5931.02014 [70]
P9DHAPPDPP-TPT141.829PC71BM4.3713.30.7741.52015 [22]
P9Suzuki1 cPDPP-TPT65 a-69PC71BM5.5010.80.8065.02010 [72]
P9Suzuki2 cPDPP-TPT721.9893PC71BM7.4014.00.8067.02013 [73]
P10DHAPPDPP-3T293.845PC71BM4.0110.30.7156.22015 [22]
P10SuzukiPDPP-3T543.1584PC71BM4.6911.80.6660.02009 [74]
P10StillePDPP-3T1502.7285PC71BM7.1015.40.6769.02013 [73]
P11DHAPPCPDTBT724.5276PC71BM3.9813.90.6345.52012 [80]
P11SuzukiPCPDTBT152.183PC71BM3.7412.70.6443.82012 [80]
P11StillePCPDTBT281.561PC61BM3.5011.80.6546.02007 [77]
P12DHAPPPDTBT152.178PC61BM3.4010.50.7245.02016 [81]
P12Stille1-HMW dPPDTBT593.379PC61BM3.8011.50.7345.02016 [81]
P12Stille2-LMW dPPDTBT162.170PC61BM2.908.880.7246.02016 [81]
P13DHAP-107.670PC71BM2.805.580.8956.02015 [23]
P13Stille-203.185PC71BM4.809.890.8160.02015 [23]
P14DHAPPBDTT-FTTE252.298PC71BM8.3615.50.7868.82016 [83]
P14StillePBDTT-FTTE252.2-PC71BM8.4014.90.7872.22016 [83]
P15DHAPPBDTT-TPD302.776PC71BM5.8410.00.9957.92016 [83]
P15StillePBDTT-TPD152.469PC71BM5.209.100.9958.72016 [83]
P16DHAPPTPD3T192.083PC71BM7.2013.30.8266.02016 [83]
P16StillePTPD3T301.894PC71BM7.3813.20.7871.12016 [83]
P17DHAPPBDTTPD12-80PC71BM5.3110.410.9256.02016 [87]
P17StillePBDTTPD10-90PC71BM4.829.170.9357.02016 [87]
P18DHAP-251.982PC71BM6.8013.8 0.91 53.52016 [19]
P18Stille-151.252PC71BM5.2010.00.8859.02013 [88]
P19DHAPPSePD3T291.657PC61BM7.1313.20.8564.02015 [89]
P20DHAPIDT-TQ271.671PC71BM5.1010.80.8953.42016 [21]
P20StilleIDT-TQ231.564PC71BM4.8210.40.8952.12016 [21]
a Different synthetic conditions were used. P1DHAP2: 1 mol % Pd(OAc)2/pivalic acid/K2CO3/DMAc/45 °C, 72 h; P1DHAP3: 1 mol % Pd2(dba)3/P(o-MeOPh)3/neodecanoic acid/Cs2CO3/MeTHF/120 °C, 12 h; b Same synthetic conditions (Pd(PPh3)4/DMF, 95 °C, 48 h) were used. Similar device fabrication procedures, except for the annealing conditions (110 °C /40 min for P1Stille1 vs. 150 °C/30 min for P1Stille2), were used; c Different synthetic conditions: P9Suzuki1: 5 mol % Pd2(dba)3/11 mol % PPh3/4 eq. K3PO4/Aliquat 336, toluene-water (8:1)/115 °C, 72 h; P9Suzuki2: 3 mol % Pd2(dba)3/12 mol % PPh3/5 eq. K3PO4/Aliquat 336, toluene-water (6:1), 115 °C, 16 h; d Different monomers were used: P12StilleLMW: (5,5′-(2,5-bis((2-hexyldecyl)oxy)-1,4-phenylene)bis(thiophene-5,2-diyl))bis(trimethylstannane) and 4,7-dibromobenzo[c][1,2,5]thiadiazole; P12StilleHMW: 4,7-bis(5-(trimethylstannyl)thiophen-2-yl)benzo[c][1,2,5]thiadiazole and 1,4-dibromo-2,5-bis((2-hexyldecyl)oxy)benzene; e A bimodal GPC trace shows two peak molecular weights at 65 kDa and 10 kDa, respectively.

Share and Cite

MDPI and ACS Style

Hendsbee, A.D.; Li, Y. Performance Comparisons of Polymer Semiconductors Synthesized by Direct (Hetero)Arylation Polymerization (DHAP) and Conventional Methods for Organic Thin Film Transistors and Organic Photovoltaics. Molecules 2018, 23, 1255. https://doi.org/10.3390/molecules23061255

AMA Style

Hendsbee AD, Li Y. Performance Comparisons of Polymer Semiconductors Synthesized by Direct (Hetero)Arylation Polymerization (DHAP) and Conventional Methods for Organic Thin Film Transistors and Organic Photovoltaics. Molecules. 2018; 23(6):1255. https://doi.org/10.3390/molecules23061255

Chicago/Turabian Style

Hendsbee, Arthur D., and Yuning Li. 2018. "Performance Comparisons of Polymer Semiconductors Synthesized by Direct (Hetero)Arylation Polymerization (DHAP) and Conventional Methods for Organic Thin Film Transistors and Organic Photovoltaics" Molecules 23, no. 6: 1255. https://doi.org/10.3390/molecules23061255

Article Metrics

Back to TopTop