Charge Mobility Enhancement for Conjugated DPP ... - ACS Publications

Apr 17, 2018 - Henning Sirringhaus,. § and Deqing Zhang*,†,‡. †. Beijing National Laboratories for Molecular Sciences, CAS Key Laboratories of ...
0 downloads 0 Views 4MB Size
Article Cite This: Chem. Mater. 2018, 30, 3090−3100

pubs.acs.org/cm

Charge Mobility Enhancement for Conjugated DPP-Selenophene Polymer by Simply Replacing One Bulky Branching Alkyl Chain with Linear One at Each DPP Unit Zhijie Wang,†,‡ Zitong Liu,*,† Lu Ning,†,‡ Mingfei Xiao,§ Yuanping Yi,†,‡ Zhengxu Cai,∥ Aditya Sadhanala,§ Guanxin Zhang,† Wei Chen,⊥,# Henning Sirringhaus,§ and Deqing Zhang*,†,‡

Downloaded via UNIV OF CALIFORNIA SANTA BARBARA on July 1, 2018 at 15:54:55 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.



Beijing National Laboratories for Molecular Sciences, CAS Key Laboratories of Organic Solids, CAS Research/Education Center for Excellence in Molecular Sciences, Institute of Chemistry, Chinese Academy of Sciences, Beijing 100190, P. R. China ‡ University of Chinese Academy of Sciences, Beijing 100049, P. R. China § Cavendish Laboratory, University of Cambridge, J. J. Thomson Avenue, Cambridge CB3 0HE, U.K. ∥ Beijing Key Laboratory of Construction Tailorable Advanced Functional Materials and Green Applications, School of Material Science & Engineering, Beijing Institute of Technology, Beijing 100081, China ⊥ Institute for Molecular Engineering and Materials Science Division, Argonne National Laboratory, 9700 Cass Avenue, Lemont, Illinois 60439, United States # Institute for Molecular Engineering, The University of Chicago, 5640 South Ellis Avenue, Chicago, Illinois 60637, United States S Supporting Information *

ABSTRACT: We demonstrate a simple, but efficient, approach for improving the semiconducting performances of DPP-based conjugated D-A polymers. This approach involves the replacement of one bulky branching alkyl chain with the linear one at each DPP unit in regular polymer PDPPSe-10 and PDPPSe-12. The UV−vis absorption, Raman spectra, PDS data, and theoretical calculations support that the replacement of bulky branching chains with linear ones can weaken the steric hindrance, and accordingly conjugated backbones become more planar and rigid. GIWAXS data show that the incorporation of linear alkyl chains as in PDPPSe-10 and PDPPSe-12 is beneficial for side-chain interdigitation and interchain dense packing, leading to improvement of interchain packing order and thin film crystallinity by comparing with PDPPSe, which contains branching alkyl chains. On the basis of field-effect transistor (FET) studies, charge mobilities of PDPPSe-10 and PDPPSe-12 are remarkably enhanced. Hole mobilities of PDPPSe-10 and PDPPSe-12 in air are boosted to 8.1 and 9.4 cm2 V−1 s−1, which are about 6 and 7 times, respectively, than that of PDPPSe (1.35 cm2 V−1 s−1). Furthermore, both PDPPSe-10 and PDPPSe-12 behave as ambipolar semiconductors under a nitrogen atmosphere with increased hole/electron mobilities up to 6.5/0.48 cm2 V−1 s−1 and 7.9/0.79 cm2 V−1 s−1, respectively.



crowding of branching alkyl chains.7−9 Electron donor units with different conjugation lengths such as thiophene, bithiophene, terthiophene, and quaterthiophene (Scheme 1a) were incorporated into DPP-based conjugated D-A polymers.24−27 The results reveal that long electron donors not only strengthen the interchain π−π interactions but also are able to separate the neighboring branching alkyl chains and thus weaken the steric hindrances, resulting in enhancement of charge mobilities for conjugated polymers.24−27 Alternatively, new branching alkyl chains and siloxaneterminated side chains (Scheme 1b) were devised and connected to DPP, isoindigo, and other acceptors.28−33 Because the branching points of these side chains are away

INTRODUCTION Conjugated donor (D)-acceptor (A) polymers have been extensively investigated for solution-processable semiconductors with high charge mobilities which are highly demanding for applications in E-paper, wearable electronics, and other devices.1−6 Diketopyrrolopyrrole (DPP) with two flanking thiophenes has been widely utilized as electron acceptor to connect with various electron donors to form the respective conjugated D-A polymers.7−30 Two alkyl groups can be incorporated into the DPP unit, which will endow the DPPbased conjugated D-A polymers with good solubilities in organic solvents.7 In order to enhance the solubilities of these DPP-based polymers, two bulky branching alkyl groups are usually connected to each DPP unit in the polymers.8,9 The presence of bulky branching groups along the polymer backbone can affect the planarity of the conjugated backbone and prevent dense packing of polymer chains due to steric © 2018 American Chemical Society

Received: March 7, 2018 Revised: April 17, 2018 Published: April 17, 2018 3090

DOI: 10.1021/acs.chemmater.8b01007 Chem. Mater. 2018, 30, 3090−3100

Article

Chemistry of Materials Scheme 1. Chemical Structures of DPP-Based Conjugated Polymersa

a

(a) Containing electron donor units with increasing lengths to weaken the steric hindrances; (b) containing branching bulky alkyl chains that are away from polymer backbones; (c) containing both branching bulky and linear alkyl chains in a random manner.

and 6 in Scheme 2) were easily synthesized with 1bromodecane or 1-bromododecane and 11-(bromomethyl)tricosane, all of which are commercially available. We report two conjugated DPP-selenophene polymers PDPPSe-10 and PDPPSe-12 (Scheme 2) which contain n-decyl and n-dodecyl as the respective linear alkyl chains linked to each DPP unit. For comparison, PDPPSe, which possesses the same con-

from the polymer backbones, the steric hindrance due to alkyl chains can be reduced, and as a result, the polymer can be more orderly packed with short π−π stacking distance. Such sidechain engineering has significantly boosted charge mobilities for conjugated D-A polymers with DPP17,28 and isoindigo31,32 as electron acceptors. However, the respective alkyl halogens, precursors to these modified side chains, are not commercially available, and their syntheses involve three or more reaction steps and purifications.34−36 This will surely add additional cost for the scalable preparation of these conjugated D-A polymers for future device applications. Some of us have recently discovered that charge mobilities of DPP-based terpolymers can be enhanced by just partial replacement of bulky branching alkyl chains with linear ones which are expected to reduce the steric hindrance and promote interdigitation of linear alkyl chains (Scheme 1c).37 These terpolymers were prepared by copolymerization of the 5,5′bis(trimethylstannyl)-2,2′-bithiophene and two DPP monomers, one of which entails two branching alkyl chains and the other contains two linear alkyl chains. Although these two DPP monomers can be easily prepared with the commercially available 11-(bromomethyl)tricosane and 1-bromododecane, branching and linear alkyl chains are randomly arranged within the resulting terpolymers, and thus batch-to-batch variation of polymer samples cannot be avoided. In this paper, we report a new and simple approach to reduce steric hindrance and thus remarkably enhance charge mobilities for DPP-based conjugated D-A polymers by replacing one bulky branching alkyl chain with the linear one at each DPP unit. The DPP monomers flanked with thiophene moieties (5

Scheme 2. Chemical Structures of PDPPSe-10, PDPPSe-12, and PDPPSe, and Synthetic Routes to PDPPSe-10 and PDPPSe-12

3091

DOI: 10.1021/acs.chemmater.8b01007 Chem. Mater. 2018, 30, 3090−3100

Article

Chemistry of Materials

Figure 1. (a) Distribution of the dihedral angles between the thiophene and selenophene units (S-C-C-Se) and between the thiophene and DPP units (S-C-C-N) in the polymer conjugated backbones obtained by molecular dynamics simulations. (b) Illustration of the corresponding dihedral angles and conceived configurations of the polymer backbones according to the simulated results.

Se dihedral angles in Figure 1) and between the thiophene and DPP units (the S−C−C−N dihedral angles) are employed to gauge the planarity of the polymer backbones. The obtained distributions of the twist angles between different conjugated units are shown in Figure 1. For PDPPSe-12, the S−C−C−N dihedral angles (between the thiophene and DPP units) are in the range of 0−47° with 18° at the maximum distribution, while the S−C−C−Se dihedral angles (between the thiophene and selenophene units) are in the range of 127−180° with 158° at the maximum distribution. In comparison, the S−C−C−N and S−C−C−Se dihedral angles are in the range of 108−180° with 143° at the maximum distribution and 107−180° with 153° at the maximum distribution, respectively, for PDPPSe. These simulation results indicate that the twists between these conjugated units are substantially smaller for PDPPSe-12 with respect to PDPPSe. That is to say, replacement of the branched alkyl groups with straight linear alkyl groups will result in better planarity for the backbone of PDPPSe-12, which is beneficial for increasing interchain π−π interaction for the polymer. In addition, as shown in Figure 1, as the S−C−C−Se dihedral angles are larger than 90° for both PDPPSe-12 and PDPPSe, the thiophene and selenophene units tend to adopt a “trans” conformation for both polymers. The thiophene and DPP units prefer to adopt a “cis” conformation for PDPPSe-12 because the S−C−C−N dihedral angles are smaller than 90°. In comparison, the S−C−C−N dihedral angles are larger than 90° for PDPPSe, and the thiophene and DPP units are in a “trans” conformation. Synthesis and Characterization. The synthesis of PDPPSe-10 and PDPPSe-12 is outlined in Scheme 2. The synthesis of monomers 3 started from compound 1 which reacted with 11-(bromomethyl)tricosane and 1-bromodecane sequentially to afford 3 in acceptable yield. Similarly, compound

jugated backbone as for PDPPSe-10 and PDPPSe-12 with two bulky branching chains, was also prepared. The results reveal that (i) conjugated backbones of PDPPSe-10 and PDPPSe-12 become more planar in comparison with that of PDPPSe based on the UV−vis absorption, Raman spectra, photothermal deflection (PDS), and theoretical calculations; (ii) polymer chains adopt edge-on orientations and are more orderly packed within thin films of PDPPSe-10 and PDPPSe-12 based on the GIWAXS data; and (iii) hole mobilities are boosted to 8.1 and 9.4 cm2 V−1 s−1 in air for PDPPSe-10 and PDPPSe-12, respectively. It is also noted that both PDPPSe-10 and PDPPSe-12 exhibit ambipolar semiconducting behavior under a N2 atmosphere, and both hole and electron mobilities are remarkably increased to 6.5/0.48 cm2 V−1 s−1 and 7.9/0.79 cm2 V−1 s−1, in comparison with those of PDPPSe (1.18/0.19 cm2 V−1 s−1). It is noted that DPP units with two different side chains have been reported and introduced into small molecules and polymers.38−43 However, these side chains entail functional group such as oligo(ethylene glycol) chains,38−40 siloxane terminated chains,40 perylene bisimide terminated chains,41 and aryl group terminated chains.42,43 Most of these DPPcontaining conjugated polymers have been employed for organic photovoltaic cells.41,42



RESULTS AND DISCUSSION

Design Rationale Based on Theoretical Calculations. In order to support our design rationale and reveal the influence of side alkyl chains on the structures of the backbones of conjugated D-A polymers, we carried out molecular dynamics simulations (see the Supporting Information for simulation details).44−46 To be simplified, five-repeated units of PDPPSe-12 and PDPPSe were calculated. The twist angles between the thiophene and selenophene units (the S−C−C− 3092

DOI: 10.1021/acs.chemmater.8b01007 Chem. Mater. 2018, 30, 3090−3100

Article

Chemistry of Materials

were detected for thin films of PDPPSe-10 and PDPPSe-12. Similarly, absorption spectra of thin films of PDPPSe-10 and PDPPSe-12 are red-shifted by comparing that of PDPPSe. Such absorption spectral red shifts can be induced by (i) the fact that the conjugated backbones of PDPPSe-10 and PDPPSe-12 become more planar than PDPPSe after replacing one branching chain with the linear one for each DPP unit; and (ii) the interchain π−π stacking within thin films of PDPPSe-10 and PDPPSe-12 and their preaggregates in solutions.47,48 The preaggregation effect in solution can be excluded by either heating the solution or lowering the concentration. As shown in Figure S6, the solution absorption spectra of PDPPSe-10 and PDPPSe-12 as well as PDPPSe were blue-shifted when the respective solutions (1 × 10−5 mol/L) were heated to 100 °C. Moreover, the absorption spectra of these three polymers at 100 °C were almost identical with the respective ones when their concentrations were decreased to 1 × 10−6 mol/L. Thus, the preaggregates formed in solutions of these polymers can be disassembled upon heating. The absorption maxima for solutions (1 × 10−5 mol/L) at 100 °C PDPPSe-10 and PDPPSe-12 appear at 848 and 846 nm, which are red-shifted in comparison with that of PDPPSe (832 nm). Such spectral red shifts imply that the conjugated backbones of PDPPSe-10 and PDPPSe-12 are more planar than PDPPSe. In fact, more planar conjugated backbones facilitate the interchain π−π interactions and thus the aggregation of polymer chains.49 This is well consistent with the observation that the red shifts of absorption maxima for PDPPSe-10 (22 nm) and PDPPSe-12 (22 nm) are larger than that of PDPPSe (12 nm) when the respective solutions were cooled from 100 °C to room temperature. HOMO/LUMO levels of PDPPSe-10 and PDPPSe-12 as well as PDPPSe were estimated with the respective onset oxidation and reduction potentials, which were determined with their cyclic voltammograms in the form of thin films (Figure S7). HOMO levels of PDPPSe-10 (−5.20 eV) and PDPPSe-12 (−5.21 eV) are slightly higher than that of PDPPSe (−5.28 eV). The LUMO level of PDPPSe-10 (−3.64 eV) is the same as PDPPSe (−3.64 eV), while the LUMO level of PDPPSe-12 (−3.66 eV) becomes slightly lower. The band gaps of PDPPSe-10 (1.56 eV) and PDPPSe-12 (1.55 eV) based on their HOMO/LUMO levels are lower than that of PDPPSe (1.64 eV). This is consistent with the fact that conjugated backbones of PDPPSe-10 and PDPPSe-12 are more planar than that of PDPPSe on the basis of their absorption spectral studies and theoretical calculations as discussed above. Raman and Photothermal Deflection Spectra. In order to provide further support for the fact that the conjugated backbones of PDPPSe-10 and PDPPSe-12 are more planar than that of PDPPSe, their Raman spectra were carefully examined. Raman spectral data support the conclusion that the backbone planarity is improved for PDPPSe-10 and PDPPSe-12. Figure S8 shows the Raman spectra of as-prepared and thermally annealed thin films of PDPPSe-10, PDPPSe-12, and PDPPSe upon excitation at 532 nm. These thin films were prepared in the same way as for FET (field-effect transistor) studies. The Raman shifts around 1366 and 1420 cm−1 correspond to the intraunit and interunit C−C, C−N stretches involving the DPP unit, while the Raman shift around 1446 cm−1 is attributed to the intraunit backbone CC symmetric stretch according to a previous report.50 As shown in Figure S8, the intensities of Raman shifts at 1446 cm−1 are higher for thin films of PDPPSe10 and PDPPSe-12 than that for PDPPSe before and after thermal annealing, indicating that π-electrons are more

4 was yielded. Compounds 3 and 4 were subjected to bromination with NBS, leading to 5 and 6 in good yields.43 Stille polycondensation of 2,5-bis(trimethylstannyl)selenophene with monomers 5 and 6, respectively, yielded PDPPSe-10 and PDPPSe-12 in high yields, after precipitation and Soxhlet extraction (see the Experimental Section). The reference polymer PDPPSe was prepared according to the reported procedure.39 The chemical structures of these polymers were verified by 1H NMR, solid-state 13C NMR, and elemental analysis (see the Experimental Section). PDPPSe-10 and PDPPSe-12 possess similar solubility as PDPPSe in o-dichlorobenzene, 1,1,2,2-tetrachloroethane, and other halogenated solvents, which enables them to be solution processable for fabrication of field effect transistors (FETs). Molecular weights (Mn) of PDPPSe-10, PDPPSe-12, and PDPPSe were measured to be 55.1, 60.7, and 76.6 kDa, with PDI of 2.52, 2.60, and 2.76, respectively, with gel permeation chromatography. As shown in Figure S4 of thermogravimetric analysis (TGA) curves, the temperatures at 5% weight loss were 393 °C for PDPPSe-10, 397 °C for PDPPSe-12, and 393 °C for PDPPSe. These illustrate that all polymers are thermally stable below 300 °C. DSC curves in Figure S5 show that there were no thermal transitions below 250 °C for PDPPSe-10, PDPPSe12 as well as PDPPSe. Absorption Spectra and HOMO/LUMO Levels. Figure 2 shows the solution absorption spectra of PDPPSe-10 and

Figure 2. Normalized absorption spectra of PDPPSe-10, PDPPSe-12, and PDPPSe in 1,2-dichlorobenzene solution (1.0 × 10−5 M) and their thin films on a quartz plate.

PDPPSe-12 as well as PDPPSe and their thin films, and the absorption spectral data are listed in Table 1. In comparison with that of PDPPSe, the absorption maxima of PDPPSe-10 and PDPPSe-12 in solutions are red-shifted by 26 and 24 nm, respectively. The 0-0/0-1 absorptions around 884 and 808 nm 3093

DOI: 10.1021/acs.chemmater.8b01007 Chem. Mater. 2018, 30, 3090−3100

Article

Chemistry of Materials

Table 1. Molecular Weights, Absorption Maxima, Redox Potentials, HOMO/LUMO Energies, and Band Gaps of PDPPSe-10, PDPPSe-12, and PDPPSe λmaxa (nm) (εmax, M−1 cm−1) polymer

Mn (kDa)/PDI

solution

film

Eredonset (V)b

ELUMO (eV)c

Eoxonset (V)b

EHOMO (eV)c

Eg,cv (eV)d

Eg,opt (eV)e

PDPPSe-10

55.1/2.52

−3.64

0.40

−5.20

1.56

1.27

60.7/2.60

−1.14

−3.66

0.41

−5.21

1.55

1.26

PDPPSe

76.6/2.76

808 884 808 884 800 868

−1.16

PDPPSe-12

870 (68000) 868 (66000) 844 (61000)

−1.16

−3.64

0.48

−5.28

1.64

1.30

Absorption maxima in 1,2-dichlorobenzene solution (1.0 × 10−5 M) and thin film. bOnset potentials (V vs Fc/Fc+) for reduction (Eredonset) and oxidation (Eoxonset). cHOMO and LUMO energies were estimated with the following equations: EHOMO = −(Eoxonset + 4.8) eV, ELUMO = −(Eredonset + 4.8) eV. dEg,cv = ELUMO − EHOMO (eV). eBased on thin film onset absorption. a

Figure 3. GIWAXS patterns of PDPPSe (a), PDPPSe-10 (b), and PDPPSe-12 (c) deposited on OTS-modified SiO2/Si substrates after thermal annealing at 180 °C. Schematic illustration for stacking orientation of PDPPSe (d, face-on) and PDPPSe-10, PDPPSe-12 (e, edge-on).

technique, and they were subjected to GIWAXS (grazing incidence wide-angle X-ray scattering) analysis.54 Figure S10 and Figure 3 show 2D GIWAXS patterns for the as-prepared thin films of PDPPSe, PDPPSe-10, and PDPPSe-12 and those after thermal annealing at 180 °C. The respective scattering peaks for the lamellar stacking of alkyl chains and interchain π−π stacking are collected in Table S1. In order to investigate the ratios of the edge-on and face-on populations in the thin films before and after thermal annealing at 180 °C, pole figures for the (100) scattering signals were constructed and are shown in Figure S11, and the area ratios of the two regions (χ = 0−45° and 45−90°) were calculated according to previous reports.55,56 The edge-on/face-on ratios were calculated to be 63.4/36.6, 96.5/3.5, and 96.5/3.5 for the as-prepared thin films of PDPPSe, PDPPSe-10, and PDPPSe12, respectively. The respective edge-on/face-on ratios were changed to 72.3/27.3, 97.7/2.3, and 97.8/2.2 after thermal annealing at 180 °C. These results demonstrate that polymer chains of PDPPSe-10 and PDPPSe-12 adopt mainly the edge-on packing mode on the substrate, whereas polymer chains of PDPPSe take both edge-on and face-on orientations. The edge-on contents were slightly enhanced after thermal annealing at 180 °C for all three polymers. As expected, more scattering signals were observed and signal intensities were enhanced for PDPPSe-10 and PDPPSe-12 as well as PDPPSe after thermal annealing at 180 °C (see Figure 3). The thermally annealed thin films of PDPPSe-10 and

delocalized across the thiophene and selenophene units within the conjugated backbones of PDPPSe-10 and PDPPSe-12.51 In addition, photothermal deflection spectra (PDS) of the spin-coated thin films of PDPPSe-10, PDPPSe-12, and PDPPSe were measured52,53 and are shown in Figure S9. On the basis of the PDS data, the respective Urbach energies were estimated to be 33, 33.5, and 36.5 meV for PDPPSe-10, PDPPSe-12, and PDPPSe, respectively, according to the literature reported method.52,53 The Urbach energies of PDPPSe-10 and PDPPSe12 are slightly smaller than that of PDPPSe. This is a significant difference, as the error in extracting the Urbach energy from our measurements is on the order of 5%. Thus, these PDS data demonstrate that PDPPSe-10 and PDPPSe-12 show a comparatively lower degree of energetic disorder than PDPPSe. It is known that more planar and rigid backbones of conjugated polymers are expected to possess low Urbach energies according to a previous report.52 Thus, the comparison of Urbach energies among PDPPSe, PDPPSe-10, and PDPPSe-12 hints that torsion angles among the conjugated moieties are reduced for PDPPSe-10 and PDPPSe-12 after replacing one branching alkyl chain with the linear one at each DPP unit. Interchain Packing and Thin Film Morphology. In the following, we show the significant effect of the replacement of one bulky branching chain with the linear one at each DPP unit on the interchain packing for thin films of PDPPSe-10 and PDPPSe-12. Thin films of PDPPSe-10 and PDPPSe-12 as well as PDPPSe were prepared with a conventional spin-coating 3094

DOI: 10.1021/acs.chemmater.8b01007 Chem. Mater. 2018, 30, 3090−3100

Article

Chemistry of Materials

thin film morphology in which nanofibers were interconnected as in the thin film of PDPPSe. But, the root-mean-square roughnesses were found to be higher for thin films of PDPPSe10 and PDPPSe-12 than that for the thin film of PDPPSe before and after thermal annealing. For instance, the rootmean-square roughnesses were measured to be 1.19, 1.24, and 0.81 nm for PDPPSe-10, PDPPSe-12, and PDPPSe, respectively, after thermal annealing at 180 °C. This may be attributed to the formation of large domains of polymer chains within thin films of PDPPSe-10 and PDPPSe-12. To conclude, the replacement of one bulky branching side chain with the linear one at each DPP unit in PDPPSe-10 and PDPPSe-12 can weaken the steric hindrance owing to branching alkyl chains in PDPPSe as illustrated in Scheme 3. This results in more orderly interchain packing and improvement of thin film crystallinity, while thin film morphology with interconnected nanofibers is kept. As illustrated in Figure 4, this simple side-chain engineering induces the interchain packing mode on the substrate to be changed from face-on to edge-on and increases the correlation lengths (LC) along the π−π stacking directions. As it will be discussed below, these structural alterations with regard to interchain packing are beneficial for improving the charge mobilities of conjugated DA polymers. Charge Mobility Enhancement. In the following, we show the remarkable enhancement of charge mobilities for thin films of PDPPSe-10 and PDPPSe-12, which is expected on the basis of the improvement of interchain packing order and thin film crystallinity. Charge mobilities of thin films of PDPPSe-10 and PDPPSe-12 as well as PDPPSe were extracted from the transfer curves of bottom-gate/bottom-contact (BGBC) fieldeffect transistors (FETs), which were fabricated with conventional procedures (see the Supporting Information) by using the respective polymer thin films as the semiconducting layers. Figure 5 shows the transfer and output curves, which were measured in air, for the respective FETs with thin films of PDPPSe-10, PDPPSe-12, and PDPPSe. As expected, all three polymers behave as p-type semiconductors. As listed in Table 2, FETs with the as-prepared thin films of PDPPSe-10 and PDPPSe-12 show similar Ion/Ioff and VTh as that with thin film of PDPPSe. However, hole mobilities of as-prepared thin films of PDPPSe-10 (2.2 cm2 V−1 s−1) and PDPPSe-12 (3.1 cm2 V−1 s−1) are ca. 5/6 times higher than that of PDPPSe (0.48 cm2 V−1 s−1). As expected, thermal annealing can enhance hole mobilities for these three polymers. The maxima/average hole mobilities can reach 8.1/5.8 cm2 V−1 s−1 and 9.4/7.5 cm2 V−1 s−1 for thin films of PDPPSe-10 and PDPPSe-12 after thermal annealing at 180 °C. In comparison, the maxima/average hole mobilities were measured to be 1.35/0.89 cm2 V−1 s−1 for the thermally annealed thin film of PDPPSe under the same condition. We note that the transfer characteristics are not ideal and typically exhibit a region of higher slope at low gate voltage and a region of smaller slope at high gate voltage. In fact, such nonideal transfer curves were observed for FETs with a number of conjugated D-A polymers,1,59 but its origins are not clearly understood.59 Therefore, we calculated the reliability factors of FETs measured in air according to the reported procedure (for details, see Figure S14).60 The average reliability factors are calculated to be 52%, 51%, and 56% for the respective 10 devices with PDPPSe-10, PDPPSe-12, and PDPPSe. Clearly, their reliability factors are similar. Therefore, comparison of their charge mobilities should be acceptable. FETs with

PDPPSe-12 show more scattering signals with higher intensities in comparison with those of PDPPSe. Therefore, it can be concluded that thin films of PDPPSe-10 and PDPPSe-12 show better crystallinities than that of PDPPSe. As shown in Figure 3 and Table S1, the thermally annealed thin film of PDPPSe-12 shows four orders of the out-of-plane scattering signals that result from the lamellar packing of side chains at qz = 0.29, 0.57, 0.84, and 1.13 Å−1, corresponding to a d-spacing of 21.9 Å. Similarly, the d-spacing due to the lamellar packing of alkyl chains was determined to be 21.2 Å based on the out-of-plane scattering signals at 0.30, 0.58, 0.87, 1.16 Å−1 for the thermally annealed thin film of PDPPSe-10. In comparison, the thin film of PDPPSe after thermal annealing shows only two orders of scattering signals in the out-of-plane direction with a d-spacing of 23.5 Å, and moreover their intensities were much weaker than those of PDPPSe-10 and PDPPSe-12. Clearly, the dspacing owing to the lamellar packing of side alkyl chains is shortened for PDPPSe-10 and PDPPSe-12. This can be attributed to the interdigitation of side alkyl chains which is facilitated by replacing the branching chains with linear ones as in PDPPSe-10 and PDPPSe-12. The thermally annealed thin films of PDPPSe-10 and PDPPSe-12 also exhibit in-plane scattering signals at 1.69 and 1.71 Å−1 (see Figure 3 and Table S1), which result from the respective interchain π−π stacking with packing distances of 3.69 and 3.67 Å. However, the thin film of PDPPSe shows a diffused out-of-plane diffraction arc at qz = 1.71 Å−1, corresponding to a π−π stacking distance of 3.68 Å. Furthermore, on the basis of the full width at hall-maximum (fwhm) of scattering signals due to π−π stacking (see Figure S12), the correlation lengths (LC) along the π−π stacking directions were estimated to be 2.26, 3.74, and 3.76 nm for PDPPSe, PDPPSe-10, and PDPPSe-12, respectively. It is known that LC is the distance over which crystalline order is preserved.57,58 Hence, thin films of PDPPSe-10 and PDPPSe12 entail longer ordering of interchain π−π stacking than that of PDPPSe (see Figure 3). As listed in Table S1, not only the fwhm’s of (010) signals but also those of (x00) ones become smaller for thin films of all three polymers after thermal annealing at 180 °C. In particular, fwhm’s of scattering signals for the thermally annealed thin films of PDPPSe-10 and PDPPSe-12 are smaller than those of PDPPSe thin film, indicating that the thermally annealed thin films of PDPPSe-10 and PDPPSe-12 show improved thin film crystallinity. Thin film morphologies were characterized with AFM. Figure 4 and Figure S13 and show AFM images of thin films of PDPPSe-10 and PDPPSe-12 as well as PDPPSe before and after thermal annealing at 180 °C. The thin film morphologies for each polymer before and after thermal annealing are similar, but root-mean-square (RMS) roughnesses increase after thermal annealing. PDPPSe-10 and PDPPSe-12 exhibit similar

Figure 4. AFM height images of the thermally (at 180 °C) annealed thin-films of PDPPSe (a), PDPPSe-10 (b), and PDPPSe-12 (c). 3095

DOI: 10.1021/acs.chemmater.8b01007 Chem. Mater. 2018, 30, 3090−3100

Article

Chemistry of Materials

Scheme 3. Schematic Illustration of the Reduction of Steric Crowding for Side Chains after the Replacement of One Branching Bulky Chain with the Linear One at Each DPP Unit

PDPPSe-10, PDPPSe-12, and PDPPSe were also measured under a nitrogen atmosphere. On the basis of transfer and output curves, thin films of PDPPSe-10 and PDPPSe-12 as well as PDPPSe exhibit ambipolar semiconducting properties under a nitrogen atmosphere. This is in a good agreement with previous reports on conjugated diketopyrrolopyrrole-selenophene polymer.29 The ambipolar semiconducting performance data are listed in Table 2. In comparison with those of PDPPSe, the respective hole and electron mobilities are obviously incremented, although they are not balanced. For instance, hole and electron mobilities are as high as 6.5 and 0.48 cm2 V−1 s−1 for thin films of PDPPSe-10 after thermal annealing at 180 °C, which are about 6 and 3 times than those of PDPPSe under the same condition. Similarly, hole and electron mobilities for the thin film of PDPPSe-12 are even boosted to 7.9 and 0.79 cm2 V−1 s−1 after thermal annealing at 180 °C, being about 7 and 4 times than those of PDPPSe under the same condition. Furthermore, it is noted that the enhancement of charge mobilities for PDPPSe-10 and PDPPSe-12 does not compromise other semiconducting parameters (VTh and Ion/off) by comparing with those of PDPPSe (see Table 2). We not only extracted the mobilities by fitting the linear part of the respective plots of the root square of IDS vs VGS at the low voltage region (see above) but also employed a “two-

Figure 5. p-type transfer and output characteristics of FETs with PDPPSe (a, d), PDPPSe-10 (b, e), and PDPPSe-12 (c, f) after thermal annealing at 180 °C, measured under ambient atmosphere. The p-type, n-type transfer and output characteristics of FETs with PDPPSe (g, j, g′, j′), PDPPSe-10 (h, k, h′, k′), and PDPPSe-12 (i, l, i′, l′) after thermal annealing at 180 °C.

Table 2. Mobilities (μh and μe), Threshold Voltages (VTh), and Ion/Ioff Ratios for FETs with Thin Films of PDPPSe-10, PDPPSe12, and PDPPSe under air polymer

temp. (°C)

μha,c (cm2 V−1 s−1)

PDPPSe-10

RT 180 RT 180 RT 180

2.2/1.8 8.1/5.8 3.1/2.2 9.4/7.5 0.48/0.30 1.35/0.89

PDPPSe-12 PDPPSe

under nitrogen

VTha (V)

Ion/Ioffa

μhb,c (cm2 V−1 s−1)

−10 to 2 −8 to3 −8 to 1 −7 to 3 −9 to 1 −10 to 0

105−106 105−106 105−106 105−106 105−106 105−106

1.7/1.2 6.5/4.1 2.4/1.6 7.9/5.3 0.43/0.26 1.18/0.85

VThb (V) −7 −4 −5 −3 −8 −7

to to to to to to

−19 −17 −16 −15 −18 −16

Ion/Ioffa

μeb,c (cm2 V−1 s−1)

VThb (V)

Ion/Ioffb

103−105 103−105 103−104 102−104 103−104 102−104

0.17/0.11 0.48/0.36 0.26/0.18 0.79/0.51 0.050/0.036 0.19/0.14

86−95 78−100 83−94 85−95 82−90 78−86

102−103 102−103 102−103 102−103 102−103 102−103

a Measured under ambient atmosphere, and the channel width (W) and length (L) were 1440 and 10 μm, respectively. bMeasured under N2, and the channel width (W) and length (L) were 1440 and 50 μm. cThe mobilities were provided in “highest/average’’ form, and the performance data were obtained based on more than 10 different FETs.

3096

DOI: 10.1021/acs.chemmater.8b01007 Chem. Mater. 2018, 30, 3090−3100

Article

Chemistry of Materials points” method to estimate charge mobilities as reported early,9 in order to avoid the mobility overestimation. Two points at the respective VTh and VGS of the plot of IDS1/2 versus VGS were selected, and the mobilities were estimated and are shown in Table S2. The details of the mobility extraction with this “twopoints” method is provided in the Supporting Information and Figure S15. The maxima hole mobilities of PDPPSe-12, PDPPSe-10, and PDPPSe in air were estimated to be 4.7, 4.1, and 0.98 cm2 V−1 s−1 after thermal annealing, whereas the maxima hole/electron mobilities for PDPPSe-12, PDPPSe-10, and PDPPSe in nitrogen were estimated to be 2.7/0.58, 2.1/ 0.35, and 0.92/0.16 cm2 V−1 s−1. Above all, both mobility estimation methods support the conclusion that PDPPSe-12 and PDPPSe-10 exhibit better semiconducting performances than PDPPSe. Oh, Yang, and their co-workers reported three siloxane terminated polymers PTDPPSe-SiC4, PTDPPSe-SiC5, and PTDPPSe-SiC629 (see Scheme 4) with the same conjugated

induced by the replacement of one branching alkyl chain with the linear one at each DPP unit, lead to remarkable enhancement of charge mobilities for PDPPSe-10 and PDPPSe-12. Hole mobilities of thin films of PDPPSe-10 and PDPPSe-12 in air are boosted to 8.1 and 9.4 cm2 V−1 s−1. It is expected that this simple and efficient approach for improving semiconducting performances is applicable for other conjugated polymers, and further investigations along this vein are underway.



EXPERIMENTAL SECTION

The reagents and starting materials were commercially available and used without any further purification, if not specified elsewhere. PDPPSe was prepared according to the reported procedure.39,43 Synthesis of Compound 2. Compound 1 (4.0 g, 13.3 mmol) and K2CO3 (3.7 g, 26.7 mmol) were added to a 250 mL double-neck round-bottom flask, and 120 mL of dry DMF was injected by syringe under a nitrogen atmosphere. The reaction was heated to 120 °C for 30 min. 11-(Bromomethyl)tricosane (5.6 g, 13.3 mmol) was slowly added dropwise over 4 h, and the reaction mixture was stirred for 12 h at 120 °C. Solvents were removed by rotary evaporation and the residue was purified by column chromatography. Red-brown crude product (2.4 g, 28%) was obtained without further purification. HRMS (MALDI-TOF): calcd. for C38H56N2O2S2 (M+) 637.3855; Found: 637.3856. Synthesis of Compound 3. Compound 2 (2.4 g, 3.8 mmol) and K2CO3 (1.04 g, 7.5 mmol) were added to a 250 mL double-neck round-bottom flask, and 100 mL of dry DMF was injected by syringe under a nitrogen atmosphere. The reaction was heated to 120 °C for 30 min. 1-Bromodecane (1.25 g, 5.7 mmol) was added, and the reaction mixture was stirred for 8 h at 120 °C. Solvents were removed by rotary evaporation and the residue was purified by column chromatography with CH2Cl2 and petroleum ether (1:3, v/v) as the eluent. Compound 3 was obtained as a red-brown solid (2.3 g, 80%). 1 H NMR (400 MHz, CDCl3): δ 8.95−8.94 (m, 1H), 8.85−8.84 (m, 1H), 7.63−7.61 (m, 2H), 7.29−7.25 (m, 2H), 4.08−4.00 (m, 4H), 1.95−1.83 (m, 1H), 1.79−1.71 (m, 2H), 1.42−1.21 (m, 54H), 0.89− 0.86 (m, 9H). 13C NMR (CDCl3, 100 MHz): δ161.74, 161.37,140.35, 140.07, 135.34, 135.03, 130.53, 130.42, 129.88, 129.82, 128.60, 128.33, 108.13, 107.62, 46.23, 42.25, 37.75, 31.91, 31.22, 29.99, 29.62, 29.53, 29.35, 29.27, 26.88, 26.23, 22.67, 14.08. HR-MS (MALDI-TOF): calcd. for C48H76N2O2S2 (M+) 777.5421; Found: 777.5421. Anal. Calcd for C48H76N2O2S2: C,74.17; H,9.86; N, 3.60; S,8.25; Found: C,74.24; H,9.78; N, 3.62; S,8.18. Synthesis of Compound 4. Compound 2 (2.0 g, 3.1 mmol) and K2CO3 (0.87 g, 6.3 mmol) were added to a 250 mL double-neck round-bottom flask; then dry DMF (100 mL) was injected by syringe under a nitrogen atmosphere. The reaction was heated to 120 °C for 30 min. 1-Bromododecane (1.17 g, 4.7 mmol) was added, and the reaction mixture was stirred for 8 h at 120 °C. Solvents were removed by rotary evaporation and the residue was purified by column chromatography with CH2Cl2 and petroleum ether (1:3, v/v) as the eluent. Compound 4 was obtained as a red-brown solid (2.0 g, 79%). 1 H NMR (400 MHz, CDCl3): δ 8.95−8.94 (m, 1H), 8.85−8.84 (m, 1H), 7.63−7.61 (m, 2H), 7.29−7.25 (m, 2H), 4.08−4.00 (m, 4H), 1.91−1.88 (m, 1H), 1.79−1.71 (m, 2H), 1.44−1.21 (m, 58H), 0.89− 0.85 (m, 9H). 13C NMR (CDCl3, 100 MHz): δ 161.75, 161.38, 140.35, 140.08, 135.34, 135.03, 130.54, 130.43, 129.88, 129.82, 128.60, 128.34, 108.13, 107.62, 46.23, 42.26, 37.75, 31.91, 31.22, 29.99, 29.62, 29.54, 29.33, 29.24, 26.89, 26.23, 22.67, 14.09. HR-MS (MALDITOF): calcd. for C50H80N2O2S2 (M+) 805.5735; Found: 805.5734. Anal. Calcd for C50H80N2O2S2: C,74.57; H, 10.01; N, 3.48; S, 7.96; Found: C, 74.49; H, 10.00; N, 3.37; S, 7.89. Synthesis of Compound 5. In a 100 mL round-bottom flask, compound 3 (1.1 g, 1.4 mmol) was dissolved in 50 mL of CHCl3 under a nitrogen atmosphere. N-Bromosuccinimide (0.55 g, 3.1 mmol) was added, and the reaction mixture was stirred for 4 h at room temperature. Solvents were removed by rotary evaporation and the

Scheme 4. Chemical Structures of PTDPPSe-SiC4/5/6 and Thin Film Charge Mobilities Measured under Nitrogen by Spin-Coating29

backbones as for PDPPSe-10 and PDPPSe-12. They were reported to show high hole and electron mobilities by using a solution-sheared technique.61 However, thin films of these siloxane terminated polymers, which were prepared with the conventional spin-coating technique, exhibited lower hole mobilities than those of PDPPSe-10 and PDPPSe-12, whereas their electron mobilities are comparable to those of PDPPSe-10 and PDPPSe-12 (see Scheme 4). Therefore, this new side-chain engineering approach for DPP-based conjugated D-A polymers, in which one bulky branching alkyl chain is replaced with the linear one at each DPP unit, is advantageous in terms of boosting charge mobilities.



CONCLUSIONS In this paper, we introduce a new side-chain engineering approach for DPP-based conjugated D-A polymers in which one bulky branching alkyl chain is replaced with the linear one at each DPP unit. With this new approach in mind, two new DPP-based conjugated polymers PDPPSe-10 and PDPPSe-12 were prepared and investigated. It is noted that the corresponding DPP monomers can be simply synthesized from commercially available compounds. The results reveal that (i) the replacement of bulky branching chains with linear ones can weaken the steric hindrance owing to branching chains, and accordingly, conjugated backbones of PDPPSe-10 and PDPPSe-12 become more planar and rigid; (ii) the incorporation of linear alkyl chains as in PDPPSe-10 and PDPPSe-12 is beneficial for side-chain interdigitation; and (iii) interchain packing order and thin film crystallinity are obviously improved for PDPPSe-10 and PDPPSe-12 by comparing with PDPPSe. These solid state structural changes, which are 3097

DOI: 10.1021/acs.chemmater.8b01007 Chem. Mater. 2018, 30, 3090−3100

Chemistry of Materials



residue was purified with column chromatography using CH2Cl2 and petroleum ether (1:3, v/v) as the eluent. Compound 5 was obtained as a red-purple solid (900 mg, 68%). 1H NMR (400 MHz, CDCl3): δ 8.70 (d, J = 4.0 Hz, 1H), 8.60 (d, J = 4.0 Hz, 1H), 7.22 (dd, J = 8.0, 4.0 Hz, 2H), 3.97 (t, J = 8.0 Hz, 2H), 3.91 (d, J = 8.0 Hz, 2H), 1.88−1.85 (m, 1H), 1.75−1.68 (m, 2H), 1.41−1.21 (m, 54H), 0.89−0.86 (m, 9H). 13C NMR (CDCl3, 100 MHz): δ 161.38, 161.01,139.29, 139.06, 135.43, 135.11, 131.64, 131.39, 131.23, 131.15, 119.06, 118.92, 108.20, 107.73, 46.37, 42.31, 37.78, 31.91, 31.87, 31.22, 29.96, 29.67, 29.62, 29.53, 29.48, 29.34, 29.26, 29.17, 26.82, 26.21, 22.67, 14.08. HR-MS (MALDI-TOF): calcd. for C48H74Br2N2O2S2 (M+) 932.3553; Found: 932.3553. Anal. Calcd for C48H74Br2N2O2S2: C, 61.66; H, 7.98; N, 3.00; S, 6.86; Found: C, 61.58; H, 8.11; N, 2.86; S, 6.83. Synthesis of Compound 6. In a 100 mL round-bottom flask, compound 3 (1.2 g, 1.5 mmol) was dissolved in 50 mL of CHCl3 under a nitrogen atmosphere. N-Bromosuccinimide (0.58 g, 3.3 mmol) was added, and the reaction mixture was stirred for 4 h at room temperature. Solvents were removed by rotary evaporation and the residue was purified by column chromatographywith CH2Cl2 and petroleum ether (1:3, v/v) as the eluent. Compound 5 was obtained as a red-purple solid (0.95 g, 66%). 1H NMR (400 MHz, CDCl3): δ 8.68 (d, J = 4.0 Hz, 1H), 8.59 (d, J = 4.0 Hz, 1H), 7.22 (dd, J = 8.0, 4.0 Hz, 2H), 3.98 (t, J = 8.0, 2H), 3.91 (d, J = 8.0 Hz, 2H), 1.88−1.85 (m, 1H), 1.76−1.68 (m, 2H), 1.41−1.22 (m, 58H), 0.90−0.86 (m, 9H). 13 C NMR (CDCl3, 100 MHz): δ 161.39, 161.02, 139.29, 139.06, 135.42, 135.10, 131.64, 131.39, 131.23, 131.15, 119.05, 118.91, 108.22, 107.74, 46.38, 42.31, 37.79, 31.92, 31.90, 31.23, 29.95, 29.67, 29.64, 29.61, 29.53, 29.48, 29.34, 29.32, 29.18, 26.83, 26.21, 22.67, 14.07. HRMS (MALDI-TOF): calcd. for C50H78Br2N2O2S2 (M+) 960.3867; Found: 960.3866. Anal. Calcd for C50H78Br2N2O2S2: C, 62.36; H, 8.16; N, 2.91; S, 6.66; Found: C, 62.07; H, 8.23; N, 2.81; S, 6.70. General Synthetic Procedures for PDPPSe-10, and PDPPSe12. Compounds 5 or 6 (1.0 equiv), (dimethyl(5-(trimethylstannyl)selenophen-2-yl)stannyl)methylium (1.0 equiv), Pd2(dba)3 (0.01 equiv), and P(o-tol)3 (0.08 equiv) were dissolved in toluene (10 mL) in a Schlenk tube. Through a freeze−pump−thaw cycle, the tube was charged with nitrogen for three times. The reaction mixture was stirred at 110 °C for 3 h. The resulting mixture was poured into methanol and stirred for 3.0 h. The mixture was poured into CH3OH and filtered. Each polymer was purified by Soxhlet extraction with various solvents (CH3 OH, acetone, hexane, and chloroform sequentially) to remove the existing oligomers and other impurities. The resulting polymer was collected and dried under vacuum at 50 °C for 48 h. Synthesis of PDPPSe-10. Compound 5 (150 mg, 0.16 mmol), compound 7 (73.3 mg, 0.16 mmol), P(o-tol)3 (3.8 mg, 0.0125 mmol), and Pd2(dba)3 (1.5 mg, 0.0016 mmol) were used. The purified PDPPSe-10 was collected to give a deep blue solid (135 mg, 93% yield). 1H NMR (500 MHz, 1,1,2,2-tetrachloroethane-d2, 100 °C): δ 8.90 (m, 2H), 7.39−7.03 (m, 4H), 4.03 (m, 4H), 2.05−2.02 (m, 3H), 1.41−1.27 (m, 54H), 0.92−0.87 (m, 9H). 13C NMR (100 MHz, solid): δ 160.45, 142.32, 137.41, 128.47, 126.28, 108.81, 45.56, 42.33, 39.39, 32.82, 30.52, 23.50, 14.82. Anal. Calcd for (C52H76N2O2S2Se)n: C, 69.05; H, 8.49; N, 3.10; S, 7.09; Found: C, 68.75; H, 8.37; N, 2.76; S, 6.88. Synthesis of PDPPSe-12. Compound 6 (150 mg, 0.156 mmol), compound 7 (71.1 mg, 0.156 mmol), P(o-tol)3 (3.8 mg, 0.0125 mmol), and Pd2(dba)3 (1.4 mg, 0.0016 mmol) were used. The purified PDPPSe-12 was collected to give a deep blue solid (134 mg, 92% yield). 1H NMR (500 MHz, 1,1,2,2-tetrachloroethane-d2, 100 °C): δ 8.90 (m, 2H), 7.40−7.03 (m, 4H), 4.02 (m, 4H), 2.05−1.99 (m, 3H), 1.41−1.27 (m, 58H), 1.01−0.89 (m, 9H). 13C NMR (100 MHz, solid): δ 160.46, 142.29, 137.31, 127.26, 126.61, 108.77, 45.37, 40.75, 39.09, 32.80, 30.54, 23.49, 14.81. Anal. Calcd for (C54H80N2O2S2Se)n: C, 69.55; H, 8.67; N, 3.00; S, 6.88; Found: C, 69.03; H, 8.43; N, 2.91; S, 7.11.

Article

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.chemmater.8b01007. TGA, DSC, Raman spectra, cyclic voltammograms, PDS, fabrication of OFETs, 1H NMR and 13C NMR data (PDF)



AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected] (D.Z.). *E-mail: [email protected] (Z.L.). ORCID

Zhijie Wang: 0000-0001-5092-0735 Zitong Liu: 0000-0003-1185-9219 Yuanping Yi: 0000-0002-0052-9364 Zhengxu Cai: 0000-0003-0239-9601 Aditya Sadhanala: 0000-0003-2832-4894 Guanxin Zhang: 0000-0002-1417-6985 Wei Chen: 0000-0001-8906-4278 Deqing Zhang: 0000-0002-5709-6088 Author Contributions

D.Z. proposed the study and designed the polymer structures. D.Z. and Z.L. analyzed the data. Z.W. and Z.L. carried out the experiments. Z.W. and G.Z. did structural characterization. L.N. and Y.Y. carried out the theoretical calculations. M.X. and A.S. carried out the PDS experiments, and H.S. explained the data. Z.C. and W.C. carried out the GIWAXS experiments. D.Z. and Z.L. prepared the manuscript. All authors discussed, revised, and approved the manuscript. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We thank the Strategic Priority Research Program of the CAS (XDB12010300), the National Key R&D Program of China (2017YFA0204701), and NSFC (21661132006) for the financial support. M.X. thanks the Cambridge Overseas Trust and Chinese Scholarship Council for Ph.D. funding. W.C. gratefully acknowledges financial support from the U.S. Department of Energy, Office of Science, Materials Sciences and Engineering Division. We also thank Dr. Joseph Strzalka and Dr. Zhang Jiang for the assistance with GIWAXS measurements. Use of the Advanced Photon Source (APS) at Argonne National Laboratory was supported by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences, under contract no. DE-AC02-06CH11357.



REFERENCES

(1) Sirringhaus, H. 25th Anniversary Article: Organic Field-Effect Transistors: The Path Beyond Amorphous Silicon. Adv. Mater. 2014, 26, 1319−1335. (2) Sokolov, A. N.; Tee, B. C.-K.; Bettinger, J. B.-H.; Tok, J. B.-H.; Bao, Z. Chemical and Engineering Approaches to Enable Organic Field-Effect Transistors for Electronic Skin Applications. Acc. Chem. Res. 2012, 45, 361−371. (3) Tee, B. C.-K.; Chortos, A.; Berndt, A.; Nguyen, A. K.; Tom, A.; McGuire, A.; Lin, Z. C.; Tien, K.; Bae, W.-G.; Wang, H.; Mei, P.; Chou, H.-H.; Cui, B.; Deisseroth, K.; Ng, T. N.; Bao, Z. A skin-

3098

DOI: 10.1021/acs.chemmater.8b01007 Chem. Mater. 2018, 30, 3090−3100

Article

Chemistry of Materials inspired Organic Digital Mechanoreceptor. Science 2015, 350, 313− 316. (4) Zang, Y.; Zhang, F.; Huang, D.; Gao, X.; Di, C.-A.; Zhu, D. Flexible Suspended Gate Organic Thin-Film Transistors for UltraSensitive Pressure Detection. Nat. Commun. 2015, 6, 6269. (5) Hwang, Y. J.; Courtright, B. A.; Ferreira, A. S.; Tolbert, S. H.; Jenekhe, S. A. 7.7% Efficient All-Polymer Solar Cells. Adv. Mater. 2015, 27, 4578−4584. (6) Kozycz, L. M.; Gao, D.; Tilley, A. J.; Seferos, D. S. One donortwo acceptor (D-A1)-(D-A2) random terpolymers containing perylene diimide, naphthalene diimide, and carbazole units. J. Polym. Sci., Part A: Polym. Chem. 2014, 52, 3337−3345. (7) Grzybowski, M.; Gryko, D. T. Diketopyrrolopyrroles: Synthesis, Reactivity, and Optical Properties. Adv. Opt. Mater. 2015, 3, 280−320. (8) Nielsen, C. B.; Turbiez, M.; McCulloch, I. Recent Advances in the Development of Semiconducting DPP-Containing Polymers for Transistor Applications. Adv. Mater. 2013, 25, 1859−1880. (9) Yi, Z.; Wang, S.; Liu, Y. Design of High-Mobility Diketopyrrolopyrrole-Basedπ-Conjugated Copolymers for Organic Thin-Film Transistors. Adv. Mater. 2015, 27, 3589−3606. (10) Yao, J.; Yu, C.; Liu, Z.; Luo, H.; Yang, Y.; Zhang, G.; Zhang, D. Significant Improvement of Semiconducting Performance of the Diketopyrrolopyrrole−Quaterthiophene Conjugated Polymer through Side-Chain Engineering via Hydrogen-Bonding. J. Am. Chem. Soc. 2016, 138, 173−185. (11) Luo, H.; Yu, C.; Liu, Z.; Zhang, G.; Geng, H.; Yi, Y.; Broch, K.; Hu, Y.; Sadhanala, A.; Jiang, L.; Qi, P.; Cai, Z.; Sirringhaus, H.; Zhang, D. Remarkable Enhancement of Charge Carrier Mobility of Conjugated Polymer Field-Effect Transistors Upon Incorporating an Ionic Additive. Sci. Adv. 2016, 2, e1600076. (12) An, T. K.; Kang, I.; Yun, H. J.; Cha, H.; Hwang, J.; Park, S.; Kim, J.; Kim, Y. J.; Chung, D. S.; Kwon, S-K.; Kim, Y.-H.; Park, C. E. Solvent Additive to Achieve Highly Ordered Nanostructural Semicrystalline DPP Copolymers: Toward a High Charge Carrier Mobility. Adv. Mater. 2013, 25, 7003−7009. (13) Xu, Y.; Sun, H.; Shin, E. Y.; Lin, Y. F.; Li, W.; Noh, Y. Y. PlanarProcessed Polymer Transistors. Adv. Mater. 2016, 28, 8531−8537. (14) Li, W.; Hendriks, K. H.; Furlan, A.; Roelofs, W. S.; Wienk, M. M.; Janssen, R. A. J. Universal Correlation between Fibril Width and Quantum Efficiency in Diketopyrrolopyrrole-Based Polymer Solar Cells. J. Am. Chem. Soc. 2013, 135, 18942−18948. (15) Bronstein, H.; Chen, Z.; Ashraf, R. S.; Zhang, W.; Du, J.; Durrant, J. R.; Tuladhar, P. S.; Song, K.; Watkins, S. E.; Geerts, Y.; Wienk, M. M.; Janssen, R. A.; Anthopoulos, T.; Sirringhaus, H.; Heeney, M.; McCulloch, I. Thieno[3,2-b]thiophene-Diketopyrrolopyrrole-Containing Polymers for High-Performance Organic FieldEffect Transistors and Organic Photovoltaic Devices. J. Am. Chem. Soc. 2011, 133, 3272−3275. (16) Guo, X.; Puniredd, S. R.; Baumgarten, M.; Pisula, W.; Müllen, K. Rational Design of Benzotrithiophene-Diketopyrrolopyrrole-Containing Donor-Acceptor Polymers for Improved Charge Carrier Transport. Adv. Mater. 2013, 25, 5467−5472. (17) Han, A. R.; Dutta, G. K.; Lee, J.; Lee, H. R.; Lee, S. M.; Ahn, H.; Shin, T. J.; Oh, J. H.; Yang, C. ε-Branched Flexible Side Chain Substituted Diketopyrrolopyrrole-Containing Polymers Designed for High Hole and Electron Mobilities. Adv. Funct. Mater. 2015, 25, 247− 254. (18) Gao, X.; Zhao, Z. High Mobility Organic Semiconductors for Field-Effect Transistors. Sci. China: Chem. 2015, 58, 947−968. (19) Liu, J.; Sun, Y.; Moonsin, P.; Kuik, M.; Proctor, C. M.; Lin, J.; Hsu, B. B.; Promarak, V.; Heeger, A. J.; Nguyen, T.-Q. TriDiketopyrrolopyrrole Molecular Donor Materials for High-Performance Solution-Processed Bulk Heterojunction Solar Cells. Adv. Mater. 2013, 25, 5898−5903. (20) 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.

(21) 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. (22) Zhang, A.; Xiao, C.; Meng, D.; Wang, Q.; Zhang, X.; Hu, W.; Zhan, X.; Wang, Z.; Janssen, R. A. J.; Li, W. Conjugated Polymers with Deep LUMO Levels for Field-Effect Transistors and Polymer− Polymer Solar Cells. J. Mater. Chem. C 2015, 3, 8255−8261. (23) Nelson, T. L.; Young, T. M.; Liu, J.; Mishra, S. P.; Belot, J. A.; Balliet, C. L.; Javier, A. E.; Kowalewski, T.; McCullough, R. D. Transistor Paint: High Mobilities in Small Bandgap Polymer Semiconductor Based on the Strong Acceptor, Diketopyrrolopyrrole and Strong Donor, Dithienopyrrole. Adv. Mater. 2010, 22, 4617−4621. (24) 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. (25) Ha, J. S.; Kim, K. H.; Choi, D. H. 2,5-Bis(2-octyldodecyl)pyrrolo[3,4-c]pyrrole-1,4-(2H,5H)-dione-Based Donor-Acceptor Alternating Copolymer Bearing 5,5′-Di(thiophen-2-yl)-2,20-biselenophene Exhibiting 1.5 cm2·V−1·s−1 Hole Mobility in Thin-Film Transistors. J. Am. Chem. Soc. 2011, 133, 10364−10367. (26) Yi, Z.; Sun, X.; Zhao, Y.; Guo, Y.; Chen, X.; Qin, J.; Yu, G.; Liu, Y. Diketopyrrolopyrrole-Based π-Conjugated Copolymer Containing β-Unsubstituted Quintetthiophene Unit: A Promising Material Exhibiting High Hole-Mobility for Organic Thin-Film Transistors. Chem. Mater. 2012, 24, 4350−4356. (27) Yi, Z.; Ma, L.; Chen, B.; Chen, D.; Chen, X.; Qin, J.; Zhan, X.; Liu, Y.; Ong, W. J.; Li, J. Effect of the Longer β-Unsubstituted Oligothiophene Unit (6T and 7T) on the Organic Thin-Film Transistor Performances of Diketopyrrolopyrrole-Oligothiophene Copolymers. Chem. Mater. 2013, 25, 4290−4296. (28) Kang, I.; Yun, H. J.; Chung, D. S.; Kwon, S. K.; Kim, Y. H. Record High Hole Mobility in Polymer Semiconductors via SideChain Engineering. J. Am. Chem. Soc. 2013, 135, 14896−14899. (29) Lee, J.; Han, A. R.; Yu, H.; Shin, T. J.; Yang, C.; Oh, J. H. Boosting the Ambipolar Performance of Solution-Processable Polymer Semiconductors via Hybrid Side-Chain Engineering. J. Am. Chem. Soc. 2013, 135, 9540−9547. (30) Lee, J.; Han, A. R.; Kim, J.; Kim, Y.; Oh, J. H.; Yang, C. Solution-Processable Ambipolar Diketopyrrolopyrrole−Selenophene Polymer with Unprecedentedly High Hole and Electron Mobilities. J. Am. Chem. Soc. 2012, 134, 20713−20721. (31) Mei, J.; Kim, D. H.; Ayzner, A. L.; Toney, M. F.; Bao, Z. Siloxane-Terminated Solubilizing Side Chains: Bringing Conjugated Polymer Backbones Closer and Boosting Hole Mobilities in Thin-Film Transistors. J. Am. Chem. Soc. 2011, 133, 20130−20133. (32) Lei, T.; Dou, J.-H.; Pei, J. Influence of Alkyl Chain Branching Positions on the Hole Mobilities of Polymer Thin-Film Transistors. Adv. Mater. 2012, 24, 6457−6461. (33) Kawabata, K.; Saito, M.; Takemura, N.; Osaka, I.; Takimiya, K. Effects of Branching Position of Alkyl Side Chains on Ordering Structure and Charge Transport Property in Thienothiophenedioneand Quinacridone-Based Semiconducting Polymers. Polym. J. 2017, 49, 169−176. (34) Meager, I.; Ashraf, R. S.; Mollinger, S.; Schroeder, B. C.; Bronstein, H.; Beatrup, D.; Vezie, M. S.; Kirchartz, T.; Salleo, A.; Nelson, J.; McCulloch, I. Photocurrent Enhancement from Diketopyrrolopyrrole Polymer Solar Cells through Alkyl-Chain Branching Point Manipulation. J. Am. Chem. Soc. 2013, 135, 11537−11540. (35) Fu, B.; Baltazar, J.; Sankar, A. R.; Chu, P.-H.; Zhang, S.; Collard, D. M.; Reichmanis, E. Enhancing Field-Effect Mobility of Conjugated Polymers Through Rational Design of Branched Side Chains. Adv. Funct. Mater. 2014, 24, 3734−3744. (36) Chen, S.; Sun, B.; Hong, W.; Aziz, H.; Meng, Y.; Li, Y. Influence of side chain length and bifurcation point on the crystalline structure and charge transport of diketopyrrolopyrrole-quaterthiophene copolymers (PDQTs). J. Mater. Chem. C 2014, 2, 2183−2190. 3099

DOI: 10.1021/acs.chemmater.8b01007 Chem. Mater. 2018, 30, 3090−3100

Article

Chemistry of Materials (37) Ma, J.; Liu, Z.; Wang, Z.; Yang, Y.; Zhang, G.; Zhang, X.; Zhang, D. Charge Mobility Enhancement for Diketopyrrolopyrrole-Based Conjugated Polymers by Partial Replacement of Branching Alkyl Chains with Linear Ones. Mater. Chem. Front. 2017, 1, 2547−2553. (38) Naik, M. A.; Venkatramaiah, N.; Kanimozhi, C.; Patil, S. Influence of Side-Chain on Structural Order and Photophysical Properties in Thiophene Based Diketopyrrolopyrroles: A Systematic Study. J. Phys. Chem. C 2012, 116, 26128−26137. (39) Yang, S.-F.; Liu, Z.-T.; Cai, Z.-X.; Dyson, M. J.; Stingelin, N.; Chen, W.; Ju, H.-J.; Zhang, G.-X.; Zhang, D.-Q. DiketopyrrolopyrroleBased Conjugated Polymer Entailing Triethylene Glycols as Side Chains with High Thin-Film Charge Mobility without Post-Treatments. Adv. Sci. 2017, 4, 1700048. (40) Zhao, X.; Xue, G.; Qu, G.; Singhania, V.; Zhao, Y.; Butrouna, K.; Gumyusenge, A.; Diao, Y.; Graham, K. R.; Li, H.; Mei, J. Complementary Semiconducting Polymer Blends: Influence of Side Chains of Matrix Polymers. Macromolecules 2017, 50, 6202−6209. (41) Li, C.; Yu, C.; Lai, W.; Liang, S.; Jiang, X.; Feng, G.; Zhang, J.; Xu, Y.; Li, W. Multifunctional Diketopyrrolopyrrole-Based Conjugated Polymers with Perylene Bisimide Side Chains. Macromol. Rapid Commun. 2017, 1700611. (42) Park, S. J.; Jung, J. E.; Kang, M. K.; Na, Y. H.; Song, H. H.; Kang, J. W.; Baek, N. S.; Kim, T.-D. Diketopyrrolopyrrole-Based Conjugated Polymers Containing Alkyl and Aryl Side-Chains for Bulk Heterojunction Solar Cells. Synth. Met. 2015, 203, 221−227. (43) Kim, J.; Kwon, J. E.; Boampong, A. A.; Lee, J.-H.; Kim, M.-H.; Park, S. Y.; Kim, T.-D.; Kim, J. H. Threshold Voltage Modulation of Polymer Transistors By Photoinduced Charge−Transfer Between Donor−Acceptor Dyads. Dyes Pigm. 2017, 142, 387−393. (44) Hess, B.; Kutzner, C.; van der Spoel, D.; Lindahl, E. GROMACS 4: Algorithms for Highly Efficient, Load-Balanced, and Scalable Molecular Simulation. J. Chem. Theory Comput. 2008, 4, 435−447. (45) Bayly, C. I.; Cieplak, P.; Cornell, W. D.; Kollman, P. A. A WellBehaved Electrostatic Potential Based Method Using Charge Restraints for Deriving Atomic Charges: The RESP Model. J. Phys. Chem. 1993, 97, 10269−10280. (46) Han, G.; Guo, Y.; Song, X.; Wang, Y.; Yi, Y. Terminal p−p Stacking Determines Three-Dimensional Molecular Packing and Isotropic Charge Transport in an A−p−A Electron Acceptor for Non-Fullerene Organic Solar Cells. J. Mater. Chem. C 2017, 5, 4852− 4857. (47) Liu, Y.; Zhao, J.; Li, Z.; Mu, C.; Ma, W.; Hu, H.; Jiang, K.; Lin, H.; Ade, H.; Yan, H. Aggregation and Morphology Control Enables Multiple Cases of High-Efficiency Polymer Solar Cells. Nat. Commun. 2014, 5, 5293. (48) Li, M.; An, C.; Marszalek, T.; Baumgarten, M.; Yan, H.; Müllen, K.; Pisula, W. Controlling the Surface Organization of Conjugated Donor-Acceptor Polymers by their Aggregation in Solution. Adv. Mater. 2016, 28, 9430−9438. (49) Nketia-Yawson, B.; Lee, H. S.; Seo, D.; Yoon, Y.; Park, W. T.; Kwak, K.; Son, H. J.; Kim, B.; Noh, Y. Y. A Highly Planar Fluorinated Benzothiadiazole-Based Conjugated Polymer for High-Performance Organic Thin-Film Transistors. Adv. Mater. 2015, 27, 3045−3052. (50) Wood, S.; Wade, J.; Shahid, M.; Collado-Fregoso, E.; Bradley, D. D. C.; Durrant, J. R.; Heeney, M.; Kim, J.-S. Natures of Optical Absorption Transitions and Excitation Energy Dependent Photostability of Diketopyrrolopyrrole (DPP)-Based Photovoltaic Copolymers. Energy Environ. Sci. 2015, 8, 3222−3232. (51) Yu, S. H.; Park, K. H.; Kim, Y.-H.; Chung, D. S.; Kwon, S.-K. Fine Molecular Tuning of Diketopyrrolopyrrole-Based Polymer Semiconductors for Efficient Charge Transport: Effects of Intramolecular Conjugation Structure. Macromolecules 2017, 50, 4227− 4234. (52) Venkateshvaran, D.; Nikolka, M.; Sadhanala, A.; Lemaur, V.; Zelazny, M.; Kepa, M.; Hurhangee, M.; Kronemeijer, A. J.; Pecunia, V.; Nasrallah, I.; Romanov, I.; Broch, K.; McCulloch, I.; Emin, D.; Olivier, Y.; Cornil, J.; Beljonne, D.; Sirringhaus, H. Approaching Disorder-Free Transport in High-Mobility Conjugated Polymers. Nature 2014, 515, 384−388.

(53) Kronemeijer, A. J.; Pecunia, V.; Venkateshvaran, D.; Nikolka, M.; Sadhanala, A.; Moriarty, J.; Szumilo, M.; Sirringhaus, H. TwoDimensional Carrier Distribution in Top-Gate Polymer Field-Effect Transistors: Correlation between Width of Density of Localized States and Urbach Energy. Adv. Mater. 2014, 26, 728−733. (54) Gann, E.; Young, A. T.; Collins, B. A.; Yan, H.; Nasiatka, J.; Padmore, H. A.; Ade, H.; Hexemer, A.; Wang, C. Soft X-Ray Scattering Facility at The Advanced Light Source with Real-Time Data Processing and Analysis. Rev. Sci. Instrum. 2012, 83, 045110. (55) 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. (56) Rivnay, J.; Steyrleuthner, R.; Jimison, L. H.; Casadei, A.; Chen, Z.; Toney, M. F.; Facchetti, A.; Neher, D.; Salleo, A. Drastic Control of Texture in a High Performance n-Type Polymeric Semiconductor and Implications for Charge Transport. Macromolecules 2011, 44, 5246− 5255. (57) Son, S. Y.; Kim, Y.; Lee, J.; Lee, G. Y.; Park, W. T.; Noh, Y. Y.; Park, C. E.; Park, T. High-Field-Effect Mobility of Low-Crystallinity Conjugated Polymers with Localized Aggregates. J. Am. Chem. Soc. 2016, 138, 8096−8103. (58) Szarko, J. M.; Rolczynski, B. S.; Lou, S. J.; Xu, T.; Strzalka, J.; Marks, T. J.; Yu, L.; Chen, L. X. Photovoltaic Function and Exciton/ Charge Transfer Dynamics in a Highly Efficient Semiconducting Copolymer. Adv. Funct. Mater. 2014, 24, 10−26. (59) McCulloch, I.; Salleo, A.; Chabinyc, M. Avoid the Kinks when Measuring Mobility. Science 2016, 352, 1521−1522. (60) Choi, H. H.; Cho, K.; Frisbie, C. D.; Sirringhaus, H.; Podzorov, V. Critical Assessment of Charge Mobility Extraction in FETs. Nat. Mater. 2018, 17, 2−7. (61) Giri, G.; Verploegen, E.; Mannsfeld, S. C.; Atahan-Evrenk, S.; Kim, D. H.; Lee, S. Y.; Becerril, H. A.; Aspuru-Guzik, A.; Toney, M. F.; Bao, Z. Tuning Charge Transport in Solution-Sheared Organic Semiconductors Using Lattice Strain. Nature 2011, 480, 504−508.

3100

DOI: 10.1021/acs.chemmater.8b01007 Chem. Mater. 2018, 30, 3090−3100