Gage T. Masona,
Chloé Lisowskia,
Piumi Kulatungaa,
Tiago C. Gomesa,
Angela Awadaa,
Yu-Xin Hsub,
Yu-Cheng Chiu
b and
Simon Rondeau-Gagné
*a
aDepartment of Chemistry and Biochemistry, University of Windsor, 401 Sunset Avenue, Windsor, ON N9B 3P4, Canada. E-mail: srondeau@uwindsor.ca
bDepartment of Chemical Engineering, National Taiwan University of Science and Technology, No. 43 Keelung Road, Sec 4, Taipei 10607, Taiwan
First published on 14th July 2025
The rise of organic electronics is unlocking exciting possibilities for the development of flexible, sustainable, and energy-efficient technologies. From wearable devices and intelligent food packaging to light-harvesting materials and smart textiles, these innovations are driving the next establishment of the Internet of Things (IoT) and other AI-driven autonomous systems, where interconnected devices and intelligent technologies seamlessly work together to transform industries and daily life. At the core of these emerging technologies are semiconducting polymers, which offer a unique combination of synthetic tunability, solution processability, and tuneable optoelectronic and mechanical properties. These features pave the way for cutting-edge fabrication techniques—such as 3D and inkjet printing—to create high-performance devices. Achieving solution processability in these polymer systems often requires the incorporation of bulky, acyclic aliphatic side chains, amongst many others, to enhance solubility and mitigate strong molecular aggregation. These side chains can significantly impact multiple properties, including charge carrier mobility through their influence on the polymer nanostructure in thin films. This work explores the utilization of dendronized side chains to modulate the properties of semicrystalline polymers. More precisely, we developed a new series of materials by integrating high-performance diketopyrrolopyrrole (DPP)-based backbones with two distinct poly(benzyl ether) side chains and two different donor units. These structural variations were found to significantly influence processing conditions, as well as the electronic and thermomechanical properties of the resulting polymers. Comprehensive characterization, including grazing incidence wide-angle X-ray scattering (GIWAXS), atomic force microscopy (AFM), quantitative nanomechanical mapping (for Young's moduli), and dynamic mechanical analysis (DMA) to determine glass transitions, was performed. The polymers were subsequently employed in the fabrication of organic field-effect transistors (OFETs) to study their impact on the electronic properties. The incorporation of novel dendron-like side chains provides a promising strategy for advancing side-chain engineering in semiconducting polymers, offering new avenues in the development of emerging printed organic electronics.
Within this growing materials toolbox, semiconducting polymers stand out due to their unique properties. These materials offer tunability, solution processability, and the ability to customize their properties through rational design.21 This enables precise control over thermomechanical properties, such as low Young's moduli and glass transition temperatures (Tg), which are crucial for achieving flexibility and durability.1,20,22 These features position semiconducting polymers as ideal candidates for the large-scale manufacturing of printed electronics, pushing the boundaries of what these emerging technologies can achieve.10,14,23,24
Among the various synthetic strategies for fine-tuning and optimizing semiconducting polymers, side-chain engineering stands out as particularly effective.21 Often employed to solubilize otherwise insoluble π-conjugated polymers, side chains have evolved into critical design elements that significantly influence key properties such as solid-state polymer chain packing, polarity, aggregation, and thermal transitions.21,25–27 Numerous studies, predominantly using long branched aliphatic chains, demonstrate how side chains can be precisely tailored to control and enhance the performance of materials in organic electronics. For example, Bao et al. reported a series of branched aliphatic side chains with different lengths in diketopyrrolopyrrole-based (DPP) polymers and found that the morphology (face-on vs. edge-on orientation) changes drastically with the side chain length. They also found that these morphological changes lead to drastic differences in charge carrier mobility, ranging between 1.4 and 3.7 cm2 V−1 s−1.28 Their findings suggest that a mixed morphology is highly beneficial for creating charge percolation pathways and an intertwined 3-D network. Beyond traditional alkyl chains, side chains in semiconducting polymers provide versatile anchor points to push functionality further. Precise design of new side chains has enabled the incorporation of a diverse range of complex moieties and functional groups, including ethylene glycol, carbohydrates, amino acids, crosslinkable groups, oligosiloxanes and more.29–34 These advancements in side chain engineering were shown to not only lead to drastically different physical and chemical properties but also broaden the application potential of conjugated polymers in next-generation electronic devices.
Notably, as demonstrated in numerous examples of side chain engineering, the bulk dimension and molecular flexibility of these solubilizing side chains are key design elements that show a profound influence on the electronic and mechanical properties of semiconducting polymers. For instance, Mei et al. reported the use of bulky oligosiloxane side chains in polyisodingo polymers.35 In this work, the authors showed that pushing the siloxane chains further away from the backbone led to an important decrease in π-stacking distances, from 3.63 to 3.37 Å. This difference also led to an important increase in device performance, reaching a charge carrying mobility as high as 4.8 cm2 V−1 s−1 in OFETs. This phenomenon can be attributed to the bulkiness of these side chains: as the side chain gets bulkier and pushed further away from the main polymer chain, the π-conjugated backbone can get closer through enhanced π-stacking, thus promoting charge transport through charge hopping and the overall charge carrier mobility measured in thin film transistors, all while remaining processable. Building from this finding, several reports focused on the rational design of side chains to increase their bulkiness and investigated not only the impact on the electronic properties but also the impact on the mechanical properties. For instance, Chen et al. prepared a series of isoindigo–bithiophene polymers with systematically varying carbosilane side chain lengths (from C6 to C10) and found that as longer side chains were incorporated into the polymers the film moduli decreased, and that the longest chain maintained the highest charge carrier mobility under strain.36
Building on previous work, the incorporation of dendron-inspired side chains into semiconducting polymers has been explored, with dendronized side chains emerging as a promising platform for advanced semiconductors. These side chains are distinguished by their high degree of functionalization, monodispersity, and bulkiness. Among others, the use of poly(benzyl ether) dendrons to modify conjugated polymers is a particularly attractive strategy that has been explored in polyphenylene and polythiophene-based conjugated polymers primarily to improve solubility and processability, though often at the cost of molecular packing and crystallinity.37–39 Leveraging the unique features of dendronized side chains, our group developed a method to integrate short, polyamidoamine (PAMAM) dendritic side chains into semicrystalline diketopyrrolopyrrole-based polymers using an azide–alkyne Huisgen 1,3-dipolar cycloaddition reaction.40 Comprehensive characterization of the resulting dendronized semiconducting polymers revealed that dendritic side chains effectively reduce polymer aggregation and crystallinity in thin films. Notably, PAMAM-containing polymers exhibited good charge transport properties in OFETs, performing in a similar manner to diketopyrrolopyrrole-based polymers featuring branched alkyl side chains. While these findings highlight the potential of dendritic side chains for designing advanced semiconducting polymers, their use remains relatively underexplored in semicrystalline donor–acceptor (D–A) semiconducting polymers, with their impact on optoelectronic, thermomechanical, and solid-state properties still challenging to predict. Exploring dendritic side chains derived from more rigid and bulky moieties, as well as their integration into D–A systems, is therefore an interesting avenue to explore more efficient organic electronics.
Herein, we report the design and synthesis of diketopyrrolopyrrole-based semiconducting polymers incorporating poly(benzyl ether) side chains (Fig. 1). Using a macromonomer approach, the dendritic side chain was expanded from generation 1 (G1) to generation 2 (G2) to investigate the effects of side chain bulkiness and rigidity on material properties. Additionally, the donor unit in the donor–acceptor polymers was changed from bithiophene to thienothiophene to explore the influence of donor design. Comprehensive characterization using optical spectroscopies, atomic force microscopy (AFM), and X-ray scattering revealed that the bulky dendronized side chains effectively reduce polymer aggregation in solution while increasing inter-chain spacing in the solid state. Thermomechanical properties were analyzed using dynamic mechanical analysis (DMA) and quantitative nanomechanical mapping (AFM). The results showed that the side chain Tg is between −75 and −85 °C, while backbone Tg remains near 0 °C. Increasing the side chain generation from G1 to G2 softened the side chains, lowering the Tg, while the backbone Tg increased with side chain bulk. Similarly, the Young's moduli decreased with higher side chain generation, highlighting the influence of dendritic architecture on mechanical properties. Finally, the materials were tested in OFETs, revealing that increasing the size of the side chain incorporated into the polymer led to reduced charge carrier mobility but resulted in a softer material. Overall, these findings demonstrate the potential of incorporating bulky, non-traditional moieties using side chain engineering as a powerful strategy to tailor the nanoscale morphology, optoelectronic properties, and thermomechanical behavior of semiconducting polymers for softer organic electronic applications. At the same time, our results underscore the critical need for thoughtful side chain design and moiety selection, as excessive side chain bulk can disrupt thin film morphology, hinder molecular packing, and ultimately compromise electronic performance.
![]() | ||
Fig. 1 Diketopyrrolopyrrole-based semiconducting polymers with dendritic poly(benzyl ether) side chains. |
Following purification, the new polymers were characterized using various techniques to probe the physical properties. The results are summarized in Table 1. First, size-exclusion chromatography was used to evaluate molecular weights and dispersity. P(G1-DPP-BT) was found to have a lower molecular weight (22 kDa) compared to P(G1-DPP-TT), which exhibited a significantly higher molecular weight of 96 kDa. In contrast, both generation 2 semiconducting polymers displayed similar molecular weights of approximately 50 kDa. This can be potentially attributed to solubility differences between the growing polymer chains, with P(G1-DPP-BT) chains possibly reaching their solubility limit earlier than their thienothiophene analogue, P(G1-DPP-TT), thereby restricting chain growth and leading to a lower molecular weight. Interestingly, the similar molecular weights observed for P(G2-DPP-BT) and P(G2-DPP-TT) can potentially indicate that the dendritic architecture imposes steric or solubility constraints that inherently limit chain extension, effectively normalizing molecular weights regardless of the monomer structure. The energy levels of the polymers were measured using cyclic voltammetry (Fig. S2, ESI†). All polymers showed HOMO levels ranging between −5.2 and −5.5 eV and LUMO levels between −3.9 and −4.0, which agrees with previous results obtained on similar polymer systems.46 Both G1 and G2 side chains showed very little influence on the bandgap and energy levels. Thermal decomposition temperatures were evaluated using thermogravimetric analysis at a 5% weight loss (Fig. S3, ESI†). All polymers showed high stability, with all polymers decomposing over 330 °C.
Polymer | Mna (kDa) | Mwa (kDa) | Đb | λmax(soln)c | λmax(film)c | Eopt![]() |
HOMOe (eV) | LUMOf (eV) | Tdg (°C) |
---|---|---|---|---|---|---|---|---|---|
a Number-average molecular weight and weight-average molecular weight estimated by high-temperature gel permeation chromatography in 1,2,4-trichlorobenzene at 140 °C using polystyrene as the standard.b Dispersity defined as Mw/Mn.c Absorption maxima determined in solution (2.5 × 10−4 g mL−1 in CHCl3) and the spin-coated thin film.d Calculated using the following equation: gap = 1240/λonset of the polymer film.e Calculated from cyclic voltammetry (potentials vs. Ag/AgCl) using 0.1 M TBAPF6 in CH3CN as an electrolyte.f Estimated from the calculated Eoptg and HOMO.g Determined from thermogravimetric analysis (TGA) at a 5% weight loss. | |||||||||
P(G1-DPP-BT) | 22.3 | 25.6 | 1.2 | 789 | 822 | 1.33 | −5.2 | −3.9 | 336 |
P(G2-DPP-BT) | 50.2 | 61.2 | 1.2 | 777 | 822 | 1.33 | −5.5 | −4.0 | 333 |
P(G1-DPP-TT) | 96.5 | 155.5 | 1.6 | 833 | 854 | 1.37 | −5.3 | −3.9 | 337 |
P(G2-DPP-TT) | 51.1 | 72.6 | 1.4 | 810 | 841 | 1.37 | −5.3 | −3.9 | 360 |
Following the initial characterization of the physical properties of the new semiconducting materials, UV-vis spectroscopy was performed in the solution and solid states to gain insight into the influence of the dendritic moieties and different backbone donating units on the optoelectronic properties. Notably, information about the molecular aggregation of polymer chains can be obtained by comparing the ratios of the 0-0 and 0-1 vibrational peaks, which correspond to aggregated and non-aggregated species, respectively.28 All polymers exhibited typical absorption spectra for DPP-based conjugated polymers, as shown in Fig. 2 and Table 1.28,47,48 The incorporation of generation 1 and 2 side chains significantly influenced the optical properties of the polymers in CHCl3. P(G1-DPP-BT) exhibited absorption bands centered at λ = 434 and 789 nm, while P(G2-DPP-BT) exhibited absorption bands centered at λ = 426 and 777 nm, respectively, which can be attributed to the π–π* transitions and donor–acceptor charge transfer vibrations, respectively. Similarly, P(G1-DPP-TT) and P(G2-DPP-TT) exhibit π–π* transitions at λ = 420 and 412 nm, respectively, while their donor–acceptor charge transfer vibrations appear at λ = 833 and 810 nm, respectively. The redshift between the BT and TT in the solution state can be associated with an increased effective conjugation when thienothiophene is used as a donor due to its planarity and rigidity, which has been previously observed in related systems.49 The resulting blueshifts between G1 and G2 for the polymers can be attributed to the decrease in aggregation properties in solution, likely due to the bulk of the side chain. In the solid state, polymer chain aggregation differed notably, as λmax for all four polymers increased by 20–40 nm for their donor–acceptor vibrations compared to their solution-state absorptions. Interestingly, all polymers, except for P(G1-DPP-BT), exhibited similar aggregation behaviors in solution, showing an overall decrease in aggregation in the solid state in comparison to the solution state. In the solid state, both P(G1-DPP-BT) and P(G2-DPP-BT) were shown to have a similar 0-0/0-1 peak ratio, attributed to the relative minor difference in aggregation. In contrast, P(G1-DPP-TT) and P(G2-DPP-TT) differed significantly in their solid-state aggregation properties, with P(G1-DPP-TT) displaying a much higher 0-0/0-1 peak ratio than P(G2-DPP-TT), indicating that P(G1-DPP-TT) aggregates more in the solid state than P(G2-DPP-TT). While molecular weight (Mn) effects cannot be ruled out, these findings suggest that increasing the side chain size leads to a significant reduction in solid-state aggregation, likely due to the increased steric bulk from G1 and G2. Notably, P(G1-DPP-TT) shows minimal change between its solution and solid-state UV-vis spectra, a behavior not observed for the other polymers. This suggests relatively weaker interchain aggregation or a more planar conformation already present in solution, which could reduce the need for substantial reorganization during film formation and potentially benefit charge transport in devices.
To gain further insight into the morphological properties and characteristics of the polymers, grazing incidence wide-angle X-ray scattering (GIWAXS) was performed (Fig. 4a–d). The intermolecular lamellar spacing between adjacent polymer chains was determined from the 2D scattering patterns (Fig. 4e), while the out-of-plane π–π stacking distances of the conjugated backbones are summarized in Table S1 (ESI†). The corresponding 1D scattering vector profiles are shown in Fig. S4 (ESI†). The d-spacing for P(G1-DPP-BT) is 39.03 Å, increasing to 43.63 Å for P(G2-DPP-BT). Similarly, P(G1-DPP-TT) exhibits a d-spacing of 39.03 Å, which increases to 42.45 Å for P(G2-DPP-TT), demonstrating that both the size and generation of the side chains significantly influence the molecular packing of these polymers. Notably, reducing the backbone length from BT to TT while increasing the planarity and rigidity resulted in no significant change in d-spacing within the G1-series, while there was a much more pronounced effect on the G2 series, where the d-spacing for P(G2-DPP-TT) was 1.18 Å less than that of P(G2-DPP-BT). The observed increase in d-spacing across generations can be closely correlated with the steric bulk and rotational freedom introduced by the G1 and G2 side chains. Higher-order reflections (100) were identified for all polymers and are presented in Fig. 4 and Fig. S4 and Table S1 (ESI†). Additionally, both generation 1 polymers exhibit some degree of π–π stacking in the out-of-plane direction, indicating the presence of both edge-on and face-on orientations relative to the OTS-functionalized substrate. This structural arrangement is further supported by the GIWAXS 2D profiles shown in Fig. 4.
To evaluate the mechanical properties of thin films and better understand the influence of both side chain generation and backbone incorporation, quantitative nanomechanical mapping (AFM) was utilized. The corresponding force plots and quantitative nanomechanical data are presented in Fig. S5 (ESI†). The results shown in Fig. 5a demonstrate that P(G1-DPP-BT) has a Young's modulus of 1560 MPa, which significantly decreases with increasing side chain size, as seen in P(G2-DPP-BT) having a modulus of 906 MPa. This suggests that larger side chains soften the polymer, likely due to an increase in the amorphous content and a reduction in molecular aggregation. A similar trend is observed when comparing P(G1-DPP-TT) (2698 MPa) to P(G2-DPP-TT) (2266 MPa). While this trend is consistent across the polymer series, the bithiophene-based polymers exhibit distinctly lower Young's moduli compared to their thienothiophene counterparts. This may be attributed to the effect from the backbone, as previous studies have shown that isolated thiophenes typically have lower elastic moduli compared to rigid, fused thiophene backbones. The increased stiffness and planarity of fused thiophenes often result in a higher Tg, contributing to the higher modulus observed in the thienothiophene series. These findings align with previous reports that bulky side chains reduce stiffness and enhance the softness of conjugated polymers, while rigid π-conjugated backbones increase stiffness and modulus. The observed decrease in Young's modulus with increasing side chain content highlights exciting potential for developing flexible and conformable technologies. By strategically tuning the size and intrinsic softness of side chains in conjugated polymers, it is possible to design materials with precisely controlled mechanical properties, enabling enhanced softness, stretchability, and flexibility.
The thermomechanical properties of the new dendron-bearing semiconducting polymers were further investigated using dynamic mechanical analysis (Fig. 5b and Fig. S6, ESI†). The detailed sample preparation procedures are described in the ESI.† Briefly, the polymers were drop-cast onto a meshed glass-fiber substrate as previously reported and dried overnight.45 To avoid artifacts and erase the thermal history of the sample, the polymers were subjected to a full heating–cooling cycle prior to analysis. The side chain Tg values are determined from the loss modulus, while the backbone Tg values are extracted from the tanδ curve. Previous reports indicate that the branched, alkyl side chain transitions typically occur at lower temperatures, at approximately −50 °C in DPP-based polymers, whereas the backbone transitions for DPP-based polymers are typically observed around 0 °C.45 In the current system, we found that as the side chain size increases and becomes increasingly soft across successive generations, the temperature range of the side chain transitions narrows and decreases drastically. Specifically, as the side chain generation increases from G1 to G2 for both donor series, the Tg is reduced for both the bithiophene and thienothiophene series to approximately −86 °C to −87 °C for P(G1-DPP-BT) and P(G2-DPP-BT), while the side chain transitions for P(G1-DPP-TT) and P(G2-DPP-TT) are approximately −79 °C and −83 °C, respectively. A similar trend is observed for the backbone transitions, where increasing side chain bulk results in a higher backbone Tg for both the G1 and G2 series. There is a noticeable increase in the backbone Tg of the BT series when increasing the side chain generation from G1 to G2, notably, from −3 to 4 °C. A similar trend was observed in the TT series, where the Tg occurs at −2 °C in G1 and increases to 3 °C for G2. This result confirms that increasing the bulkiness of the side chain can lead to lower Tg for side chain transitions and an increase in backbone transitions.
The fabrication details and testing procedures for the prepared OFET devices are detailed in the ESI.† Briefly, all polymers were dissolved in anhydrous chlorobenzene (5 mg mL−1) and stirred for 1 hour at 80 °C. The solution was then spin-coated onto an OTS-functionalized, bottom-gate top-contact (BGTC) silicon substrate. Mobilities were extracted by linear fitting of the IDS1/2 vs. VGS transfer curves in the saturation regime, using the following equation: IDS(sat) = (WC/2L)μsat(VG − Vth)2. The results obtained for all polymers are summarized in Fig. 6, with transfer and output characteristic curves in Fig. 6 and Fig. S7 and S8 (ESI†), respectively. All polymers tested showed typical transfer and output characteristics, except P(G2-DPP-BT), which showed non-ideal output characteristics and lower device yield (50%). P(G1-DPP-BT) and P(G1-DPP-TT) exhibited charge carrier mobilities comparable to those of other semiconducting polymers with aliphatic side chains. Among the new polymers tested, both generation 1 side chain bearing polymers showed good charge mobility. P(G1-DPP-TT) showed the highest performance, with an average mobility of 2.61 × 10−3 cm2 V−1 s−1, along with an on/off current ratio of 103 and an average threshold voltage of −18.7 V. In comparison, P(G1-DPP-BT) exhibited an average hole mobility of 6.03 × 10−4 cm2 V−1 s−1, a higher on/off current ratio of 104, and an average threshold voltage of −9.7 V. Upon increasing the side chain size from G1 to G2, a significant decline in charge carrier mobility was observed. P(G2-DPP-BT) exhibited a charge carrier mobility of 4.39 × 10−4 cm2 V−1 s−1, an on/off current ratio of 101, and an average threshold voltage of −29.7 V, while also demonstrating non-ideal device behavior. The mobility of P(G2-DPP-TT) decreased even more substantially to an average of 5.81 × 10−6 cm2 V−1 s−1, although it displayed typical output characteristics. This overall reduced performance can be attributed primarily to the structural effects introduced via side chain engineering. While the bulkier generation 2 poly(benzyl ether) side chains improve solubility and processability, they also introduce greater steric hindrance and rotational freedom, which can disrupt the conjugation length, hinder molecular packing, and impair charge transport pathways. In the case of P(G2-DPP-TT), the mobility decline is particularly pronounced. This may arise from the interplay between the bulky G2 side chains and the more planar backbone of the DPP-TT unit. The thienothiophene-based backbone generally favors stronger π–π stacking and crystallinity in the absence of steric disruption. However, the introduction of large side chains likely exacerbates steric congestion around this otherwise planar backbone, not only impeding π-stacking more severely than in P(G2-DPP-BT) but also potentially creating greater torsional disorder and severely limiting charge percolation pathways.
The observed charge carrier mobility trends align closely with the results obtained by GIWAXS; the decrease in mobility in OFETs corresponds to reduced crystallinity and increased π-stacking and lamellar d-spacing distances. P(G1-DPP-TT) shows the highest mobility, consistent with its tighter π-stacking (3.48 Å) and pronounced high-order (100) reflections, potentially indicative of a higher degree of crystallinity. This finding is in line with previous studies demonstrating that enhanced π-stacking and interdigitation promote charge carrier mobility.35,50 In contrast, the increased π-stacking and lamellar distances in P(G2-DPP-BT) and P(G2-DPP-TT) are correlated with reduced mobility and more negative threshold voltages. This decline reflects less efficient intra- and intermolecular charge transport due to the bulkier side chains, which reduce crystalline order, increase amorphous content, and compromise polymer packing efficiency in the thin films.
Footnote |
† Electronic supplementary information (ESI) available: Detailed experimental procedures for all new compounds and a complete characterization of materials and devices. See DOI: https://doi.org/10.1039/d5ma00446b |
This journal is © The Royal Society of Chemistry 2025 |