Rebecca F.
Meacham
a,
Heejung
Roh
a,
Camille E.
Cunin
a,
Eric R.
Lee
a,
Wenhao
Li
c,
Yan
Zhao
c,
Sanket
Samal
*ab and
Aristide
Gumyusenge
*a
aDepartment of Materials Science & Engineering, Massachusetts Institute of Technology, 77 Massachusetts Ave, Cambridge, MA 02139, USA. E-mail: aristide@mit.edu; ssamal@mit.edu
bJames Tarpo Jr. and Margaret Tarpo Department of Chemistry, Purdue University, 560 Oval Drive, West Lafayette, IN 47907, USA
cDepartment of Materials Science, Laboratory of Molecular Materials and Devices, State Key Laboratory of Molecular Engineering of Polymers, Fudan University, 220 Handan Rd, Shanghai, 200433, China
First published on 22nd October 2024
In mixed ionic–electronic conductive polymers, electronic conduction is optimal in tightly packed flat chains, while ionic conduction benefits from free volume accommodating large ions. To this end, polymers with high crystallinity are often excluded from structure–property studies of high-performing mixed conductors due to their unbalanced transport, which favors electronic charges over ionic ones. Herein, we investigated how mixed conduction can be achieved in ordered conjugated polymers by systematically combining interchain order with side chain engineering. We synthesized a series of isoindigo (IID)-based copolymers with varying amounts of aliphatic and hydrophilic side chains and examined the impact of interchain order on mixed conduction. Through crystallographic, spectro-electrochemical, and molecular dynamics studies, we demonstrated that systematically introducing hydrophilic side chains reduces the bulk order and long-range aggregation by increasing chain flexibility while preserving the interchain stacking distances within crystalline domains. Testing these IID polymers in transistor devices revealed that ion insertion and device transconductance strongly depend on the amount of hydrophilic side chains. We demonstrated that glycol side chains can enhance mixed conduction while maintaining interchain order. Our findings suggest that the IID system is promising for designing polymers that can accommodate ionic species without compromising the chain ordering required for electronic conduction.
It is difficult to concomitantly meet the requirements for ionic and electronic conduction when designing new mixed conductive polymers. Electronic conduction requires close packing and high molecular weight for charges to travel along the polymer backbone and avoid trapping in disordered regions.8,9 Ionic conduction, on the other hand, requires swelling of the polymer film upon contact with the electrolyte and having sufficient space throughout the film bulk for ions and their solvation shells.10 Swelling and deswelling of the polymer film require dynamic creation of spaces between polymer chains and can disrupt the packing needed for electronic conduction leading to decreased electronic performance. Additionally, hydrophilic side chains, such as ethylene glycol which are common in OMIECs,11–14 are flexible and can disrupt packing. It is thus important to understand how polymer morphology and interchain ordering impact mixed conduction to better inform polymer design and device performance.
One of the understudied cores in mixed conductive polymers is isoindigo (IID). This electron-deficient core is attractive particularly in donor–acceptor type of polymers due to its simple synthesis and stability in ambient conditions,15,16 but only recently it has become of interest in designing OMIEC materials.17,18 This is partly due to the highly planar nature of IID which imparts strong π–π stacking and backbone planarity into polymers, advantageous for electronic conduction but unfavorable for ion insertion.19 This strong stacking and aggregation have often excluded IID from OMIEC designs since an ideal mixed conductor must concomitantly accommodate both ionic species and electronic charge carriers. As a result, IID has been less studied compared to other less crystalline cores such as ethylenedioxythiophene (EDOT) and diketopyrrolopyrrole (DPP).19,20 With these acceptor cores, copolymerization offers a method to incorporate monomers with different side chains and create polymer structures capable of mixed conduction.21,22 A mix of hydrophobic long alkyl side chains and hydrophilic side chains such as ethylene glycol are used to impart solubility for solution processing as well as enhanced solvation of electrolyte and ionic conduction.8–11 In the case of IID, the copolymerization often leads to greater degrees of π–π stacking,23–28 which makes its use in OMIECs challenging. In fact, the successful examples of mixed conductors based on IID have been based on copolymerizing with monomers which intentionally disrupt the order and chain planarity to maximize electrolyte uptake and redox activity.17,18,29
In this work, we aimed to investigate side chain engineering as a versatile approach to modulate the electrochemical response in conjugated polymers without sacrificing their inherent interchain order. We designed a series of IID-based copolymers varying the ratio of aliphatic and hydrophilic side chains to systematically modulate bulk crystallinity and ion uptake. We used thin film morphology analyses and molecular dynamics simulations to understand the relationship between structure and crystallization behavior with increasing content of hydrophilic side chains. The IID series revealed that tethering branched glycol side chains onto backbones with inherent interchain order leads to tunable morphology and electrochemical performance resulting from (i) a change in the π–π stacking from an edge-on to a face-on orientation; (ii) reduced long-range aggregation in the film bulk; and (iii) an overall coiling of chain bundles, while conserving the π–π distance of the parent IID core within the crystalline regions. This conservation of inherent interchain distances while accommodating ions through side chains leads to enhanced transconductance in OECT devices, which proved tunable based on the amount of hydrophilic side chains present, as determined by morphology characterization and electrochemical operation in OECTs. With this study, we garnered insights that we believe will spark further investigations on highly crystalline organic mixed conductors, where the balance between interchain order to favor electronic transport and bulk ion transport is attainable through side chain engineering. Such investigations would thus tap into the unexploited potential of order polymers as mixed conductive materials.
The target mg-IID polymer series was synthesized using Stille polycondensation and obtained as dark reddish, shiny solids (full synthetic details in the ESI†). 1H NMR was used to confirm the polymer structures as well as the ratios of a-IID and g-IID incorporated (Fig. 1B). The signature chemical shifts for the alkyl side chain appear at 0.97–1.85 ppm and at 3.93 ppm, while that of glycol side chain appear at 2.55 ppm, 3.41–3.74 ppm and 4.01 ppm. As the amount of g-IID in the feedstock increased, the corresponding NMR peaks for the glycol side chain also increased, while the peaks for the alkyl side chain decreased indicating proportional and gradual incorporation of g-IID into the polymers. The ratios of a-IID and g-IID in each polymer were consistent with the feedstock ratio within 2% for 25g-IID and 50g-IID. The 75g*-IID feedstock (25% a-IID and 75% g-IID) resulted in 34% a-IID and 66% g-IID incorporated into the 75g*-IID polymer, denoted throughout as 75g*-IID, given the target ratio was 75% g-IID. The NMR peaks became sharper with increased g-IID concentration possibly due to increased solubility in the TCE-D2 solvent. To confirm the respective molecular weights, high-temperature gel permeation chromatography (HT-GPC) was used, but due to poor solubility of the mg-IID polymers containing glycol side chains in the mobile phase, the molecular weight estimation was only successful for 0g-IID. Such poor solubility is common in IID-based structures which complicates accurate characterization of the polymers.18,28 We then employed Diffusion Ordered Spectroscopy (DOSY), as described in the ESI,† to estimate the molecular weight of g-IID containing polymers. The molecular weights of the mg-IID polymers are found to be about 14–65 kg mol−1 (Table 1), which are in good agreement with previously reported values.18 Note that the DOSY method yields values as a singular average, as opposed to a conventional distribution, but it does circumvent the solubility limitations and was thus used herein to confirm the polymer structure and compare the molecular weights of mg-IID polymers.
MWa (kg mol−1) | T d (°C) | θ (°) | γ (mJ m−2) | E g (eV) | |
---|---|---|---|---|---|
a 0g-IID Mn obtained from GPC, others M obtained from DOSY. b Decomposition temperature defined as 1% mass loss. c Optical band gap calculated using Tauc plot with UV-Vis-NIR absorption spectroscopy. | |||||
0g-IID | 65.5 | 390 | 104.3 | 20.45 | 1.61 |
25g-IID | 33.7 | 358 | 99.9 | 23.06 | 1.61 |
50g-IID | 14.8 | 338 | 82.7 | 33.78 | 1.61 |
75g*-IID | 23.5 | 334 | 80.6 | 35.14 | 1.60 |
100g-IID | 34.1 | 325 | 61.6 | 46.85 | 1.60 |
All polymers were then fully characterized and key parameters are shown in Table 1. Notably, all mg-IID polymers showed little or no melting and recrystallization peaks in DSC and were thermally stable up to 300 °C as determined by TGA with decreasing decomposition temperatures as the g-IID content was increased (Fig. S1†). UV-Vis-NIR absorption spectra were recorded for thin films of the mg-IID series to probe the inter-chain order and π–π stacking of the copolymers. UV-Vis-NIR absorption spectra of all polymers showed prominent 0–0 and 0–1 vibronic peaks, characteristic of IID units (Fig. 1C).16 The high relative intensity of the 0–0 peak (∼720 nm) in the 0g-IID and 25g-IID thin films indicated strong inter-chain stacking. A decrease in intensity was observed for the 50g-IID and 75g*-IID thin films, indicating a disruption in order. As the concentration of g-IID increased, the relative intensity at ∼720 nm continued to decrease, however remained well defined even in the 100g-IID thin film. This indicates that the introduction of large, flexible glycolated side chains does disrupt the packing in polymer thin films containing g-IID as compared to the parent 0g-IID polymer thin film. However, some order in the thin film is still retained even at the highest g-IID concentration (100g-IID). UV-Vis-NIR absorption of the polymers in chloroform showed some aggregation, indicating stacking behavior even in solution. The 0–0 peak remained prominent for 0g-IID in solution, was less prominent in the 25g-IID solution, and did not appear in the 50g-IID, 75g*-IID, and 100g-IID solutions indicating that formation into thin films enhanced ordering and π–π stacking particularly for the less aggregated solutions with higher g-IID concentration (Fig. S2†). Optical band gaps were extracted from thin film UV-Vis-NIR absorption profiles and used with X-ray Photoelectron Spectroscopy (XPS) to calculate HOMO and LUMO energy levels to determine the impact of copolymerization (Fig. 1D).
X-ray diffraction patterns further revealed that glycol side chains have a significant impact on the preferential orientation of the IID backbone relative to the substrate surface. Fig. 2A illustrates different modes of packing observed in the IID polymer thin films, and Fig. 2B–F shows the respective 2D diffraction patterns of mg-IID polymer thin films. As expected, 0g-IID yielded well-ordered and highly crystalline films, indicated by high intensity diffraction peaks (between 0.1 and 1.2 Å−1) detected in the out-of-plane direction corresponding to the lamellae stacking (Fig. 2B and G). Additionally, 0g-IID chains exhibit a preferentially edge-on stacking behavior, as the in-plane direction shows no lamellae peaks, and instead, a sharp (010) peak centered at 1.75 Å−1, assigned to the π–π stacking. Additionally, the typical halo between Q values of 1.4–1.5 Å−1, assigned to the amorphous region, is very dim indicating high film crystallinity.
Compared to the parent 0g-IID polymer, the glycolated copolymers showed an overall loss in film crystallinity, as hinted by UV-Vis-NIR absorption spectra (Fig. 1C & Table S3†). With the increase in g-IID content within the polymer two key features emerged from the GIXRD data: film crystallinity decreases and the chain orientation flips. Starting with 25g-IID, the polymers exhibited a gradual decrease in the intensity of the lamellar stacking peaks in the out-of-plane direction. Additionally, the π–π stacking peak in the out-of-plane direction gradually decreases in intensity, and eventually vanishes in the case of 100g-IID (Fig. 2C–G). Concomitantly, a new (010) peak centered at 1.75 Å−1 starts to emerge in the out-of-plane direction indicating a predominantly face-on π–π stacking as the amount of glycol side chains increases. This orientation was also supported by the presence of new lamellar stacking peaks in the in-plane direction (Fig. 2C–G). It was thus obvious that the branched glycol side chains not only impacted the overall film crystallinity, but also the packing behavior between neighboring chains. Also noteworthy was the mixed orientation observed when the polar and non-polar side chains are systematically well balanced, i.e., 50g-IID and 75*g-IID thin films exhibit a mixed phase, which might be of interest for directional ion insertion.32 As ion injection is fastest in parallel to lamellar stacking, changing from edge-on to face-on can have large impacts on ionic response time depending on vertical or lateral ion injection operation, with films of edge-on morphology responding faster when ion injection is vertical.32 This ability to tune the film crystallinity, the lamellar spacing, as well as the chain orientation is of interest, especially when designing materials with co-optimized electronic and ionic transport. Note that despite the lowered degree of crystallinity and change in orientation, the π–π stacking distance (around 3.56 Å, similar to previously reported IID-based structures16,23,28,33,34) remains unchanged, and the variations are accommodated through change in lamellar spacing (Table S3†). This modular behavior thus shows the versatility of the IID core and its ability to offer tunable mixed conduction properties.
The 0g-IID polymer chains exhibited a more orderly packing arrangement, with all the chains situated in the same plane. In contrast, the 100g-IID polymer chains deviated away from the plane, leading to a coiled architecture (Fig. 3B). However, despite this deviation, they still maintain the π–π interaction, which is also consistent with GIXRD results, where a strong π–π interaction is still observed as the glycol percentage is increased even though the overall film crystallinity is decreased. Due to this reason, the 0g-IID chains arrange parallelly with tight side chain interdigitation, whereas the 100g-IID arrange themselves in a disordered and twisted manner (Fig. 3C). We believe that this disorder in the 100g-IID polymers arises due to the closely positioned branched glycol side chains near the polymer backbone with only one linker carbon, forcing the IID core to twist while maintaining the strong π–π interaction.
Further insights into ion insertion were garnered from the bleaching behavior of the polymer thin films during spectro-electrochemical analysis. We thus tracked the absorption spectra of the annealed mg-IID thin films during the cyclic voltammetry. Full absorption spectra from all polymers are shown in Fig. S8† and more details of the measurement setup can be found in the ESI.† As shown in Fig. 4B, 0g-IID thin films did not begin to bleach until >0.4 V was applied onto the polymer-coated working electrode. Even with 0.8 V applied, the thin film still showed strong absorption in the visible region. On the contrary, 25g-IID, 50g-IID, 75g*-IID, and 100g-IID thin films showed observable drop in the absorption peak intensities with as low as 0.2 V, and significant bleaching above 0.4 V, indicative of low on-set potentials when the glycol content is increased in the polymer thin film, as discussed above. Note that the bleaching in the visible region was accompanied by the rise in the NIR absorption peaks associated with the formation of polarons in all polymers upon electrochemical doping (Fig. S8†). We could thus conclude that increasing the amount of hydrophilic side chain in the polymer enables greater ion insertion, as indicated by the bleaching behavior and the formation of polarons, even in highly stacked cores such as IID. This is further supported by ionic conductivity of each film determined through electrochemical impedance spectroscopy (EIS) (Fig. S9†). The ionic conductivity at 1 kHz of the 0g-IID was 1.73 × 10−8 S cm−1 and was highest in the 100-gIID film at 4.22 × 10−8 S cm−1 showing a 4× increase with the presence of branched glycol side chains as compared to the branched alkyl side chains present in the 0g-IID film. The hydrophobic alkyl side chains increase the resistance to ion insertion in the films, and increasing the amount of glycol side chain decreases the resistance, leading to higher ionic conductivity.
As channel materials in OECT devices, all mg-IID polymers exhibited relatively modest performances as shown in Fig. 4C & D. Details on the devices fabrication and characterization are provided in the ESI and corresponding device metrics are summarized in Table S3.† Also found in the ESI are the characteristic transfer curves, exhibiting conventional p-type turn-on behaviors in all polymers (Fig. S10†). OECT device operation varied depending on hydrophilic sidechain concentration within the polymer. Higher voltage was required to turn on devices with low g-IID concentration, consistent with the spectro-electrochemical analysis indicating higher voltage needed for ion insertion (Fig. 4A–D). Increasing the concentration of glycol side chains increased the device current and transconductance. Increasing the amount of g-IID increased device performance with the 100g-IID film reaching 100 times more device current than 25g-IID and the maximum transconductance of the series studied (Fig. 4C, D & Table S3†). Higher concentration of g-IID increased ion insertion due to the hydrophilicity and disorder created by the flexible chains. Note that as expected, the 0g-IID and 25g-IID films did not show high performance (compared to state-of-the-art and less crystalline counterparts achieving μC* up to 522 F cm−1 V−1 s−1 (ref. 4 and 20)) in OECTs due to the lack of space for ions in highly ordered regions but still showed to outperform previously reported IID-based analogues, achieving comparable normalized transconductance17 and improved μC*.18
To further quantify the impact of interchain order on mixed ionic-electronic conduction, the characteristic metrics, namely μC*, were extracted from OECT measurements, where μ is the charge carrier mobility and C* is the volumetric capacitance. More details on the extraction and calculations of the device metrics are available in the ESI.† As mentioned above, the mg-IID series performed rather modestly; the extracted μC* values could only reach as high as 12.26 F cm−1 V−1 s−1 as the g-IID ratio increased to 100%. However, as we hypothesized and in agreement with our morphological and electrochemical results, the ionic-electronic conduction increases with the glycol content, without sacrificing electronic conduction (Table S3†). We calculated C* values from electrochemical measurements discussed above and extracted OECT mobilities to decouple the contributions from the two modes of transport across the IID series. The extracted OECT mobility was used as a parameter to compare electronic conduction due to low performance in dry-state OFET devices. Among the polymer series, 75g*-IID film shows the most balanced performance, with the highest OECT mobility and second highest C*. We found that by increasing the concentration of glycol side chains, the volumetric capacitance increases by over 10× upon the incorporation of polar and branched sidechains, while minimal changes occur in the extracted μOECT, particularly when comparing the 0g-IID and 100g-IID films (Table S3†). This increase was in contrast with the sidechain-insensitive π–π stacking distances as revealed by the GIXRD results. Instead of driving the chains further apart,18 the chain coiling and lowered long-range aggregation (upon the incorporation of polar glycol sidechains) improve the ability of IID films to uptake ions more efficiently. This is in contrast with the common understanding for optimizing redox activity where the addition of polar sidechains tends to greatly disrupt chain-stacking, or the parent systems are designed to be inherently amorphous. Here, to accommodate for the ionic species, the IID chains can maintain their favored spacing, and instead adapt to a more coiled configuration. As result, the thin films reveal a face-on orientation and an overall lowered crystallite aggregation. Such chain flexibility and in-bulk reconfiguration, associated to the incorporation of the branched glycol sidechains, could thus rationalize the observed mixed conduction behaviors across the IID series, albeit the conserved and close interchain distances.
Footnote |
† Electronic supplementary information (ESI) available: Experimental methods, synthetic procedures for monomers and polymers, 1H NMR spectra, thermogravimetric analysis, differential scanning calorimetry, UV-Vis-NIR solution absorption, DFT simulations, atomic force microscopy, grazing incidence wide angle X-ray diffraction spacing, cyclic voltammetry, spectroelectrochemical analysis, electrochemical impedance spectroscopy, OECT characteristic transfer curves. See DOI: https://doi.org/10.1039/d4lp00272e |
This journal is © The Royal Society of Chemistry 2024 |