Side Chain Regioregularity Enables High-Performance and Sustainable Organic Electrochemical Transistors

基督教牧师 北京 钥匙(锁) 可穿戴计算机 中国科学院 纳米技术 材料科学 工程类 计算机科学 中国 嵌入式系统 哲学 神学 计算机安全 政治学 法学
作者
Liuyuan Lan,Junxin Chen,Huiqing Hou,Jiayao Duan,Yiming Wang,Yuze Lin,Iain McCulloch,Wan Yue
出处
期刊:CCS Chemistry [Chinese Chemical Society]
卷期号:: 1-14 被引量:1
标识
DOI:10.31635/ccschem.024.202404746
摘要

Open AccessCCS ChemistryRESEARCH ARTICLES13 Aug 2024Side Chain Regioregularity Enables High-Performance and Sustainable Organic Electrochemical Transistors Liuyuan Lan†, Junxin Chen†, Huiqing Hou, Jiayao Duan, Yiming Wang, Yuze Lin, Iain McCulloch and Wan Yue Liuyuan Lan† State Key Laboratory of Optoelectronic Materials and Technologies, Key Laboratory for Polymeric Composite and Functional Materials of Ministry of Education, Guangzhou Key Laboratory of Flexible Electronic Materials and Wearable Devices, School of Materials and Engineering, Sun Yat-Sen University, Guangzhou 510275 , Junxin Chen† State Key Laboratory of Optoelectronic Materials and Technologies, Key Laboratory for Polymeric Composite and Functional Materials of Ministry of Education, Guangzhou Key Laboratory of Flexible Electronic Materials and Wearable Devices, School of Materials and Engineering, Sun Yat-Sen University, Guangzhou 510275 , Huiqing Hou Beijing National Laboratory for Molecular Sciences, CAS Key Laboratory of Organic Solids and Laboratory of Molecular Nanostructure and Nanotechnology, Institute of Chemistry, Chinese Academy of Sciences, Beijing 100190 , Jiayao Duan State Key Laboratory of Optoelectronic Materials and Technologies, Key Laboratory for Polymeric Composite and Functional Materials of Ministry of Education, Guangzhou Key Laboratory of Flexible Electronic Materials and Wearable Devices, School of Materials and Engineering, Sun Yat-Sen University, Guangzhou 510275 , Yiming Wang State Key Laboratory of Optoelectronic Materials and Technologies, Key Laboratory for Polymeric Composite and Functional Materials of Ministry of Education, Guangzhou Key Laboratory of Flexible Electronic Materials and Wearable Devices, School of Materials and Engineering, Sun Yat-Sen University, Guangzhou 510275 , Yuze Lin Beijing National Laboratory for Molecular Sciences, CAS Key Laboratory of Organic Solids and Laboratory of Molecular Nanostructure and Nanotechnology, Institute of Chemistry, Chinese Academy of Sciences, Beijing 100190 , Iain McCulloch Department of Chemistry, Chemistry Research Laboratory, University of Oxford, Oxford OX1 3TA Andlinger Center for Energy and the Environment, Department of Electrical and Computer Engineering, Princeton University, Princeton, NJ 08544 and Wan Yue *Corresponding author: E-mail Address: [email protected] State Key Laboratory of Optoelectronic Materials and Technologies, Key Laboratory for Polymeric Composite and Functional Materials of Ministry of Education, Guangzhou Key Laboratory of Flexible Electronic Materials and Wearable Devices, School of Materials and Engineering, Sun Yat-Sen University, Guangzhou 510275 Cite this: CCS Chemistry. 2024;0:1–14https://doi.org/10.31635/ccschem.024.202404746 SectionsSupplemental MaterialAboutAbstractPDF ToolsAdd to favoritesDownload CitationsTrack Citations ShareFacebookTwitterLinked InEmail High-performance organic electrochemical transistors (OECTs) with sustainable processes are crucial for bioelectronics and integration applications, but still face challenges in molecular design, as well as solvent-device compatibility. Herein, we introduced a unique synthetic protocol focused on regioselective chemistry for the development of a donor–acceptor polymeric mixed ionic–electronic conductor (PMIEC) with a well-defined side chain arrangement and demonstrated the superiority of side chain regioregularity in enhancing OECT performance. Furthermore, we pioneered the utilization of a green solvent, 2-methyl tetrahydrofuran (MeTHF), for depositing the active OECT channel layers. We found that the regioregular copolymer exhibited over three times higher μC* of up to 810 F cm−1 V−1 s−1 compared to its regioirregular counterpart, thanks to improved crystallinity, reduced trap density of states (tDOS), and enhanced OECT hole mobility. Notably, this was achieved without the need for additional film post-treatments or specialized polymer fractionation techniques and stood among the highest values reported to date for green-solvent-processed OECTs. Our work represents a significant advancement in sustainable OECTs and highlights the importance of precise control over side chain regioregularity in developing high-performance PMIECs. Download figure Download PowerPoint Introduction Sustainable bioelectronics not only paves the way for a more technologically advanced future but also addresses the pressing need for environmental consciousness.1–5 As we delve deeper into research in this field, it becomes increasingly crucial to integrate sustainability at every stage of development—from materials selection to processing techniques and device architectures.6–8 In this respect, the emergence of organic electrochemical transistors (OECTs) has opened up exciting opportunities by offering numerous advantages over traditional electronic components such as low operating voltage, high sensitivity, flexibility, and compatibility with aqueous media.9–12 OECTs hold tremendous promise in fields such as biosensing, medical diagnostics, artificial synapses, and neuromorphic computing.13–17 To further advance these practical applications, continuous material innovation is imperative to enhance performance and functionality while implementing sustainable manufacturing processes to minimize environmental impact. OECTs utilize organic mixed ionic–electronic conductors (OMIECs) as channel materials, enabling the penetration of ions from an electrolyte into the bulk of the channel material under a gate bias. This process, known as electrochemical doping/de-doping, plays a crucial role in altering the channel conductivity. A figure of merit commonly used for performance benchmarking is the product of charge carrier mobility and volumetric capacitance (μC*).18 From the perspective of molecular design for OMIECs, the incorporation of oligoethylene glycol (OEG) side chains in conjunction with backbone engineering offers a robust approach to enhance ionic/electronic conduction in conjugated polymers (CPs), due to its synthetic versatility and ability to fine-tune material properties by tailoring the length, branching position, and composition of hydrophilic side chains.19–21 In the past few years, driven by the development of OEG functionalized CPs, OECTs have made significant strides in terms of performance and functionality, achieving impressive μC* values exceeding 500 F cm−1 V−1 s−1.22–30 It is noteworthy that the donor–acceptor (D–A) copolymers, characterized by an alternating arrangement of D and A units along their backbones, exhibit strong interchain interactions, tunable electronic properties, and versatile design options. These attributes position them as a highly promising class of mixed conductors that have garnered significant attention in the OECT community.31–40 McCulloch's group33 reported a series of D–A copolymers based on diketopyrrolopyrrole (DPP), with the glycolated polymer p(gDPP-T2) (Figure 1a) showing improved polaron delocalization and better OECT performance (μC* = 377 F cm−1 V−1 s−1). Yue's group39 synthesized a range of D–A copolymers via post-polymerization, and the utilization of g2T2-gBT4 in OECTs yielded a μC* of 359 F cm−1 V−1 s−1. Very recently, Yoon's group40 developed DPP-based D–A copolymers through hybrid side chain engineering, resulting in a high μC* of 702 F cm−1 V−1 s−1 for PDPP-4EG-based OECTs, which was further increased to 809 F cm−1 V−1 s−1 using a unidirectional film alignment technique. Though progress in exploring D–A copolymers for OECTs, the current landscape remains limited in both quantity and diversity, with device performance lagging behind that of all-donor counterparts.30 Given their highly tunable energy levels and significant promise in mitigating noncapacitive Faradaic side reactions during OECT operation,33,41 it is worthwhile to focus more on optimizing the molecular design to fully unlock the potential of D–A copolymers for OECT applications. Figure 1 | (a) The representative D–A copolymers previously reported in the literature for OECTs. (b) The regioirregular and regioregular D–A copolymers developed in this work for OECTs. Download figure Download PowerPoint Enhancing the ionic and electrical conductivities in D–A polymeric mixed conductor necessitates comprehensive considerations involving meticulous tailoring of the donor/acceptor structure, fine-tuning of molecular packing/orientation, and precise control over intermolecular interactions. Conventional wisdom suggests that manipulating regioregularity, which refers to the orientation or orderliness of asymmetric units along the backbone vector, generally exerts a positive influence on charge transport property by enhancing intermolecular interactions and reducing energetic disorder in organic field-effect transistors (OFETs) field.42–45 Keeping in mind the fact that, in OFETs, which are solely dependent on electronic processes, the arrangement of densely packed polymer chains or crystalline domains is crucial for efficient charge carrier transport. For OECTs, however, charge conduction is primarily influenced by volumetric doping/dedoping, induced by ions injection, achieving an optimal device performance relies on maintaining a delicate balance between electronic and ionic conductivities due to their intrinsic contradictions.46–49 The fundamental difference in charge transport mechanisms between OFETs and OECTs underscores the need to further elucidate the relationship between charge/ion transport and the regioregularity effect in OMIECs, which has not yet been reported. Furthermore, it is important to note that current high-performance OECT devices rely heavily on the usage of chloroform (CHCl3), a toxic solvent for processing purposes, which poses a substantial risk to both environmental and human health. In this regard, it is becoming increasingly important to consider sustainable manufacturing processes (using environmentally friendly "green" solvents) for these devices, particularly as we delve deeper into the exploration of OECT technology for bioelectronics requiring compatibility with biological systems and adherence to stringent safety standards.1,50 Hydrophilic OEG functionalized CPs are known for their high polarity, potentially conferring solubility in polar green solvents such as acetone, 2-methyl tetrahydrofuran (MeTHF), and anisole based on the principle of "like dissolves like."51 Despite these properties, there has been very limited exploration into the utilization of these green solvents for the fabrication of OECTs,52–54 mainly due to the challenge associated with the compromise between solvent compatibility and device performance. To this end, developing OMIECs that can be processed from green solvents while achieving favorable transistor characteristics is highly desired for realizing sustainable and high-throughput transistors for real bioapplications. Herein, for the first time, we exploited the regioselective chemistry within OEG-substituted benzothiadiazole (BT) and successfully developed a novel D–A polymeric mixed conductor with side chain regioregularity (R-gBT-gCPDT, Figure 1b) for OECTs. For comparison, an irregioregular counterpart (iR-gBT-gCPDT, Figure 1b) was also synthesized, sharing the same backbone but differing in the arrangement of OEG side chains on the BT moiety. Extensive investigations were conducted to thoroughly examine the influence of side chain regioregularity on their electronic, optical, charge transport, and coupled ionic–electronic transport properties. It is noteworthy that, in line with sustainable practices, MeTHF was selected as the solvent for fabricating OECT devices due to its commercial availability and biorenewable nature derived from furfural,55 and its ability to effectively dissolve both polymers (shown in Scheme 1). We found that the regular arrangement of OEG chains on BT unit along the backbone of polymer R-gBT-gCPDT caused an edge-on packing motif, enhanced crystallinity, and reduced trap density of states (tDOS), thereby resulting in a more than 6-fold increase in hole mobility up to 5.6 cm2 V−1 s−1, as compared to the regiorandom counterpart (iR-gBT-gCPDT). Impressively, R-gBT-gCPDT achieved a maximum μC* value of 810 F cm−1 V−1 s−1 in OECTs, setting a new record for D–A copolymers and ranking highest among the reported values thus far for OECTs processed using environmentally friendly solvents. This work provides valuable insights into the development of high-performance OECTs through precise control of side chain regioregularity within CP structures while prioritizing sustainable electronics through green solvent selection. Scheme 1 | The synthetic routes for copolymers iR-gBT-gCPDT and iR-gBT-gCPDT. Next to the resultant polymer lies a photograph depicting its dissolution in MeTHF at a concentration of 5 mg/mL. Download figure Download PowerPoint Experimental Methods Materials All starting materials were purchased from commercial suppliers and used as received unless stated otherwise. The detailed synthetic procedures of CPs were described in the Supporting Information. Materials characterization 1H NMR and 13C NMR spectra were recorded in deuterated solvents on a AVANCE 400 NMR Spectrometer (Bruker, Germany). 1H NMR chemical shifts are reported in ppm downfield from tetramethylsilane (TMS) reference using the residual protonated solvent as an internal standard. The number-average molecular weight (Mn) and polydispersity index of polymers were determined by a PL-GPC 220 in hexafluoroisopropanol at 45 °C using a calibration curve with standard linear polystyrene as a reference. UV–vis-NIR absorption spectra were recorded on a Cary Win spectrophotometer (Cary 5000/6000i, Agilent Technologies). Polymer thin films were deposited onto glass substrates by spin-coating using a hexafluoroisopropanol solution (5 mg mL−1) under ambient conditions. Cyclic voltammetry (CV) measurements were performed with a CHI520E electrochemical workstation (Shanghai Chenhua Instrument Co. Ltd., China) with a three-electrode single-compartment cell equipped with a cylindrical platinum working electrode as the working electrode, Pt wire as the counter electrode, and Ag/AgCl electrode as the reference electrode. 0.1 M tetrabutylammonium hexafluorophosphate (Bu4NPF6) dissolved in acetonitrile and aqueous solution of 0.1 M NaCl were used as the supporting electrolytes, respectively. Spectroelectrochemistry measurements Spectroelectrochemical measurements were performed using the Cary Win spectrophotometer (Cary 5000/6000i; Agilent Technologies) coupled with an electrochemical analyzer (Shanghai Chenhua Instrument Co. LTD., CHI520E, China) in a three-electrode setup. The polymers were deposited on indium tin oxide (ITO)-coated glass slides by spin-coating a MeTHF solution (5 mg mL−1) at 1500 rpm for 30 s and 3000 rpm for 30 s, which acted as the working electrode and were placed in quartz cuvette equipped with a custom-made Teflon lid. Pt wire and Ag/AgCl electrode were used as the counter electrode and the reference electrode, respectively. The supporting electrolyte used was a 0.1 M NaCl aqueous solution. A background measurement was taken with a clean bare ITO substrate. The indicated voltages were applied for 10 s until the current stabilized prior to the spectrum measurement. Device fabrication and characterization The glass substrates were cleaned via sonication in acetone, isopropyl alcohol, and deionized water and then dried by nitrogen flow. Gold contacts were patterned and deposited onto glass slides to form source/drain electrodes by a standard photolithographic/lift-off process and thermal evaporation, respectively. The channel length (L) and width (W) were 10 and 100 μm, respectively. After cleaning the substrate using piranha solution and subsequently blow-drying with a nitrogen gun, the polymer semiconductors were deposited via spin-coating from a MeTHF solution (5 mg mL−1) at 1500 rpm for 30 s and then at 3000 rpm for another 30 s, under ambient conditions. A 0.1 M NaCl aqueous solution was used as the electrolyte, while an Ag/AgCl pellet served as the gate electrode. A small poly(dimethylsiloxane) (PDMS) well was used to confine the electrolyte. The fabricated devices were characterized in the air at room temperature using a semiconductor parameter analyzer (B1500A, Keysight Technologies, Beijing, China) and an electrical probe station. Atomic force microscopy measurements Atomic force microscopy (AFM) measurements were performed in tapping mode, using Bruker dimension icon AFM with a silicon tip (Bruker FastScan®). The polymer films were spin-coated on ITO substrates from MeTHF solution (5 mg mL−1) under ambient conditions. The doping of polymer films was conducted in 0.1 M NaCl solution with a three-electrode system using an electrochemical analyzer (CHI520E; Shanghai Chenhua Instrument Co. LTD., China), with polymer-coated ITO as working electrode, Ag/AgCl pellet as reference electrode and a Pt wire as the counter electrode and applying a voltage at 0.7 V versus Ag/AgCl for 10 mins. Then the film was washed with deionized water and dried with nitrogen flow. The cyclic dedoping-doping of polymer films was conducted by subjecting them to a square-wave voltage switching between 0 and 0.7 V versus Ag/AgCl. Electrochemical impedance spectroscopy measurements Electrochemical impedance spectroscopy (EIS) measurements were performed using an impedance analyzer (CHI520E, Shanghai Chenhua Instrument Co. Ltd., China) with a three-electrode set-up identical to that used for CV measurements. Polymer-coated gold electrode with a defined area of 0.2 cm2 served as the working electrode. A platinum wire and an Ag/AgCl were employed as the counter electrode and reference electrode, respectively. The measurements were conducted in 0.1 M NaCl at a DC offset potential of 0.8 V and a sinusoidal alternating current (AC) amplitude of 10 mV, and the frequency range was set to span from 10 kHz to 1 Hz. Capacitance was calculated using the equation below: C = 1 2 π f × | I mg ( Z ) | where C is the capacitance, f is the frequency, and |Img(Z)| is the magnitude of the imaginary part of the impedance. The capacitance values extracted at 1 Hz were divided by film volume to determine the volumetric capacitance (C*). Grazing-incidence wide-angle X-ray scattering (GIWAXS) Samples for X-ray scattering were prepared using the same procedures for the preparation of active channel layer in device fabrication except using pure silicon wafers as the substrates. An incidence angle of 0.18° and a photon energy of 8 keV were used to record the scattering patterns. The two-dimensional (2D) GIWAXS patterns were collected from films with a resistivity of 0.001–30 Ohm cm−1. Trap density of states (tDOS) tDOS were performed on Keysight 4980A precision inductance, capacitance, and resistance (LCR) meter (Keysight Technologies, Beijing, China) and analyzed by using the thermal admittance spectroscopy (TAS) method under angular frequency-dependent capacitance measurement (0.02–2000 kHz). The energetic profile of tDOS was derived according to the following equation:56 N T ( E ω ) = − V bi q W d C d ω ω k B T where kB is Boltzmann's constant, ω is the angular frequency, C is the capacitance, T is the absolute temperature, W is the depletion width, and Vbi is the built-in potential. The applied angular frequency ω is defined by the following formula:56 E ω = k B T ln ( ω 0 ω )where ω0 is the attempt-to-escape frequency. The device architecture for TAS measurement was ITO/ZnO/Polymer/MoO3/Ag. Results and Discussion The synthetic routes for regioirregular and regioregular D–A copolymers, iR-gBT-gCPDT and R-gBT-gCPDT, are illustrated in Scheme 1, and the detailed synthesis procedures can be found in the Supporting Information. The incorporation of triethylene glycol (TEG) side chains into both polymers was aimed to facilitate ions insertion/transport while taking into account their molecular packing and solution processability. For the preparation of side chain regioirregular polymer (iR-gBT-gCPDT), the asymmetric electron-deficient compound gBT was first synthesized by a nucleophilic substitution reaction of 4,7-dibromo-5-fluoro-2,1,3-benzothiadiazole (fBT) and TEG monomethyl ether using sodium hydride as a base in anhydrous tetrahydrofuran (THF), which was then subjected to a standard Stille-coupling polymerization with electron-donating monomer 4,4-di[2-[2-(2-methoxyethoxy)ethoxy]ethyl]-2,6-bis(tributylstannyl)-4H-cyclopenta[2,1-b:3,4-b']dithiophene (CPDT-Sn2). To access the sidechain regioregular polymer (R-gBT-gCPDT), our focus was placed on the regiochemistry of gBT. We sought to understand and control the specific position where a one-fold cross-coupling reaction occurred between compound gBT and 4,4-di[2-[2-(2-methoxyethoxy)ethoxy]ethyl]-2-tributylstannyl-4H-cyclopenta[2,1-b:3,4-b']dithiophene (CPDT-Sn1). For our experimental attempts, we opted for Pd2(dba)3/P(o-tolyl)3 as the catalyst while employing THF as the solvent. The isolated product from this reaction was analyzed utilizing 1H–1H nuclear overhauser effect (NOE) spectroscopy. As shown in Supporting Information Figure S1, it revealed an absence of any cross-correlation peaks between the proton resonances of cyclopenta[2,1-b:3,4-b']dithiophene (CPDT) and BT, demonstrating that a new compound with high regioselectivity had been successfully formed by connecting the site on BT adjacent to TEG group to the CPDT moiety, namely gBT-gCPDT (as depicted in Scheme 1). Regarding the distinct site-selective reactivity of the C-Br bond on BT, we speculated that there could be an electrophilic-like reaction occurring within this system; the electron-donating effect of the TEG group made its neighboring C-Br bond more susceptible to attack. This selective reaction strategy will be helpful in the future design and preparation of regioregular mixed conductors. Subsequently, gBT-gCPDT was deprotonated by lithium diisopropylamide and then quenched with tributylstannyl chloride to afford compound gBT-gCPDT-Sn1 containing two reactive functional groups. Finally, employing Pd2(dba)3/P(o-tolyl)3 as the catalyst in toluene, the self-polymerization of gBT-gCPDT-Sn1 led to the formation of target polymer R-gBT-gCPDT, thereby ensuring a well-ordered arrangement of the side chains on BT units along the polymer backbone. Both iR-gBT-gCPDT and R-gBT-gCPDT were readily soluble in chloroform and can be effectively dissolved into the green solvent MeTHF (Scheme 1), enabling subsequent high-quality thin film preparation through the spin-coating method. The number-average molecular weights (Mn) and weight-average molecular weights (Mw) of the two copolymers were estimated by the gel permeation chromatography (GPC) using hexafluoroisopropanol as the eluent at 45 °C. The Mn values obtained were 163 and 184 kDa for iR-gBT-gCPDT and R-gBT-gCPDT, respectively, while the Mw values were found to be 222 kDa for iR-gBT-gCPDT and 306 kDa for R-gBT-gCPDT ( Supporting Information Figures S2 and S3). Density functional theory (DFT) calculations at B3LYP/6-31G(d) level were carried out to investigate the impact of side chain regioregularity on the geometric and electronic properties of these two polymers. To simplify the calculations, we substituted the long OEG side chains on the CPDT and BT building blocks with methyl and methoxy groups, respectively, while utilizing four repeating units of the polymers. To approximate a regiorandom tetramer for iR-gBT-gCPDT, a random number generator was employed to arbitrarily determine the positioning of each BT unit's side chain (right-right-left-left orientation). The optimized geometrics of the polymers are presented in Figure 2a,b. Both polymers exhibited slightly twisted backbones, as manifested by calculating the average dihedral angle (φavg) between CPDT and adjacent BT units over seven instances. The φavg value in R-gBT-gCPDT is 2.87°, smaller than that of iR-gBT-gCPDT (φavg = 3.96°), potentially conferring a conformational advantage that contributed to the formation of crystalline domains and enhanced charge transport properties. We further calculated the electrostatic potential (ESP) distribution of the two regioisomers with an electron density isosurface value set at 0.0004 a.u., as shown in Figure 2c,d, where regions of low potential are represented by red color, while areas of high potential are depicted in blue. For both neutral polymers, negative ESP values were observed to be distributed along the conjugated backbone, mainly around the nitrogen and sulfur atoms, indicating their electron-rich nature. In contrast, the electrostatic distribution maps of the oxidized counterparts iR-gBT-gCPDT+ and R-gBT-gCPDT+ exhibited significantly positive charge characteristics across the entire molecule. This is particularly pronounced in the oxygen regions of these oxidation species, indicating their ability to coordinate with negatively charged carriers in the electrolyte solution. Moreover, the calculations indicate that R-gBT-gCPDT exhibited a larger dipole moment of 8.30 Debye, in comparison to the value of 5.37 Debye for iR-gBT-gCPDT, expected to facilitate long-range intermolecular interactions.57 Furthermore, we provided the optimized geometrics of iR-gBT-gCPDT with two additional side chain orientations (left-left-left-right and left-right-left-right), as shown in Supporting Information Figure S11. The φavg values in iR-gBT-gCPDT with left-left-left-right and left-right-left-right orientations were 3.94° and 3.65°, respectively. Similar to iR-gBT-gCPDT with a right-right-left-left orientation, negative ESP values were distributed along the conjugated backbone in the neutral state for both orientations, while the electrostatic distribution maps in their oxidized state showed pronounced positive charge features throughout the molecule. Figure 2 | Optimized molecular geometries based on DFT calculations at B3LYP/6-31G(d) level for (a) regioirregular iR-gBT-gCPDT and (b) regioregular R-gBT-gCPDT. (c) Molecular ESP maps of iR-gBT-gCPDT and its oxidation species (iR-gBT-gCPDT+). (d) Molecular ESP maps of R-gBT-gCPDT and its oxidation species (R-gBT-gCPDT+). Download figure Download PowerPoint The optical properties of the copolymers were accessed through UV–vis-NIR absorption measurement and the corresponding results are summarized in Supporting Information Table S1. As depicted in Figure 3a, both regioisomers presented typical absorption profiles of D–A polymers characterized by dual-band absorptions in solution and thin films, including a narrow high-energy band in the regime of 350–510 nm attributed to π–π* transitions, and a broad low-energy band spanning from 520 to 1030 nm originating from the intramolecular charge transfer (ICT) between the donor and acceptor moieties. We observed that there was no distinct difference in the absorption maxima (λmax) between iR-gBT-gCPDT and R-gBT-gCPDT, whether in solution or solid state, with the λmax in high and low energy bands located at around 400 and 800 nm, respectively ( Supporting Information Table S1). When going from solution to film state, both polymers exhibited a small variation in λmax, yet displayed a broader low-energy band extending into the lower energy region in absorption spectra, indicating that the polymer chains were more densely packed in their solid form.58 Furthermore, the thin-film absorption onset for R-gBT-gCPDT was found to be 988 nm, slightly larger than that of iR-gBT-gCPDT (978 nm), suggesting an enhanced aggregation which might be facilitated by the regioregular arrangement of OEG side chains; based on these absorption onsets the optical bandgaps (Egopt) were calculated to be 1.27 eV for iR-gBT-gCPDT and 1.26 eV for R-gBT-gCPDT. Figure 3 | (a) UV–vis-NIR absorption spectra of copolymers iR-gBT-gCPDT and R-gBT-gCPDT in MeTHF solution and a thin film. (b) Cyclic voltammograms of polymer thin films in a 0.1 М NaCl aqueous electrolyte. UV–vis-NIR spectroelectrochemistry of polymer thin films for (c) iR-gBT-gCPDT and (d) R-gBT-gCPDT in 0.1 M NaCl aqueous electrolyte under different potentials ranging from −0.1 to 0.8 V. (e) The absolute changes in polaron absorption
最长约 10秒,即可获得该文献文件

科研通智能强力驱动
Strongly Powered by AbleSci AI
更新
大幅提高文件上传限制,最高150M (2024-4-1)

科研通是完全免费的文献互助平台,具备全网最快的应助速度,最高的求助完成率。 对每一个文献求助,科研通都将尽心尽力,给求助人一个满意的交代。
实时播报
garden发布了新的文献求助10
刚刚
我我我发布了新的文献求助10
刚刚
蟹蟹发布了新的文献求助10
1秒前
啦啦啦完成签到 ,获得积分10
1秒前
1秒前
徐墨玄完成签到,获得积分10
1秒前
1秒前
VPN不好用发布了新的文献求助10
2秒前
李爱国应助橘子海冒险采纳,获得10
2秒前
吹雪完成签到,获得积分0
2秒前
4秒前
GillianRan发布了新的文献求助10
5秒前
6秒前
DAISHU发布了新的文献求助10
7秒前
7秒前
7秒前
7秒前
L100_plus完成签到,获得积分10
7秒前
雪白问柳完成签到,获得积分10
8秒前
8秒前
大黄应助dyfsj采纳,获得10
9秒前
10秒前
李健应助APP采纳,获得10
10秒前
sine_mora完成签到,获得积分10
10秒前
平方完成签到,获得积分10
11秒前
毒蛇如我发布了新的文献求助10
11秒前
乐观帅哥完成签到,获得积分10
11秒前
高长生发布了新的文献求助50
12秒前
12秒前
13秒前
雪白问柳发布了新的文献求助10
13秒前
爱听歌雨真完成签到,获得积分10
14秒前
Winkhl发布了新的文献求助10
14秒前
15秒前
Orange应助DAISHU采纳,获得10
15秒前
16秒前
可靠半青完成签到 ,获得积分10
17秒前
17秒前
Owen应助hh采纳,获得20
17秒前
18秒前
高分求助中
Evolution 3rd edition 1500
Lire en communiste 1000
Mantiden: Faszinierende Lauerjäger Faszinierende Lauerjäger 700
PraxisRatgeber: Mantiden: Faszinierende Lauerjäger 700
Elastic local buckling behaviour of corroded cold-formed steel columns 500
A new species of Coccus (Homoptera: Coccoidea) from Malawi 500
2-Acetyl-1-pyrroline: an important aroma component of cooked rice 500
热门求助领域 (近24小时)
化学 医学 生物 材料科学 工程类 有机化学 生物化学 物理 内科学 纳米技术 计算机科学 化学工程 复合材料 基因 遗传学 催化作用 物理化学 免疫学 量子力学 细胞生物学
热门帖子
关注 科研通微信公众号,转发送积分 3180437
求助须知:如何正确求助?哪些是违规求助? 2830752
关于积分的说明 7980700
捐赠科研通 2492416
什么是DOI,文献DOI怎么找? 1329494
科研通“疑难数据库(出版商)”最低求助积分说明 635742
版权声明 602954