Next Article in Journal
A New Strategy to Improve the Toughness of Epoxy Thermosets—By Introducing Poly(ether nitrile ketone)s Containing Phthalazinone Structures
Next Article in Special Issue
Defect Detection and Depth Estimation in Composite Materials for Pulsed Thermography Images by Nonuniform Heating Correction and Oriented Gradient Information
Previous Article in Journal
Effect of Random Lateral Ballast Resistance on Force-Deformation Characteristics of CWR with a Small-Radius Curve
Previous Article in Special Issue
Fiber-Bragg-Grating-Based Displacement Sensors: Review of Recent Advances
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Side-Chain-Assisted Transition of Conjugated Polymers from a Semiconductor to Conductor and Comparison of Their NO2 Sensing Characteristics

Department of Organic and Nano System Engineering, Konkuk University, 120 Neungdong-ro, Seoul 05029, Republic of Korea
*
Author to whom correspondence should be addressed.
Materials 2023, 16(7), 2877; https://doi.org/10.3390/ma16072877
Submission received: 7 March 2023 / Revised: 31 March 2023 / Accepted: 3 April 2023 / Published: 4 April 2023
(This article belongs to the Special Issue Application of Emerging Materials for Advanced Imaging and Sensing)

Abstract

:
To investigate the effect of a side chain on the electrical properties of a conjugated polymer (CP), we designed two different CPs containing alkyl and ethylene glycol (EG) derivatives as side chains on the same conjugated backbone with an electron donor-acceptor (D-A) type chain configuration. PTQ-T with an alkyl side chain showed typical p-type semiconducting properties, whereas PTQ-TEG with an EG-based side chain exhibited electrically conductive behavior. Both CPs generated radical species owing to their strong D-A type conjugated structure; however, the spin density was much greater in PTQ-TEG. X-ray photoelectron spectroscopy analysis revealed that the O atoms of the EG-based side chains in PTQ-TEG were intercalated with the conjugated backbone and increased the carrier density. Upon application to a field-effect transistor sensor for PTQ-T and resistive sensor for PTQ-TEG, PTQ-TEG exhibited a better NO2 detection capability with faster signal recovery characteristics than PTQ-T. Compared with the relatively rigid alkyl side chains of PTQ-T, the flexible EG-based side chains in PTQ-TEG have a higher potential to enlarge the free volume as well as improve NO2-affinity, which promotes the diffusion of NO2 in and out of the PTQ-TEG film, and ultimately resulting in better NO2 detection capabilities.

1. Introduction

Conjugated polymers (CPs) are promising active materials that can be used in various electronic devices, such as organic photovoltaic cells, field-effect transistors (FETs), and thermoelectric devices [1,2,3,4,5,6]. The electrical properties of CPs strongly depend on their chemical structures, which determine the degree of p-orbital overlap in the conjugated frameworks [7,8,9]. Recently, the insertion of electron-accepting dopants into electron-donating semiconductors has been reported to efficiently induce electron transfer or form a charge transfer (CT) complex, resulting in electrical conductors [10]. However, because infinite p-orbital overlap through conjugated structures is entropically unfavorable, most linear CPs exhibit a kinked chain configuration, resulting in semiconducting properties unless charge carriers are intentionally generated via chemical or electrical doping processes [11,12,13,14]. Few examples of CPs exhibiting essentially conductor-like properties have been reported. As a result, the design strategies for devising conductive CPs have not been fully established.
CPs have received great attention as versatile gas detection platforms because of their ease of structural modification to improve the detection suitability for target gases [15,16,17]. For example, NO2 is a toxic gas that is abundantly released from industrial sources, and its strong oxidizing properties can critically damage the human respiratory system. Therefore, the accurate and fast detection of NO2 is necessary to ensure workplace safety. Recently, our research group reported that EG-based side chains have good affinity for polar NO2 molecules, providing better and faster NO2 detection properties [18]. The electronic performance of CP can be dramatically altered by side-chain engineering without modifying the conjugated backbone [19,20,21,22,23]. Therefore, the design of a side chain capable of specific interaction with an analyte to improve the detection sensitivity of CP-based electronic sensors is in high demand. In addition, since the porous morphology of CP can facilitate gas diffusion and enlarge the reactive surface area [24,25,26,27,28], CP films with porous structures have been reported to provide enhanced gas detection capabilities [29,30].
CP-based electronic sensors have two different device configurations: chemiresistive- and FET-type sensors. Chemiresistive sensors utilize a change in resistance to detect the target gas; therefore, CP as a detection platform should have electrical conductor-like properties [31,32,33]. FET-based sensors, on the other hand, adopt semiconducting CPs to monitor current changes due to charge carrier mobility and/or concentration, which enable the delivery of multiparameter response characteristics, such as charge–carrier mobility, threshold voltage, on/off current, and conductivity [34,35,36]. Various CPs have been implemented in chemiresistive- and FET-type electronic sensors, and their advantages have been reported in terms of detection sensitivity, selectivity, and fast recovery. For example, an ion-in-conjugation polymer, p-polyphenyl squaraine (p-PPS), has been adopted as a chemiresistive sensor for NO2 gas detection and exhibited a high sensitivity of 100 ppb with a detection limit of 40 ppt [37]. In another example, a diketopyrrolopyrrole (DPP)-based CP bearing NO2-affinitive ethylene glycol (EG) side chains has registered sub-ppm detectivity when adopted in FET-type sensors [18]. However, it is unreasonable to directly compare the gas detection performances of chemiresistive- and FET-type sensors because their working mechanisms and structures of the applied polymers are different.
In this study, we designed two different CPs that share the same thiadiazolo–quinoxaline-based conjugated backbone with different side chains. In the designed CPs, alkyl- and EG-based side chains were introduced into the same quinoxaline moiety, and their electrical properties were characterized using a FET-based electrode configuration. While PTQ-T containing alkyl side chains exhibited typical p-type semiconductor-like behavior, PTQ-TEG having EG-based side chains showed a conductor-like current-voltage (I-V) response. The difference in the electrical properties of the two CPs was discussed by comparing their molecular properties, such as molecular energy level, film morphology, and atomic binding energy. Finally, the NO2 detection performances of PTQ-T and PTQ-TEG were compared by applying them to chemiresistive- and FET-type sensors, respectively. Although both CPs have similar chemical structure, it was found that PTQ-TEG implemented in the chemiresistive sensor marked higher NO2 sensitivity and fast recovery characteristics.

2. Materials and Methods

2.1. Materials

The reagents used in this study were purchased from commercial suppliers (Sigma-Aldrich, Tokyo Chemical Industry, and Alfa Aesar) and used without further purification. Compounds 1a and 1b were synthesized similarly to the previously reported methods [18,38]. The chemical structure of CP was confirmed using Fourier transform infrared (FT-IR) and proton nuclear magnetic resonance (1H-NMR) spectroscopy (Figures S1 and S2). The molecular weight and polydispersity index (PDI) were determined by gel permeation chromatography with chloroform as an eluent (Figure S3).

2.2. PTQ-T and PTQ-TEG

In a dried Schlenk flask, 1a (0.5 g, 0.40 mmol) and 2,5-bis(trimethylstannyl)thiophene (0.16 g, 0.40 mmol) were dissolved in anhydrous xylene (2 mL) under an argon atmosphere. After adding tris(dibenzylideneacetone)dipalladium (2.2 mol%) and tris(o-tolyl) phosphine (3.2 mol%), the mixture was stirred in a microwave reactor at 140 ℃ for 2 h. Then, 2-tributylstannyl thiophene (5 mol%) and 2-bromothiophene (5 mol%) were added at 30 min intervals to terminate polymerization. The solution was then poured into methanol, and the precipitate was collected using a nylon membrane filter. To remove the low molecular weight portion, the obtained solid was further purified by Soxhlet extraction using methanol, acetone, hexane, and chloroform. The portion soluble only in chloroform was reprecipitated in methanol, and the final PTQ-T was collected in 75% yield by filtration using a nylon membrane: Mw (18.7 kDa) and PDI (3.10). Similar to the method for the polymerization of PTQ-T, PTQ-TEG was obtained in 68% yield through copolymerization with 1b (0.5 g, 0.40 mmol) and 2,5-bis(trimethylstannyl)thiophene (0.16 g, 0.398 mmol): Mw (16.3 kDa) and PDI (2.52).

2.3. FET and Gas Sensor Fabrication

A heavily n-doped silicon wafer containing a SiO2 layer (300-nm-thick with a capacitance of 10.8 nF/cm2, Fine Science) was cut into 2 cm × 2 cm pieces. After cleaning the pieces by ultrasonication in acetone and isopropyl alcohol for 20 min, they were dried under N2 flow and then subjected to a 20 min UV–ozone treatment to modify the silicon wafer surface. After forming a self-assembled monolayer of octadecyl trichlorosilane on the substrate, each CP solution dissolved in chloroform (5 mg/mL) was spin-cast onto the self-assembled monolayer at 1500 rpm for 60 s. The formed CP film was dried in a vacuum oven at 25 °C for 4 h to remove residual solvent. To investigate the electrical properties of obtained CP thin films (57.0 ± 5.0 nm), source and drain electrodes (Au, 50 nm) with a channel length of 100 µm and a width of 2000 µm were thermally deposited on the CP layer. The FET active channel was placed in a gas sensor device and wired with Ag wire.

2.4. Electrical and Gas Sensing Properties

The electrical characteristics of the FET devices were measured using a Keithley 4200-SCS semiconductor parameter analyzer (Keithley, Cleveland, OH, USA) connected to a probe station (MS TECH, Suwon, Republic of Korea). To obtain the transfer curve, the gate bias (VG) was swept from 40 V to −80 V in −1.0 V increments, while the source-drain voltage (VDS) was fixed at −80 V. The output curve was collected by sweeping VDS from 0 to −80 V in −1.0 V increments while fixing VG at 0 V, −20 V, −40 V, −60 V, and −80 V, respectively. The gas-sensing properties of CPs were measured using a gas sensor (Precision Sensor System Inc., Daejeon, Republic of Korea). The applied voltage (VG and VDS) was fixed at −10 V in the FET-type sensor, and the resistor-type sensor applied VDS (−10 V) at zero VG to measure the sensitivity to the analyte gas. The gas detection sensitivity was identified as the average value through repeated exposure to the analyte gas (50 ppm) for 50 s and filling with N2 three times for 900 s.

2.5. Characterization

Polystyrene was used as a standard for gel permeation chromatography (Waters, Worcester County, MA, USA) to determine the molecular weights of the obtained CPs. UV-visible (UV-vis) absorption spectra of CP were characterized using UV-vis spectroscopy (Agilent, Santa Clara, CA, USA) in solution (chloroform) and film states, respectively. Ultraviolet photoelectron spectroscopy (UPS, Riken, Tokyo, Japan) was performed to determine the energy levels of the CPs. The surface morphologies of CP thin films were characterized using atomic force microscopy (AFM, Park Systems, Suwon, Republic of Korea) in a non-contact mode. The degree of CP chain assembly in the film state was analyzed using 2D-GIXRD (Xenocs, Grenoble, France), and the atomic binding energies of the CPs were determined using X-ray photoelectron spectroscopy (XPS, Thermo-Fisher, Seoul, Republic of Korea). All measurements using the CP thin films were performed with the same thickness (57.0 ± 5.0 nm) applied to the gas sensor device.

3. Results and Discussion

As shown in Scheme S1 and Figure 1a, two CPs containing the same thiadiazolo-quinoxaline-based conjugated backbone but different side chains were designed to compare their electrical properties. Both CPs were obtained through a Stille-type cross-coupling reaction between thiadiazolo-quinoxaline and thiophene monomers, and their chemical structures were confirmed using FT-IR and 1H-NMR spectroscopy. Both CPs exhibited the same skeletal vibrations corresponding to thiophene (1603 cm−1) and quinoxaline (1513 cm−1) moieties in the FT-IR spectra. However, the stretching vibrations corresponding to C-H (2700–3000 cm−1) and C-O-C (1025–1100 cm−1) were clearly distinguished (Figure S1) because the side chains introduced into each CP are different. In addition, judging from the peak position and the integral ratio of aromatic and aliphatic protons in the 1H-NMR spectra (Figure S2), it was confirmed that polymerization was carried out successfully. Specifically, the integral ratio of aromatic and aliphatic protons was 1:7.1 for PTQ-T and 1:4.5 for PTQ-TEG, which was consistent with the theoretical values at a similar level (1:5.9 for PTQ-T and 1:4.1 for PTQ-TEG). The electrical properties of CPs have been known to be partially affected by the side chain. For example, the aggregation and chain assembly propensities of CPs depend on the side chains introduced onto the conjugated backbone, which critically affect the charge-carrier mobility of the CPs [7]. To compare the electrical properties of the CPs, alkyl and flexible EG-based side chains were introduced into the same conjugated backbone. In the conjugated framework of the obtained CPs, the thiadiazolo-quinoxaline derivative is a strong electron-accepting moiety, and the connected thiophenes have electron-donating characteristics. Therefore, the obtained CPs have an electron donor-acceptor (D-A) chain configuration. The typical characteristics of D-A type CPs include a bimodal-shaped absorption spectrum and narrow energy band-gap [39,40]. It has been known that, in the bimodal-shaped absorption spectrum of the D-A type CPs, the absorption in the shorter wavelength region is from π–π* transitions, and the red-shifted absorption originates from the intramolecular charge transfer (CT) between electron donating and accepting moieties that are covalently interconnected in the conjugated skeleton [41].
As shown in Figure 1b, when the absorption tendencies of the obtained CPs were characterized using UV-vis absorption spectroscopy, although both CPs have the same conjugated backbone, PTQ-T exhibited more red-shifted absorption than PTQ-TEG in both solution and film states. In general, alkyl side chains have a relatively rigid nature compared to EG-based side chains due to higher rotation barrier energy. Therefore, introducing alkyl side chains into CP (PTQ-T in this study) is likely to enhance the chain stiffness, which can promote non-covalent interactions between the CP chains and result in red-shifted absorption compared to analogous CPs with EG-based side chains [34,42]. Interestingly, although both CPs exhibited bimodal-shaped absorption spectra commonly seen in D-A type CPs [43,44], the absorption in the film state was found to be more red-shifted than that in the solution state, only in the longer wavelength region. This can be attributed to facile intramolecular CT interactions between D-A moieties via partial chain planarization or enhanced non-covalent interactions of CP chains in the film state [45]. When the highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) of each CP were identified using AC2 and the absorption edge, both CPs were confirmed to have similar molecular energy levels with an extremely narrow band-gap of 0.9 eV (Figure 1c). The narrow band-gap in both CPs implies the strong electron-accepting property of the thiadiazolo-quinoxaline unit because the band-gap of quinoxaline-based CPs with similar structures commonly exceeds 1.0 eV [46,47].
Although PTQ-T and PTQ-TEG share the same conjugated framework, they exhibited completely different electrical properties. To compare the electrical properties of the CPs, a FET device with a bottom-gate top-contact structure was fabricated. The PTQ-T-based FET showed a typical p-type charge-transport behavior, and the hole mobility was determined to be 0.002 cm2/V·s from the charge-carrier transfer curve (Figure 2a and Figure S4a). However, as shown in Figure 2b, PTQ-TEG exhibited a considerably high current level with no off-current region in the VG sweep from −80 V to 40 V. In the output curve, even when VG was 0 V, PTQ-TEG exhibited a proportional trend in the output current with increasing VDS (Figure S4b), indicating electrically conducting characteristics. Typically, the electrical conductivity of semiconducting CPs does not exceed 10−7 S/cm because of the insufficient charge-carrier density unless they are chemically or electrically doped [48]. However, the obtained conductivity of PTQ-TEG was 3.0 × 10−4 S/cm, which is too high of a value for PTQ-TEG to be labeled as a conventional semiconducting CP.
The distinct electrical properties of PTQ-TEG, compared to PTQ-T, could be due to the contribution of EG-based side chains because both CPs have the same conjugated framework. To ascertain the effect of the EG-based side chain on the electrical properties of PTQ-TEG, the film morphologies and chain assembly features of both CPs were characterized using AFM and 2D-GIXRD, respectively, because the electrical properties of CPs are known to be sensitively affected by the degree of chain aggregation and assembly [49,50]. As shown in Figure 3a,b, both CPs exhibited featureless smooth surfaces with root-mean-square (RMS) roughness values of 0.478 nm and 0.355 nm for PTQ-T and PTQ-TEG, respectively. Compared to PTQ-T, PTQ-TEG showed a slightly reduced RMS roughness. The reduced RMS roughness can be attributed to the enhanced fluidity of the EG-based side chain, which has a lower rotational barrier energy than the alkyl side chain. In addition, the chain assembly tendencies of the CPs in the film state were examined using 2D-GIXRD. As shown in Figure 3c,d, both CPs exhibit diffused diffraction tendencies, indicating amorphous-like weak crystalline characteristics. When comparing the diffraction intensities in the low-q region of both CPs (Figure S5a,b), the diffraction pattern in the in-plane direction (xy-axis, qxy) was confirmed to be more prominent than that in the out-of-plane direction (z-axis, qz), implying that both CPs prefer face-on-type chain assemblies in the film state. AFM and 2D-GIXRD measurements indicated that the CPs had similar morphologies, including a chain assembly tendency, which implies that the difference in the electrical properties between PTQ-T and PTQ-TEG did not originate from the film morphology.
To confirm whether the difference in the electrical properties of PTQ-T and PTQ-TEG was related to their inherent chemical structures, the binding energies of the elements present in the CPs were analyzed using XPS. As shown in Figure 4, when the binding energies of the C, N, and O atoms were compared, the binding-energy patterns of the N atoms were almost the same in both CPs; however, the binding energies of C and O were clearly distinguished. As shown in Figure 4a, the binding-energy difference between the C atoms can be inferred to be due to the C atoms in the EG-based side chain introduced in PTQ-TEG. The C atom bonded to the O atom in the EG-based side chain of PTQ-TEG partially loses electrons and shows a relatively high binding energy because the O atom is more electronegative than the C atom. However, it can be clearly noted that the binding energy of O atoms in PTQ-TEG has shifted to the higher binding-energy region compared to PTQ-T. Upon further analysis of the binding-energy distribution of the O atoms in PTQ-TEG, the binding-energy spectrum could be separated into two different types of O atoms. One type showed the same binding energy as PTQ-T, whereas the other type had a higher binding energy. Furthermore, as depicted in Figure 4b, the O atoms with a higher binding energy (C-O*) occupied a larger portion than the neutral O atoms (C-O) in PTQ-TEG. The shift to a higher binding energy for the O atoms implies that they donate electrons. Although the exact interpretation of the binding-energy shift is limited, charge carriers were likely to be generated, because of electron donation by the O atoms in the EG-based side chains, resulting in the electrical conducting properties of PTQ-TEG.
To compare the carrier densities quantitatively, the polaron spin densities of both CPs were measured using electron spin resonance (ESR). As shown in Figure 5, both CPs clearly show similar ESR signals without any artificial charge generation treatment, such as doping. Radical cations and anions have been reported to be generated by CT interaction when strong electron donating and accepting molecules are electronically intercalated [10]. Radical generation via CT interactions can also occur when D-A type structures form within a conjugated framework. For example, when pure organic materials contain strong electron acceptors in their conjugated structures, radicals have been generated via CT interactions and exhibit organic magnetism [51,52]. Therefore, it can be speculated that the ESR signals in both CPs originate from the strong intramolecular CT interaction between the strong electron accepting thiadiazolo-quinoxaline and electron donating thiophene, because similar ESR signals appear in both CPs with the same conjugated framework. In addition, the asymmetric ESR signals appearing for both CPs indicate that they are the result of multiple radical species rather than a single type of radical. When further analyzed using a Lorentz fitting, the ESR signal was deconvoluted into two different ESR signals with g-factors of 2.0100 and 2.0027. A g-factor of 2.0100 is comparable to that of radical anions, as commonly shown in organic molecules containing strong electron-accepting moieties, and a g-factor of 2.0027 coincides with unstable carbon radicals [35,37]. When the area of each ESR component was compared through double integration and calibration with a 2,2,6,6-tetramethylpiperidine-1-oxyl free radical (TEMPO) solution (Figure S6), PTQ-TEG exhibited a higher spin density than PTQ-T. In particular, the difference in the electrical properties of the CPs can be interpreted by comparing the g-factor 2.0027 component in each ESR signal because the g-factor of the carbon radical is comparable to that of a free electron [53]. When comparing the spin density corresponding to free electrons (g-factor of 2.0027), PTQ-TEG exhibited a 4.5 times higher value than PTQ-T. Therefore, PTQ-TEG has sufficient potential to exhibit higher electrical conductivity than PTQ-T, which could be the reason for its electrical conducting properties.
Because NO2 has strong electron-accepting properties, its detection generally utilizes the CT interaction with electron-donating CPs, which can generate hole carriers and increase the current level in electrical sensors, such as FET- and resistive-type electrical devices [18,37]. Because PTQ-T and PTQ-TEG exhibit semiconducting and conducting properties, respectively, their NO2 sensing performances were compared using an FET-type sensor for PTQ-T and a resistive-type sensor for PTQ-TEG. When the NO2 detection capability was evaluated by repeatedly injecting NO2 (50 ppm) for 50 s and N2 for 900 s, the source-drain current (IDS) of both CPs markedly increased. As shown in Figure 6a, when repeatedly exposed to NO2 (50 ppm) under VG and VDS of −10 V, PTQ-T in a FET-type senor exhibited a gradual increase in current change upon each NO2 exposure, indicating that the NO2 from the previous exposure did not completely escape during N2-filling for 900 s. The IDS(t)/IDS(0) value of PTQ-T did not fully recover to its initial value during N2 charging, indicating that NO2 was likely trapped in the PTQ-T film. In contrast, in the case of PTQ-TEG applied in a resistive-type sensor, the IDS quickly recovered to its initial level when exposed to N2, demonstrating uniform detectivity during repeated NO2 sensing evaluations. Although NO2 traps in CP chains may overestimate the NO2 detectivity of the PTQ-T-based FET sensor, the NO2 detection capabilities of both CPs were quantified using the average values of sensitivity and recovery for three repeated NO2 detection experiments. PTQ-TEG applied to the resistive-type sensor exhibited a detectivity of 6.9%/ppm and recovery of 96%, which exceeded the detectivity and recovery of PTQ-T applied to the FET-type sensor (4.0%/ppm and 31%). The better NO2 detection characteristics of PTQ-TEG, compared with those of PTQ-T, can be attributed to the difference in the chemical structure rather than the type of electrical sensor applied to each CP. As shown in Figure S7, when the surface energies of both CPs were determined by measuring contact angles using three different solvents (water, diiodomethane, and glycerol), PTQ-TEG (35.16 mN/m) was confirmed to have a higher surface energy than PTQ-T (21.41 mN/m). This result indicates that PTQ-TEG has a higher affinity to polar NO2. In addition, compared to the alkyl side chain in PTQ-T, the EG-based side chain can increase the free volume of the PTQ-TEG film because of its more flexible nature. Indeed, PTQ-TEG exhibited stronger amorphous hollow diffraction than PTQ-T in the high-q region of 2D-GIXRD (Figure 3d and Figure S5c). Therefore, the high affinity for NO2 and the enlarged free volume of PTQ-TEG facilitated NO2 diffusion into the resistive-type sensor, resulting in better NO2 detection ability, including faster recovery, compared to PTQ-T.
The selectivity and limit of detection (LOD) are important parameters of gas sensors. To demonstrate that the EG-modified polar side chain of PTQ-TEG did not impair the NO2 detection selectivity for other polar gases, the gas-sensing characteristics of the resistive sensor adopting PTQ-TEG were further evaluated using SO2, NH3, and CO2 with electron-withdrawing, electron-donating, and non-polar neutral properties, respectively. As shown in Figure 7a and Figure S8, under the same conditions as NO2 detection, the detection sensitivities were in the following order: NO2 (6.9%/ppm), CO2 (0.19%/ppm), NH3 (0.05%/ppm), and SO2 (0.03%/ppm). This result clearly indicates that the introduced EG-based side chains in PTQ-TEG only marginally impaired the selective detection of NO2. Interestingly, while most CP-based electrical sensors respond sensitively to NH3, the devised resistive sensor adopting PTQ-TEG exhibited a negligible response to NH3. Additionally, when the theoretical LOD was determined by exposure to NO2 at concentrations of 100, 300, 600, and 900 ppb, the PTQ-TEG resistive sensor responded linearly to the NO2 concentration (Figure S9). From the signal calibration with the signal-to-noise ratio and RMS noise, the slope extracted from the linear curve fitting using the obtained IDS(t)/IDS(0) value of each NO2 concentration marked a theoretical LOD of 1.59 ppb (Figure 7b). The extracted LOD was comparable to that of the most sensitive electrical NO2 sensor that adopted crystalline CPs with good electrical properties [31].

4. Conclusions

Two CPs with the same conjugated backbone containing different side chains were designed to investigate the effect of the side chain on their electrical properties. Although both CPs exhibited similar smooth morphologies and weak crystalline chain assemblies in the film state, they exhibited completely different electrical properties. Specifically, PTQ-T with alkyl side chains showed typical p-type semiconducting characteristics, whereas PTQ-TEG with EG-based side chains exhibited electrical conducting behaviors. It was confirmed that, although both CPs have radical species owing to their strong D-A type conjugated structure, the O atoms of the EG-based side chains can additionally intercalate with the conjugated backbone, increase the carrier density, and ultimately generate the conductor-like properties of PTQ-TEG. When PTQ-T and PTQ-TEG were applied to FET- and resistive-type sensors, respectively, PTQ-TEG exhibited higher NO2 sensitivity with a faster recovery tendency than PTQ-T. The flexible EG-based side chains increased the free volume of the CP chains as well as the affinity with polar NO2 molecules, which facilitated NO2 diffusion in and out of the PTQ-TEG film, resulting in better sensitivity to NO2 than PTQ-T. In addition, it was confirmed that the EG-based side chains in PTQ-TEG barely impaired the detection selectivity for other common gases, such as SO2, NH3, and CO2.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma16072877/s1, Scheme S1: Detailed chemical structures and polymerization scheme of CPs; Figure S1: 1H-NMR spectra of each CP; Figure S2: FT-IR spectra of each CP; Figure S3: Molecular weight of obtained CPs, which were determined using gel permeation chromatography with elution of chloroform; Figure S4: Output curve characteristics of both CPs according to VG in FET devices; Figure S5: 1D XRD diffractograms; Figure S6: ESR intensity according to TEMPO concentration; Figure S7: contact angle information used to determine the surface energy of each CP; Figure S8: Response pattern of PTQ-TEG resistive sensor to common gases; Figure S9: Response pattern of PTQ-TEG resistive sensor according to NO2 concentration.

Author Contributions

Y.A. characterized the performance of electrical sensors and prepared the manuscript. S.H., H.K. and M.S.K. synthesized side chains, monomers, and polymers together. W.H.L. discussed the result of gas sensor properties, and B.-G.K. directed this research. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable, all data obtained from this study are already given in the article.

Acknowledgments

This work was supported by Konkuk University in 2020.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liang, S.; Jiang, X.; Xiao, C.; Li, C.; Chen, Q.; Li, W. Double-Cable Conjugated Polymers with Pendant Rylene Diimides for Single-Component Organic Solar Cells. Acc. Chem. Res. 2021, 54, 2227–2237. [Google Scholar] [CrossRef] [PubMed]
  2. Zhu, L.; Zhang, M.; Zhong, W.; Leng, S.; Zhou, G.; Zou, Y.; Su, X.; Ding, H.; Gu, P.; Liu, F.; et al. Progress and prospects of the morphology of non-fullerene acceptor based high-efficiency organic solar cells. Energy Environ. Sci. 2021, 14, 4341–4357. [Google Scholar] [CrossRef]
  3. Yao, Z.-F.; Wang, J.-Y.; Pei, J. High-performance polymer field-effect transistors: From the perspective of multi-level microstructures. Chem. Sci. 2021, 12, 1193–1205. [Google Scholar] [CrossRef] [PubMed]
  4. Yang, Y.; Hong, Y.; Wang, X. Utilizing the Diffusion of Fluorinated Polymers to Modify the Semiconductor/Dielectric Interface in Solution-Processed Conjugated Polymer Field-Effect Transistors. ACS Appl. Mater. Interfaces 2021, 13, 8682–8691. [Google Scholar] [CrossRef] [PubMed]
  5. Han, J.M.; Yoon, S.E.; Jung, K.H.; Bae, O.; Kim, D.; Kim, U.; Seo, H.; Kim, F.S.; Kim, K.C.; Kim, J.H.; et al. Dopant-dependent thermoelectric performance of indoloindole-selenophene based conjugated polymer. Chem. Eng. J. 2022, 431, 133779. [Google Scholar] [CrossRef]
  6. Fontana, M.T.; Stanfield, D.A.; Scholes, D.T.; Winchell, K.J.; Tolbert, S.H.; Schwartz, B.J. Evaporation vs. Solution Sequential Doping of Conjugated Polymers: F4TCNQ Doping of Micrometer-Thick P3HT Films for Thermoelectrics. J. Phys. Chem. C 2019, 123, 22711–22724. [Google Scholar] [CrossRef]
  7. Jeon, S.; Lee, J.H.; Park, J.I.; Jo, B.; Whang, D.R.; Ahn, T.K.; Park, H.J.; Kim, S.D.; Lee, W.H.; Kim, B.-G. Manipulation of Chain Conformation for Optimum Charge-Transport Pathways in Conjugated Polymers. ACS Appl. Mater. Interfaces 2017, 9, 22757–22763. [Google Scholar] [CrossRef]
  8. Fratini, S.; Nikolka, M.; Salleo, A.; Schweicher, G.; Sirringhaus, H. Charge transport in high-mobility conjugated polymers and molecular semiconductors. Nat. Mater. 2020, 19, 491–502. [Google Scholar] [CrossRef] [Green Version]
  9. Park, K.S.; Kwok, J.J.; Dilmurat, R.; Qu, G.; Kafle, P.; Luo, X.; Jung, S.-H.; Oliver, Y.; Lee, J.-K.; Mei, J.; et al. Tuning conformation, assembly, and charge transport properties of conjugated polymers by printing flow. Sci. Adv. 2019, 5, eaaw7757. [Google Scholar] [CrossRef] [Green Version]
  10. Méndez, H.; Heimel, G.; Opitz, A.; Sauer, K.; Barkowski, P.; Oehzelt, M.; Soeda, J.; Okamoto, T.; Takeya, J.; Arlin, J.-B.; et al. Doping of Organic Semiconductors: Impact of Dopant Strength and Electronic Coupling. Angew. Chem. Int. Ed. 2013, 125, 7905–7909. [Google Scholar] [CrossRef]
  11. Yoon, S.E.; Park, J.; Kwon, J.E.; Lee, S.Y.; Han, J.M.; Go, C.Y.; Choi, S.; Kim, K.C.; Seo, H.; Kim, J.H.; et al. Improvement of Electrical Conductivity in Conjugated Polymers through Cascade Doping with Small-Molecular Dopants. Adv. Mater. 2020, 32, 2005129. [Google Scholar] [CrossRef]
  12. Park, J.; Yoon, S.E.; Lee, J.; Whang, D.R.; Lee, S.Y.; Shin, S.J.; Han, J.M.; Seo, H.; Park, H.J.; Kim, J.H.; et al. Unraveling doping capability of conjugated polymers for strategic manipulation of electric dipole layer toward efficient charge collection in perovskite solar cells. Adv. Funct. Mater. 2020, 30, 2001560. [Google Scholar] [CrossRef]
  13. Yoon, S.E.; Kang, Y.; Jeon, G.G.; Jeon, D.; Lee, S.Y.; Ko, S.-J.; Kim, T.; Seo, H.; Kim, B.-G.; Kim, J.H. Exploring wholly doped conjugated polymer films based on hybrid doping: Strategic approach for optimizing electrical conductivity and related thermoelectric properties. Adv. Funct. Mater. 2020, 30, 2004598. [Google Scholar] [CrossRef]
  14. Yoon, S.E.; Kang, Y.; Noh, S.Y.; Park, J.; Lee, S.Y.; Park, J.; Lee, D.W.; Whang, D.R.; Kim, T.; Kim, G.-H.; et al. High efficiency doping of conjugated polymer for investigation of intercorrelation of thermoelectric effects with electrical and morphological properties. ACS Appl. Mater. Interfaces 2020, 12, 1151–1158. [Google Scholar] [CrossRef]
  15. Cheon, H.J.; Li, X.; Jeong, Y.J.; Sung, M.J.; Li, Z.; Jeon, I.; Tang, X.; Girma, H.G.; Kong, H.; Kwon, S.-K.; et al. A novel design of donor–acceptor polymer semiconductors for printed electronics: Application to transistors and gas sensors. J. Mater. Chem. C 2020, 8, 8410–8419. [Google Scholar] [CrossRef]
  16. Cheon, H.J.; Shin, S.Y.; Tran, V.V.; Park, B.; Yoon, H.; Chang, M. Preparation of conjugated polymer/reduced graphene oxide nanocomposites for high-performance volatile organic compound sensors. Chem. Eng. J. 2021, 425, 131424. [Google Scholar] [CrossRef]
  17. Kwon, E.H.; An, H.; Park, M.B.; Kim, M.; Park, Y.D. Conjugated polymer–zeolite hybrids for robust gas sensors: Effect of zeolite surface area on NO2 sensing ability. Chem. Eng. J. 2021, 420, 129588. [Google Scholar] [CrossRef]
  18. Kang, Y.; Kwak, D.H.; Kwon, J.E.; Kim, B.-G.; Lee, H. NO2-Affinitive Conjugated Polymer for Selective Sub-Parts-Per-Billion NO2 Detection in a Field-Effect Transistor Sensor. ACS Appl. Mater. Interfaces 2021, 13, 31910–31918. [Google Scholar] [CrossRef]
  19. Lin, Y.-C.; Lu, Y.-J.; Tsao, C.-S.; Saeki, A.; Li, J.-X.; Chen, C.-H.; Wang, H.-C.; Chen, H.-C.; Meng, D.; Wu, K.-H.; et al. Enhancing photovoltaic performance by tuning the domain sizes of a small-molecule acceptor by side-chain-engineered polymer donors. J. Mater. Chem. A 2019, 7, 3072. [Google Scholar] [CrossRef]
  20. Otep, S.; Lin, Y.-C.; Matsumoto, H.; Mori, T.; Wei, K.-H.; Michinobu, T. Diketopyrrolopyrrole–thiophene–methoxythiophene based random copolymers for organic field effect transistor applications. Org. Electron. 2020, 87, 105986. [Google Scholar] [CrossRef]
  21. Jia, S.; Qi, S.; Xing, Z.; Li, S.; Wang, Q.; Chen, Z. Effects of Different Lengths of Oligo (Ethylene Glycol) Side Chains on the Electrochromic and Photovoltaic Properties of Benzothiadiazole-Based Donor-Acceptor Conjugated Polymers. Molecules 2023, 28, 2056. [Google Scholar] [CrossRef]
  22. Lin, K.; Wu, C.; Zhang, G.; Wu, Z.; Tang, S.; Lin, Y.; Li, X.; Jiang, Y.; Lin, H.; Wang, Y.; et al. Toward High-Performance Electrochromic Conjugated Polymers: Influence of Local Chemical Environment and Side-Chain Engineering. Molecules 2022, 27, 8424. [Google Scholar] [CrossRef] [PubMed]
  23. Shi, Y.; Li, J.; Sun, H.; Li, Y.; Wang, Y.; Wu, Z.; Jeong, S.Y.; Woo, H.Y.; Fabiano, S.; Guo, X. Thiazole Imide-Based All-Acceptor Homopolymer with Branched Ethylene Glycol Side Chains for Organic Thermoelectrics. Angew. Chem. Int. Ed. 2022, 61, e202214192. [Google Scholar] [CrossRef] [PubMed]
  24. Saber, A.F.; Elewa, A.M.; Chou, H.-H.; EL-Mahdy, A.F.M. Donor to Acceptor Charge Transfer in Carbazole-based Conjugated Microporous Polymers for Enhanced Visible Light-Driven Photocatalytic Water Splitting. ChemCatChem 2023, 15, e202201287. [Google Scholar] [CrossRef]
  25. Barman, S.; Singh, A.; Rahimi, F.A.; Maji, T.K. Metal-Free Catalysis: A Redox-Active Donor-Acceptor Conjugated Microporous Polymer for Selective Visible-Light-Driven CO2 Reduction to CH4. J. Am. Chem. Soc. 2021, 143, 16284. [Google Scholar] [CrossRef] [PubMed]
  26. Saber, A.F.; Elewa, A.M.; Chou, H.-H.; EL-Mahdy, A.F.M. Donor-acceptor carbazole-based conjugated microporous polymers as photocatalysts for visible-light-driven H2 and O2 evolution from water splitting. Appl. Catal. B Environ. 2022, 316, 121624. [Google Scholar] [CrossRef]
  27. Chang, C.-L.; Elewa, A.M.; Wang, J.H.; Chou, H.-H.; EL-Mahdy, A.F.M. Donor-acceptor conjugated microporous polymers based on Thiazolo[5,4-d]thiazole building block for high-performance visible-light-induced H2 production. Microporous Mesoporous Mater. 2022, 345, 112258. [Google Scholar] [CrossRef]
  28. Xie, P.; Han, C.; Xiang, S.; Jin, S.; Ge, M.; Zhang, C.; Jiang, J.-X. Toward high-performance dibenzo[g,p]chrysene-based conjugated polymer photocatalysts for photocatalytic hydrogen production through donor-acceptor-acceptor structure design. Chem. Eng. J. 2023, 459, 141553. [Google Scholar] [CrossRef]
  29. Gao, L.; Liu, C.; Peng, Y.; Deng, J.; Hou, S.; Cheng, Y.; Huang, W.; Yu, J. Ultrasensitive flexible NO2 gas sensors via multilayer porous polymer film. Sens. Actuators B Chem. 2022, 368, 132113. [Google Scholar] [CrossRef]
  30. Luo, H.; Kaneti, Y.V.; Ai, Y.; Wu, Y.; Wei, F.; Fu, J.; Cheng, J.; Jing, C.; Yuliarto, B.; Eguchi, M.; et al. Nanoarchitectured Porous Conducting Polymers: From Controlled Synthesis to Advanced Applications. Adv. Mater. 2021, 33, 2007318. [Google Scholar] [CrossRef]
  31. Park, H.; Kim, D.-H.; Ma, B.S.; Shin, E.; Kim, Y.; Kim, T.-S.; Kim, F.S.; Kim, I.-D.; Kim, B.J. High-Performance, Flexible NO2 Chemiresistors Achieved by Design of Imine-Incorporated n-Type Conjugated Polymers. Adv. Sci. 2022, 9, 2200270. [Google Scholar] [CrossRef]
  32. Chen, Z.; Yu, C.; Bai, W.; Ye, W.; Wang, J.; Wei, J.; Wang, Y.; He, J.; Lu, J. Surface functionalization of ion-in-conjugation polymer sensors for humidity-independent gas detection at room temperature. Sens. Actuators B Chem. 2022, 372, 132654. [Google Scholar] [CrossRef]
  33. Ngai, J.H.L.; Gao, X.; Kumar, P.; Polena, J.; Li, Y. A Highly Stable Diketopyrrolopyrrole (DPP) Polymer for Chemiresistive Sensors. Adv. Electron. Mater. 2021, 7, 2000935. [Google Scholar] [CrossRef]
  34. Chae, H.; Han, J.M.; Ahn, Y.; Kwon, J.E.; Lee, W.H.; Kim, B.-G. NO2-Affinitive Amorphous Conjugated Polymer for Field-Effect Transistor Sensor toward Improved NO2 Detection Capability. Adv. Mater. Technol. 2021, 6, 2100580. [Google Scholar] [CrossRef]
  35. Yang, Y.; Liu, Z.; Zhang, G.; Zhang, X.; Zhang, D. The Effects of Side Chains on the Charge Mobilities and Functionalities of Semiconducting Conjugated Polymers beyond Solubilities. Adv. Mater. 2019, 31, 1903104. [Google Scholar] [CrossRef] [Green Version]
  36. Lu, C.-F.; Liao, S.-F.; Chen, I.-F.; Chen, C.-T.; Chao, C.-Y.; Su, W.-F. Detecting Minute Chemical Vapors via Chemical Interactions between Analyte and Fluorinated Thiophene-Isoindigo Conjugated Polymer Transistor. ACS Appl. Electron. Mater. 2019, 1, 1873–1880. [Google Scholar] [CrossRef]
  37. Yu, C.; Lin, H.-Z.; Zhou, J.; Cheng, X.-F.; He, J.-H.; Li, H.; Xu, Q.-F.; Li, N.-J.; Chen, D.-Y.; Lu, J.-M. An ion-inconjugation polymer enables the detection of NO2 with parts-per-trillion sensitivity and ultrahigh selectivity. J. Mater. Chem. A 2020, 8, 1052–1058. [Google Scholar] [CrossRef]
  38. Perzon, E.; Zhang, F.; Andersson, M.; Mammo, W.; Inganäs, O.; Andersson, M.R. A Conjugated Polymer for Near Infrared Optoelectronic Applications. Adv. Mater. 2007, 19, 3308–3311. [Google Scholar] [CrossRef]
  39. Zhao, X.; Cai, H.; Deng, Y.; Jiang, Y.; Wang, Z.; Shi, Y.; Han, Y.; Geng, Y. Low-Band gap Conjugated Polymers with Strong Absorption in the Second Near-Infrared Region Based on Diketopyrrolopyrrole-Containing Quinoidal Units. Macromolecules 2021, 54, 3498–3506. [Google Scholar] [CrossRef]
  40. Chen, H.; Cai, G.; Guo, A.; Zhao, Z.; Kuang, J.; Zheng, L.; Zhao, L.; Chen, J.; Guo, Y.; Liu, Y. Low Band Gap Donor–Acceptor Conjugated Polymers with Indanone-Condensed Thiadiazolo[3,4-g]quinoxaline Acceptors. Macromolecules 2019, 52, 6149–6159. [Google Scholar] [CrossRef]
  41. Kim, B.-G.; Zhen, C.-G.; Jeong, E.J.; Kieffer, J.; Kim, J. Organic dye design tools for efficient photocurrent generation in dye-sensitized solar cells: Exciton binding energy and electron Acceptors. Adv. Funct. Mater. 2012, 22, 1606–1612. [Google Scholar] [CrossRef] [Green Version]
  42. Kang, Y.; Eun, H.J.; Kye, H.; Kim, D.; Heo, J.; Kim, J.H.; Kim, B.-G. Side-chain engineering of conjugated polymers toward highly efficient near-infrared organic photo-detectors via morphology and dark current management. J. Mater. Chem. C 2020, 8, 7765. [Google Scholar] [CrossRef]
  43. Hou, Y.; Long, G.; Sui, D.; Cai, Y.; Wan, X.; Yu, A.; Chen, Y. Different donor–acceptor structures of dithiafulvalene-fused semiconducting polymers with different band gaps. Chem. Commun. 2011, 47, 10401. [Google Scholar] [CrossRef] [PubMed]
  44. Jiao, Z.; Jiang, T.; Zhou, Z.; Qin, C.; Long, J.; Liu, Y.; Jiang, Y. Identification of a bridge-specific intramolecular exciton dissociation pathway in donor–π–acceptor alternating conjugated polymers. Nanoscale Res. Lett. 2021, 16, 51. [Google Scholar] [CrossRef]
  45. Ahmed, L.R.; Gilmanova, L.; Pan, C.-T.; Kaskel, S.; EL-Mahdy, A.F.M. Hollow Spherical Covalent Organic Frameworks from Nonplanar or Planar Monomers for the Fluorescence Detection of Telomere DNA: Role of the 2-(2-Azidoethoxy)ethoxy Group. ACS Appl. Polym. Mater. 2022, 4, 9132. [Google Scholar] [CrossRef]
  46. Lee, S.W.; Hussain, W.; Lee, J.; Whang, D.R.; Jeong, W.H.; Choi, H.; Chang, D.W. Effect of Chlorine Substituents on Photovoltaic Properties of Monocyanated Quinoxaline-Based Polymers. ACS Appl. Mater. Interfaces 2023, 15, 5547–5555. [Google Scholar] [CrossRef]
  47. Sagita, C.P.; Yoon, J.W.; Keong, M.; Lee, S.W.; Bae, H.; Whang, D.R.; Choi, H.; Ko, S.-J.; Chang, D.W. Improved Photovoltaic Performance of Quinxoalne-Based Polymers by Systematic Modulation of Electron-Withdrawing Substituents. J. Mater. Chem. C 2022, 10, 10338–10346. [Google Scholar] [CrossRef]
  48. Schwarze, M.; Gaul, C.; Scholz, R.; Bussolotti, F.; Hofacker, A.; Schellhammer, K.S.; Nell, B.; Naab, B.D.; Bao, Z.; Spoltore, D.; et al. Molecular parameters responsible for thermally activated transport in doped organic semiconductors. Nat. Mater. 2019, 18, 242–248. [Google Scholar] [CrossRef]
  49. Deng, J.; Zheng, L.; Ding, C.; Guo, Y.; Xie, Y.; Wang, J.; Ke, Y.; Li, M.; Li, L.; Janssen, R.A.J. Determinant Role of Solution-State Supramolecular Assembly in Molecular Orientation of Conjugated Polymer Films. Adv. Funct. Mater. 2023, 33, 2209195. [Google Scholar] [CrossRef]
  50. Moro, S.; Siemons, N.; Drury, O.; Warr, D.A.; Moriarty, T.A.; Perdigão, L.M.A.; Pearce, D.; Moser, M.; Hallani, R.K.; Parker, J.; et al. The Effect of Glycol Side Chains on the Assembly and Microstructure of Conjugated Polymers. ACS Nano 2022, 16, 21303–21314. [Google Scholar] [CrossRef]
  51. Mahmood, J.; Park, J.; Shin, D.; Choi, H.-J.; Seo, J.-M.; Yoo, J.-W.; Baek, J.-B. Organic Ferromagnetism: Trapping Spins in the Glassy State of an Organic Network Structure. Chem 2018, 4, 2357–2369. [Google Scholar] [CrossRef] [Green Version]
  52. Chen, Z.X.; Li, Y.; Huang, F. Persistent and Stable Organic Radicals: Design, Synthesis, and Applications. Chem 2021, 7, 288–332. [Google Scholar] [CrossRef]
  53. Sharma, V.; Puthumana, U.; Karak, P.; Koner, A.L. Visible-Light-Triggered Generation of Ultrastable Radical Anion from Nitro-substituted Perylenediimides. J. Org. Chem. 2018, 83, 11458–11462. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) Chemical structures; (b) absorption tendencies, and (c) electronic energy levels of the obtained CPs.
Figure 1. (a) Chemical structures; (b) absorption tendencies, and (c) electronic energy levels of the obtained CPs.
Materials 16 02877 g001
Figure 2. FET curves of (a) PTQ-T and (b) PTQ-TEG.
Figure 2. FET curves of (a) PTQ-T and (b) PTQ-TEG.
Materials 16 02877 g002
Figure 3. Surface morphologies and chain assembly tendencies characterized using AFM and 2D-GIXRD for (a,c) PTQ-T and (b,d) PTQ-TEG.
Figure 3. Surface morphologies and chain assembly tendencies characterized using AFM and 2D-GIXRD for (a,c) PTQ-T and (b,d) PTQ-TEG.
Materials 16 02877 g003
Figure 4. (a) Atomic binding-energy distributions of PTQ-T and PTQ-TEG and (b) comparison of intercalated O atoms in PTQ-TEG. “*” indicates the presence of O atoms having a higher binding energy compared to the pristine O atoms.
Figure 4. (a) Atomic binding-energy distributions of PTQ-T and PTQ-TEG and (b) comparison of intercalated O atoms in PTQ-TEG. “*” indicates the presence of O atoms having a higher binding energy compared to the pristine O atoms.
Materials 16 02877 g004
Figure 5. ESR signals and corresponding g-factors of (a) PTQ-T and (b) PTQ-TEG and comparison of the spin density of each ESR component. (c) indicates the spin density corresponding to each g-factor in ESR signals of both CPs.
Figure 5. ESR signals and corresponding g-factors of (a) PTQ-T and (b) PTQ-TEG and comparison of the spin density of each ESR component. (c) indicates the spin density corresponding to each g-factor in ESR signals of both CPs.
Materials 16 02877 g005
Figure 6. NO2 detection characteristics of (a) PTQ-T and (b) PTQ-TEG and comparison of their sensitivity and recovery tendency. (c) indicates the NO2 detectivity and recovery tendency of each CP.
Figure 6. NO2 detection characteristics of (a) PTQ-T and (b) PTQ-TEG and comparison of their sensitivity and recovery tendency. (c) indicates the NO2 detectivity and recovery tendency of each CP.
Materials 16 02877 g006
Figure 7. (a) Detection selectivity and (b) the calculated theoretical LOD of a resistive sensor adopting PTQ-TEG.
Figure 7. (a) Detection selectivity and (b) the calculated theoretical LOD of a resistive sensor adopting PTQ-TEG.
Materials 16 02877 g007
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ahn, Y.; Hwang, S.; Kye, H.; Kim, M.S.; Lee, W.H.; Kim, B.-G. Side-Chain-Assisted Transition of Conjugated Polymers from a Semiconductor to Conductor and Comparison of Their NO2 Sensing Characteristics. Materials 2023, 16, 2877. https://doi.org/10.3390/ma16072877

AMA Style

Ahn Y, Hwang S, Kye H, Kim MS, Lee WH, Kim B-G. Side-Chain-Assisted Transition of Conjugated Polymers from a Semiconductor to Conductor and Comparison of Their NO2 Sensing Characteristics. Materials. 2023; 16(7):2877. https://doi.org/10.3390/ma16072877

Chicago/Turabian Style

Ahn, Yejin, Sooji Hwang, Hyojin Kye, Min Seon Kim, Wi Hyoung Lee, and Bong-Gi Kim. 2023. "Side-Chain-Assisted Transition of Conjugated Polymers from a Semiconductor to Conductor and Comparison of Their NO2 Sensing Characteristics" Materials 16, no. 7: 2877. https://doi.org/10.3390/ma16072877

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop