Next Article in Journal
A Critical Analysis on the Sensitivity Enhancement of Surface Plasmon Resonance Sensors with Graphene
Next Article in Special Issue
Flower-like ZnO Nanostructures Local Surface Morphology and Chemistry
Previous Article in Journal
Green and Efficient Preparation of Tailed Lignin Nanoparticles and UV Shielding Composite Films
Previous Article in Special Issue
Recent Progress on Nanomaterials for NO2 Surface Acoustic Wave Sensors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Microstructural Control of Soluble Acene Crystals for Field-Effect Transistor Gas Sensors

1
Department of Materials Science and Engineering, Seoul National University, Seoul 08826, Korea
2
Department of Organic and Nano System Engineering, Konkuk University, Seoul 05029, Korea
3
Department of Physics, Konkuk University, Seoul 05029, Korea
4
Division of Chemical Engineering, Konkuk University, Seoul 05029, Korea
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(15), 2564; https://doi.org/10.3390/nano12152564
Submission received: 22 June 2022 / Revised: 21 July 2022 / Accepted: 22 July 2022 / Published: 26 July 2022
(This article belongs to the Special Issue Nanomaterials in Gas Sensors)

Abstract

:
Microstructural control during the solution processing of small-molecule semiconductors (namely, soluble acene) is important for enhancing the performance of field-effect transistors (FET) and sensors. This focused review introduces strategies to enhance the gas-sensing properties (sensitivity, recovery, selectivity, and stability) of soluble acene FET sensors by considering their sensing mechanism. Defects, such as grain boundaries and crystal edges, provide diffusion pathways for target gas molecules to reach the semiconductor-dielectric interface, thereby enhancing sensitivity and recovery. Representative studies on grain boundary engineering, patterning, and pore generation in the formation of soluble acene crystals are reviewed. The phase separation and microstructure of soluble acene/polymer blends for enhancing gas-sensing performance are also reviewed. Finally, flexible gas sensors using soluble acenes and soluble acene/polymer blends are introduced, and future research perspectives in this field are suggested.

1. Introduction

Solution-processed small-molecule semiconductors (so-called soluble acenes) have attracted much attention as alternative field-effect transistors (FETs) for flexible switching elements in displays and sensors [1,2,3,4,5,6,7,8]. Control of the crystalline microstructure during the solution processing of soluble acene is critical for its desired function in FETs and sensors [9,10,11,12,13,14,15]. Typically, highly crystalline films are desirable for the effective π–π stacking of conjugated acene moieties and the corresponding lateral transport of charge carriers in FETs. Although the fabrication of uniform and large-area single-crystal films is a target goal for achieving the best performance in FETs, solution processing of soluble acene typically leads to a polycrystalline film with defects inside the film. Accordingly, the fine control of solvent evaporation during the formation of crystals is necessary. Several well-written review articles focusing on the microstructural and morphological control during solvent evaporation in the casted soluble acene solution are available [9,10,11,16,17,18,19,20,21]. Strategies, such as the control of solution composition, printing parameters (for example, spin speed in spin-casting and jetting condition in inkjet printing), and surface energy of the substrate, have been well documented in recent publications for high-performance soluble acene FETs [9,16,17,18,22,23,24].
In addition to the performance of FETs (for example, field-effect mobility and on–off current ratio threshold voltage), sensors require other performance parameters [25,26,27,28,29,30,31]. FET-based gas sensors particularly require a diffusion pathway for the active semiconductor channel for gas molecules to reach the semiconductor-dielectric interface. Note that the semiconducting layer near the semiconductor-dielectric interface is the region where most of the field-effect charge carriers pass through in a given gate bias of FETs. The gas molecules affect these charge carriers at the interface, leading to a current change. The current change is monitored at the constant source-drain bias condition in FETs, and sensitivity, recovery, selectivity, and stability are the required performance parameters in FET-based gas sensors [26,27,32,33,34].
In FET-based gas sensors, the formation of defects (such as grain boundaries, crystal edges, and pores) is key to guaranteeing a diffusion pathway for sensitive gas detection [15,23,26,35,36,37]. However, review articles focusing on the defect engineering of soluble acene crystals with a particular focus on the employed strategies are currently unavailable, although there have been many reports on controlling these defects in soluble acene FET-based gas sensors. The use of soluble acene in the semiconducting layers of FET-based gas sensors has several advantages. First, soluble acene FETs can be used for detecting environmentally harmful gases (such as nitrogen dioxide and ammonia) and healthcare-related exhaled gases (such as formaldehyde and acetone) with a detection limit as low as ppb. Second, selectivity to the target gas is feasible when using a specific soluble acene, which can be obtained by the synthesis of a new soluble acene. Here, the manipulation of specific interactions between the soluble acene and target gas molecules is required. Third, the printing technique can be utilized to fabricate large-area/flexible gas sensors on plastic substrates as it is better than a vacuum-based evaporation tool for reducing costs and increasing manufacturing speed [23,38,39,40,41,42]. Typically, a polymer binder is added to a soluble acene solution to fabricate printing-based, flexible gas sensors. The polymer can increase the solution viscosity while reducing the dewetting of soluble acene crystals. Here, inducing vertical phase separation in soluble acene/polymer blends is also necessary for fabricating FETs. Readers can read recent review papers focusing on obtaining high field-effect mobility in FETs based on soluble acene/polymer blends [17,18,43,44,45,46,47,48,49]. Instead of introducing strategies for vertical phase separation, this review focuses on introducing key concepts in the microstructural control of soluble acene/polymer blends for high-performance gas sensors.
This focused review paper introduces several representative works on the microstructural control in soluble acene and soluble acene/polymer blends for high-performance FET-based gas sensors. Section 2 introduces the working principle of the soluble acene FET-based gas sensor. Section 3 reviews the control of the soluble acene microstructure to enhance the gas-sensing performance. Section 4 reviews the microstructural control of soluble acene/polymer blends for improving the gas-sensing performance. Section 5 introduces representative studies on flexible soluble acene gas sensors. Finally, the conclusions and future research perspectives are presented.

2. Soluble Acene FET-based Gas Sensors: Working Principle

This review uses the word “soluble acene” to mention solution-processed small molecular semiconductors. Typically, the semiconducting properties of organic molecules are induced by extending conjugation in fused acenes or hetero-acenes [50,51]. However, fused acene or hetero-acene (e.g., pentacene) are not soluble in common solvents, although some studies solubilized pentacene with harsh chemical treatments [52,53,54]. The addition of solubilizing groups to the six and 13 positions in pentacene can increase its solubility while changing the packing motif from herringbone stacking to co-facial brick wall stacking (Figure 1a) [55]. Figure 1 shows the chemical structures of representative soluble acenes, (e.g., 6,13-bis(triisopropylsilylethynyl)pentacene (TIPS-pentacene)). In addition to the commonly used spin-coating or inkjet printing, solution-shearing-assisted bar coating leads to a high-performance TIPS-pentacene FET with field-effect mobility exceeding 4.6 cm2/Vs [56,57,58]. The addition of triethylsilylethynyl groups to the anthradithiophene backbone is also an effective method for increasing both solubility and solid-state ordering (Figure 1b) [59]. Interestingly, a 5,11-bis(triethylsilylethynyl)anthradithiophene (TES-ADT) film was crystallized by solvent vapor annealing, and millimeter-sized TES-ADT spherulites were easily fabricated using this method [60,61,62]. The further addition of fluorine atoms to TES-ADT led to exceptionally small π–π stacking due to fluorine-fluorine interactions (Figure 1c) [63]. Accordingly, 2,8-difluoro-5,11-bis(triethylsilylethynyl)anthradithiophene (diF-TES-ADT) exhibited an ultrafast crystallization capability during solution processing. High mobility exceeding 5 cm2/Vs has been reported in FETs based on diF-TES-ADT/polymer blends [6,64,65,66]. Instead of bulky silylethynyl groups, alkyl chains (such as hexyl and octyl) can also be attached to the edges of the acene backbone to enable solution processability (Figure 1d,e). 2,7-Dihexyl-dithieno[2,3-d;2′,3′-d′]benzo[1,2-b;4,5-b′]dithiophene (C6-DTBDT) and 2,7-dioctylbenzothieno[2,3-b]benzothiophene (C8-BTBT) are examples that show excellent field-effect mobilities in the application of FETs [49,67,68,69].
Figure 2 illustrates the gas-sensing mechanism when these soluble acenes are used as the active layer in FET gas sensors. When a gate bias is applied in FETs, the polarization of the insulator leads to an accumulation of charge carriers at the semiconducting layer near the semiconductor-insulator interface. Because all the soluble acenes in Figure 1 are operated in the p-type mode, a negative gate bias is applied, and hole carriers are generally accumulated at the semiconducting layer. When gas molecules penetrate the semiconducting layer near the semiconductor-insulator interface, hole carriers increase (or decrease) depending on the type of gas molecules [25,70,71,72]. In the case of oxidizing gases, such as NO2, the electron-withdrawing character of the gas induces hole accumulation (that is, an increase in carrier density). Thus, the source-drain current increases upon exposure to the oxidizing gas. In contrast, the electron-donating character of reducing gases, such as NH3, induces hole depletion (that is, a decrease in carrier density), leading to a decrease in the source-drain current. The change in the field-effect mobility can also be monitored in FET gas sensors. If the semiconductor is not a perfect single crystal (this is the typical case in soluble acene films manufactured using solution processing), the trap filling behavior in the oxidizing gas leads to an increase in the field-effect mobility. In contrast, it decreases upon exposure to a reducing gas. However, the change in mobility is also affected by the scattering of charge carriers by gas molecules. In that case, the field-effect mobility decreases in the presence of reducing/oxidizing gases. The charge–dipole interaction between the semiconductor and gas molecules could also change the mobility of gas sensors. Accordingly, changes in the electrical performance of FETs depend on the specific soluble acene type and target gas molecules.

3. Control of Soluble Acene Microstructure for Gas Sensors

Soluble acene microstructure could be manipulated by the careful control of solvent evaporation during solution processing (e.g., drop-casting, spin-coating, inkjet printing, bar coating). Especially, the use of solvent with a high boiling point is preferable for inducing slow evaporation of solvent, thereby increasing crystallinity with structural perfectness. However, soluble acene with low film-forming capability triggers thin-film dewetting during solution processing. Thus, solvent with appropriate vapor pressure and viscosity is used in solution processing such as spin-coating and bar coating. Because crystalline microstructure (e.g., crystallinity, grain boundary, misorientation) of soluble acene is highly dependent on the processing condition, fine control solvent evaporation is indispensable for the proper functions in FETs and sensors. Because several well-written review articles are available in the morphological and structural control of soluble acene [9,10,11,16,17,18,19,20,21], control of the microstructure focusing on gas sensing performance is introduced in this section. In addition, the gas sensing mechanism in the representative works is reviewed by considering the specific interaction between soluble acene and gas molecules.
Soluble acene is composed of an aromatic conjugated core (e.g., fused acene, hetero-acene) and a solubilizing group (e.g., aliphatic alkyl chain) (Figure 1). In soluble acene FET gas sensors, it is important to examine the specific interaction between soluble acene and gas molecules from the prototypical soluble acene, TIPS-pentacene. Because TIPS-pentacene is oriented with silylethynyl groups on the substrate surface (Figure 1a) [55], the adsorption of gas molecules on TIPS-pentacene may be weaker than pentacene without insulating silylethynyl groups. The pentacene backbone cannot interact directly with the gas molecules in TIPS-pentacene. However, a recent report by Wang et al., revealed that TIPS-pentacene FETs could detect ppm levels of NO2 better than pentacene FETs [73]. Figure 3a shows the device structure of TIPS-pentacene FET sensors. Para-sexiphenyl (p-6P) was used as the buffer layer to obtain a terrace-like TIPS-pentacene film. Although TIPS-pentacene was thermally evaporated on the p-6P film, the sensing modality of TIPS-pentacene could be measured using the gas testing setup (Figure 3b). Dynamic tests with periodic exposure to NO2 showed that the TIPS-pentacene sensor exhibited excellent sensitivity (>1000%/ppm), recovery (>90%), and reproducibility (Figure 3c). When cross-sensitivity was measured with different gases (Figure 3d), the TIPS-pentacene sensor exhibited selective detection of NO2. The responsivities and sensitivities to SO2, wet air, and NH3 were low. Interestingly, the TIPS-pentacene FETs exhibited better sensing performance than the pentacene FETs. The combined effect of low intrinsic conductivity and efficient charge transport ability in TIPS-pentacene films leads to a high gas-on/gas-off conductivity, resulting in excellent sensitivity toward NO2. A similar performance enhancement of the TIPS-pentacene sensor toward NH3 compared to pentacene sensor has also been reported [74].
Because the diffusion of gas molecules into the semiconducting layer near the semiconductor-dielectric interface is important, defects, grain boundaries, and molecular ordering in soluble acene films affect the gas sensing performance of soluble acene FET-based gas sensors. Thus, it is important to control the soluble acene microstructure during solution processing. Shao et al., examined the effect of crystallinity and grain boundaries on the NO2 sensing performance of TIPS-pentacene FETs using different types of processing solvents (Figure 4a) [75]. When four different types of solvents (namely, o-xylene, toluene, chlorobenzene (CB), and 1,2-dichlorobenzene (1,2-DCB)) were used, the spin-cast films exhibited different morphologies and microstructures. Figure 4b shows the NO2 sensing performance for the different processing solvents. Upon exposure to NO2, the hole-carrier density and field-effect mobility increased. The responses were ordered as follows: o-xylene > toluene > CB > 1,2-DCB. Atomic force microscopy images and X-ray diffraction results indicated that TIPS-pentacene from o-xylene exhibited a highly crystalline structure with a high density of grain boundaries (Figure 4c). In contrast, the CB sample exhibited a lower grain boundary density. Thus, TIPS-pentacene FETs fabricated from CB exhibited low sensitivity, regardless of their high field-effect mobility. The 1,2-DCB sample had a loosely connected microstructure with the lowest crystallinity, degrading the field-effect mobility and sensitivity. Using the TIPS-pentacene film fabricated from o-xylene, the limit of detection of NO2 could be lowered to 1.93 ppb. Seo et al., found that the gas-sensing performance was enhanced by increasing the grain boundary density in TES-ADT spherulites [76]. The grain boundary density was regulated by changing the mixing time of the TES-ADT solution [77,78]. Subsequent solvent vapor annealing of the spin-cast TES-ADT films led to the formation of TES-ADT spherulites with different grain boundary densities (Figure 4d). A mixing time of 5 min resulted in a large grain size and a corresponding high field-effect mobility of 0.3 cm2/Vs. However, a mixing time of 12 h led to a smaller grain size and moderate field-effect mobility of 0.16 cm2/Vs. The mixing-induced self-aggregation behavior of TES-ADT molecules triggers aggregates in the spin-cast TES-ADT film, and subsequent solvent vapor annealing leads to a high nucleation density and corresponding small spherulites with a high grain boundary density. The response rate and sensitivity of the TES-ADT film with a high grain boundary density (sample of 12 h) exhibited significantly better performance, notwithstanding the low field-effect mobility (Figure 4e). Similarly, the increase in grain boundaries by vertical annealing or solvent vapor annealing could also be an effective method for increasing the sensitivity to NO2 [79,80]. From these results, grain boundaries provide a pathway for the target gas molecules to reach the semiconductor-insulator interface.
The diffusion pathway for the gas molecules can also be enlarged by fabricating soluble acene stripes. Gas molecules easily reach the semiconductor-dielectric interface through these crystal edges. Li et al., demonstrated that ultrathin C6-DTBDT microstripes could be a good sensing platform for adsorption, diffusion, interaction, and desorption activities [81]. They fabricated C6-DTBDT microstripes using dip coating (Figure 5a), and the sensor based on C6-DTBDT microstrips exhibited an excellent response toward NH3. The electrostatic interaction between the electron-deficient thiophene unit and electron-rich NH3 led to the adsorption of NH3, thereby inducing traps and de-doping inside the channel (Figure 5b). The combined effect of de-doping and dipole-charge interactions led to an abrupt decrease in the channel current. Here, the form and microstripes were better than the film because the interaction could be facilitated by the efficient pathway between the microstripes (Figure 5c). Although the sensor responded to other analytes, it exhibited the best performance (for example, high sensitivity and low response/recovery times) for NH3. C6-DTBDT microstripes across Au source/drain electrodes were fabricated using the wetting-dewetting concept with evaporated Au electrodes and octadecyltrichlorosilane-treated SiO2 (Figure 5d) [82]. It was possible to fabricate C6-DTBDT microstripes across the channel region, and the sensor exhibited an excellent response to NH3.
The patterning (such as rectangle or line) of soluble acene is important for guaranteeing the diffusion pathway for target gas molecules. Kwak et al., systematically examined the relationship between the dimensions of rectangular/line patterns and the gas-sensing performance [83]. They used a rectangular/line-patterned polydimethylsiloxane (PDMS) mold to fabricate TES-ADT crystal arrays (Figure 6a, top). The contact region in the TES-ADT film could be etched by the solvent-soaked PDMS mold, whereas solvent vapor-assisted crystallization was facilitated in the non-contact region [15,84,85]. Changing the pattern dimensions in the PDMS mold, it was possible to fabricate TES-ADT crystal arrays with different widths (Figure 6a, bottom). The gas-sensing performance was heavily dependent on the pattern type. With the narrowest line width, Type C exhibited the highest performance (such as response rate, recovery rate, and sensitivity) for NO2 (Figure 6b). Because the total pathway length for gas diffusion can be estimated from the sum of the grain boundary and edge lengths, the sensitivity is proportional to the total pathway for gas diffusion (Figure 6c). From these results, it was concluded that the ultrathin microstripe pattern provides an efficient gas-sensing platform for reversible gas adsorption and desorption (Figure 6d).

4. Microstructural Control of Soluble Acene/Polymer Blends for Gas Sensors

Because soluble acene has a low film-forming capability, polymer binders are typically added to prepare soluble acene solutions, and spin-casting, dip-coating, inkjet printing, and bar coating are applied to fabricate soluble acene/polymer blend films [17,39,43,44,45]. In addition to the insulating binder polymers, poly(triaryl amine) (PTAA) with the highest occupied molecular orbital level similar to that of soluble acene is used as the counterpart polymer in the preparation of the blend solution [6,64]. However, there have not been any reports on soluble acene/PTAA blends in gas sensor applications; therefore, papers on soluble acene/PTAA blends are excluded in this review which focuses on FET gas sensors. As the charge carrier path is substantially lateral to the substrate surface, the formation of a vertically phase-separated structure is indispensable. The authors reviewed well-written review articles that introduce recent papers dealing with phase separation in soluble acene/insulating polymer blends [17]. For gas sensor applications, the active layer should be exposed to ambient air; therefore, the bottom-gate structure is preferable to the top-gate structure. Figure 7 shows a schematic of the phase-separation behavior of TIPS-pentacene/polymer blends during the evaporation of the residual solvent. Insulating polymers (such as poly(methyl methacrylate) (PMMA), poly(alpha-methylstyrene) (PαMS), and polystyrene (PS)) can be used as binder polymers, which increase the solution viscosity and reduce the dewetting of soluble acene. The TIPS-pentacene-top/polymer-bottom structure is spontaneously induced during spin-casting, mainly because of the lower surface energy of TIPS-pentacene compared to its counterpart polymer. Here, the residual solvent at a given spin time determines the phase-separated structure and crystallization behavior of the soluble acene (Figure 7, bottom schematic) [13,86].
Lee et al., examined the phase separation and structural development of soluble acene/polymer blends by changing the spin time of a spin-casting blend solution [13]. 1,2-dichlorobenzene with a high boiling point (that is, low solvent evaporation rate) was used to amplify the effect of the residual solvent (Figure 7). A vertically phase-separated structure consisting of a soluble acene-top and polymer-bottom was induced, regardless of the spin time. In a short spin time of 5 s, the excess residual solvent resulted in the flow-induced growth of needle-like one-dimensional (1D) crystals from the edge to the center position (Figure 8a). In contrast, the optimum residual solvent at a spin time of 50 s resulted in two-dimensional (2D) spherulite crystals. The 2D crystals at 50 s exhibit significantly higher mobility (approximately 1 cm2/Vs) in FET applications than the 1D crystals at 5 s because of the higher crystal perfection and coverage of soluble acene crystals on the phase-separated insulating polymer. The response rate, recovery rate, and sensitivity toward NO2 of the 2D structure were better than those of the 1D structure (Figure 8b) [37]. Although the 1D microstructure provides a microscale route for gas adsorption/desorption, the inverse structure in the 2D crystals has many nanoscale holes (Figure 8b, bottom image). A porous structure with a lower film thickness and nanoscale holes in 2D crystals is preferable for gas diffusion, whereas higher field-effect mobility facilitates fast detection in FET gas sensors. It should be emphasized that the 2D structure was better than the 1D structure in the proposed experimental system. Although other results opposite to this trend are possible, a porous structure with many holes provides an excellent sensing platform for FET-based gas sensors.
Although porous structures have been proposed for FET-based gas sensors, a general route for fabricating porous structures (including soluble acene films) is necessary. Zhang et al., proposed a methodology for fabricating porous semiconducting films with organic semiconductor/PS blends (Figure 9a) [87]. In the so-called breath figure method, water condensation under high humidity conditions (approximately 60%) and subsequent evaporation of water under thermal annealing led to porous films. Figure 9b shows a C8-BTBT/PS film fabricated by the breath figure method. Porous structures were easily fabricated with other blend solutions (such as p-type poly(3-hexylthiophene)/PS, n-type poly[N,N′-bis(2-octyldodecyl)-naphthalene-1,4,5,8-bis(dicarboximide)-2,6-diyl]-alt-5,5′-(2,2′-bithio-phene) (N2200)/PS). Figure 9c compares the gas-sensing properties of dense and porous blend films. Porous films exhibit considerably better sensitivities to NH3 than dense films. This result indicates that a more accessible morphology with nanoscale holes in the active layer is advantageous for guaranteeing a diffusion pathway in FET gas sensors.
Soluble acene/polymer blends are beneficial in terms of long-term device stability (such as environmental and gate-bias stability). In particular, the gate-bias instability of soluble acene FETs can be reduced by incorporating an insulating polymer, which is easily achieved by blending with an insulating polymer [88]. The FET-based gas sensors are also operated under continuous bias conditions. Accordingly, the gate-bias screens the hole carriers in p-type organic semiconductors, leading to on-current decay and a negative shift in the threshold voltage. It is necessary to decrease the gate-bias instability in FET gas sensors using several strategies. Kwak et al., enhanced the device stability of TES-ADT microstripe sensors using TES-ADT/PMMA blends [15]. A patterning method similar to Figure 6a was used, and spin-casting of the TES-ADT/PMMA blend solution and subsequent stamping with a solvent-soaked/line-patterned PDMS mold led to highly crystalline TES-ADT-top/PMMA-bottom microcrystal arrays (Figure 10a). Figure 10b shows polarized optical microscopy images of the pristine TES-ADT and TES-ADT/PMMA microstripes with different blend ratios. Although the pristine TES-ADT microstripes exhibit a clearly patterned image, the blend samples show residue in the etched region. Additionally, the proportion of lateral phase separation increased with an increasing PMMA content. Consequently, the patterned TES-ADT/PMMA blend FETs exhibited lower field-effect mobility and sensitivity than the patterned TES-ADT FETs (Figure 10c). However, the patterned TES-ADT/PMMA blend FETs showed a significantly lower on-current decay under the sensor operating conditions. When the gate-bias instabilities were measured, the blend FETs exhibited lower on-current decays and corresponding higher extracted characteristic times (τ) (Figure 10d), indicating that the charge carriers in the mobile states increased with the ratio of PMMA in the TES-ADT/PMMA blends. Incorporating PMMA by blending with PMMA provides a route for decreasing the gate-bias instability in soluble acene FET gas sensors.

5. Flexible Soluble Acene Gas Sensors

Soluble acenes are promising candidates for use in flexible printed electronics and sensors. Yu et al., developed a flexible TIPS-pentacene FET gas sensor using spray coating [89]. Spray coating of TIPS-pentacene on a PMMA/substrate led to a highly crystalline TIPS-pentacene film for high-performance OFETs (Figure 11a). Exposure to the NH3 target gas resulted in a substantial decrease in the source-drain current in the FETs, mainly due to the electron-donating character of NH3 (Figure 11b). Figure 11c shows a flexible gas sensor fabricated on a plastic substrate using spray coating, and the change in the transfer curve under NH3 exposure is shown in Figure 11d. As the NH3 concentration increased, the turn-on voltage decreased, which is consistent with the decrease in the source-drain current at the given gate and source-drain voltages. Although a flexible gas sensor on a plastic substrate was demonstrated with spray coating, the operating voltage in this study was still high.
Feng et al., developed a low-voltage-driven NH3 gas sensor based on a TIPS-pentacene/PS blend [90]. Figure 12a displays the device structure of the plastic substrate. They used low-k poly(vinyl cinnamate) (PVC) as the gate dielectric and inkjet-printed Ag as the gate and source-drain electrodes. Dropping the TIPS-pentacene/PS blend onto the inclined PVC gate dielectric resulted in the 1D growth of TIPS-pentacene crystals (Figure 12b). Flow-induced 1D growth of TIPS-pentacene on an inclined substrate has previously been reported [91]. In addition to the flexibility shown in Figure 12c, the device exhibited remarkable FET characteristics (such as a low subthreshold slope, low-voltage operation, and operational stability). In particular, the FET exhibited a significant response toward NH3 under ambient conditions with a minimum power consumption of 50 nW. Figure 12d shows the readout circuit of the NH3 FET gas sensor, which successfully changes the current signal to the voltage output signal (Vout) using a load resistor. The high signal at the inlet of the ambient NH3 and stable background at the outlet of the ambient NH3 demonstrated the stable operation of the NH3 sensor under ambient conditions.

6. Conclusions and Future Perspective

This paper reviews soluble acene-based FET gas sensors, focusing on the microstructural control of soluble acene films and soluble acene/polymer blend films. Soluble acene-based FET sensors exhibit different sensing modalities depending on the type of soluble acene; however, in most cases, they can detect NO2 and NH3 with dipolar characteristics. Considering the diffusion-limited adsorption/desorption of target gas molecules, the sensitivity can be enhanced by grain boundary engineering, patterning, and generation of pores in the active layer of FET gas sensors. Soluble acene/insulating polymer blends have also been suggested as sensing layers for FETs. The microstructure and vertical phase separation must be finely tuned to enhance the sensitivity, recovery, and stability of the gas sensors. In particular, the growth characteristics of soluble acene on an insulating polymer determine the FET performance and sensor characteristics. Recent studies have also been conducted on flexible soluble acene gas sensors and circuits.
Notwithstanding the introduced studies, the selectivity of the target gas is not easily attainable for soluble acene FET gas sensors. Because NO2 is the strongest oxidizing gas in the harmful gas libraries, the sensitivity toward NO2 is generally the highest in soluble acene gas sensors. TIPS-pentacene exhibited the highest selectivity for NO2. However, it is not easy to select other gases with low oxidation or reducing strength. It should be emphasized that many types of gases exist under atmospheric conditions, which may interfere with the target gas, interrupting the sensing signal from the target gas. Thus, a selective sensing mechanism that manipulates the soluble acene–gas interaction must be developed. Under atmospheric conditions, water and oxygen molecules always affect sensitivity and long-term stability. Thus, a soluble acene gas sensor that is stable under humid atmospheric conditions should be developed. Although a flexible gas sensor has been demonstrated, a soluble acene-based gas sensor with other form factors (such as being stretchable and rollable) must be developed to enable wearable electronics. For applications in wearable electronics, exhale breath sensors employing soluble acene FETs will be particularly useful for the real-time analysis of human diseases.

Author Contributions

Review suggestion and section organization, W.H.L.; manuscript writing, J.H.L., J.H.C. and W.H.L.; manuscript correction, H.-J.C.; proofreading, W.H.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Korea Institute for Advancement of Technology (KIAT) grant funded by the Korea Government (MOTIE) (P0012770) and grants from the Basic Science Research Program of the National Research Foundation of Korea (NRF) funded by the Ministry of Science and ICT (2019R1A2C1010723).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gasperini, A.; Jeanbourquin, X.A.; Rahmanudin, A.; Yu, X.; Sivula, K. Enhancing the Thermal Stability of Solution-Processed Small-Molecule Semiconductor Thin Films Using a Flexible Linker Approach. Adv. Mater. 2015, 27, 5541–5546. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Paek, S.; Cho, N.; Cho, S.; Lee, J.K.; Ko, J. Planar star-Shaped organic semiconductor with fused triphenylamine core for solution-processed small-molecule organic solar cells and field-Effect transistors. Org. Lett. 2012, 14, 6326–6329. [Google Scholar] [CrossRef] [PubMed]
  3. Haase, K.; Teixeira da Rocha, C.; Hauenstein, C.; Zheng, Y.; Hambsch, M.; Mannsfeld, S.C. High-Mobility, solution-processed organic field-effect transistors from C8-BTBT: Polystyrene blends. Adv. Electron. Mater. 2018, 4, 1800076. [Google Scholar] [CrossRef]
  4. Hwang, D.K.; Dasari, R.R.; Fenoll, M.; Alain-Rizzo, V.; Dindar, A.; Shim, J.W.; Deb, N.; Fuentes-Hernandez, C.; Barlow, S.; Bucknall, D.G. Stable Solution-Processed Molecular n-Channel Organic Field-Effect Transistors. Adv. Mater. 2012, 24, 4445–4450. [Google Scholar] [CrossRef]
  5. Rodriquez, D.; Savagatrup, S.; Valle, E.; Proctor, C.M.; McDowell, C.; Bazan, G.C.; Nguyen, T.-Q.; Lipomi, D.J. Mechanical properties of solution-Processed small-Molecule semiconductor films. ACS Appl. Mater. Interfaces 2016, 8, 11649–11657. [Google Scholar] [CrossRef]
  6. Smith, J.; Zhang, W.; Sougrat, R.; Zhao, K.; Li, R.; Cha, D.; Amassian, A.; Heeney, M.; McCulloch, I.; Anthopoulos, T.D. Solution-Processed small molecule-polymer blend organic thin-Film transistors with hole mobility greater than 5 cm2/Vs. Adv. Mater. 2012, 24, 2441–2446. [Google Scholar] [CrossRef]
  7. Peng, B.; Huang, S.; Zhou, Z.; Chan, P.K.L. Solution-Processed Monolayer Organic Crystals for High-Performance Field-Effect Transistors and Ultrasensitive Gas Sensors. Adv. Funct. Mater. 2017, 27, 1700999. [Google Scholar] [CrossRef]
  8. Zhang, F.; Qu, G.; Mohammadi, E.; Mei, J.; Diao, Y. Solution-Processed nanoporous organic semiconductor thin films: Toward health and environmental monitoring of volatile markers. Adv. Funct. Mater. 2017, 27, 1701117. [Google Scholar] [CrossRef]
  9. Lim, J.A.; Lee, H.S.; Lee, W.H.; Cho, K. Control of the morphology and structural development of solution-Processed functionalized acenes for high-Performance organic transistors. Adv. Funct. Mater. 2009, 19, 1515–1525. [Google Scholar] [CrossRef]
  10. Gundlach, D.J.; Royer, J.E.; Park, S.K.; Subramanian, S.; Jurchescu, O.D.; Hamadani, B.H.; Moad, A.J.; Kline, R.J.; Teague, L.C.; Kirillov, O. Contact-Induced crystallinity for high-Performance soluble acene-Based transistors and circuits. Nat. Mater. 2008, 7, 216–221. [Google Scholar] [CrossRef]
  11. Kline, R.J.; Hudson, S.D.; Zhang, X.; Gundlach, D.J.; Moad, A.J.; Jurchescu, O.D.; Jackson, T.N.; Subramanian, S.; Anthony, J.E.; Toney, M.F. Controlling the microstructure of solution-Processable small molecules in thin-Film transistors through substrate chemistry. Chem. Mater. 2011, 23, 1194–1203. [Google Scholar] [CrossRef]
  12. Chen, J.; Tee, C.K.; Shtein, M.; Anthony, J.; Martin, D.C. Grain-Boundary-Limited charge transport in solution-processed 6, 13 bis (tri-Isopropylsilylethynyl) pentacene thin film transistors. J. Appl. Phys. 2008, 103, 114513. [Google Scholar] [CrossRef]
  13. Lee, J.H.; Choi, H.H.; Park, Y.D.; Anthony, J.E.; Lim, J.A.; Cho, J.; Chung, D.S.; Hwang, J.; Jang, H.W.; Cho, K. 1D versus 2D Growth of Soluble Acene Crystals from Soluble Acene/Polymer Blends Governed by a Residual Solvent Reservoir in a Phase-Separated Polymer Matrix. Adv. Funct. Mater. 2018, 28, 1802875. [Google Scholar] [CrossRef]
  14. Niazi, M.R.; Li, R.; Qiang Li, E.; Kirmani, A.R.; Abdelsamie, M.; Wang, Q.; Pan, W.; Payne, M.M.; Anthony, J.E.; Smilgies, D.-M. Solution-printed organic semiconductor blends exhibiting transport properties on par with single crystals. Nat. Commun. 2015, 6, 8598. [Google Scholar] [CrossRef] [Green Version]
  15. Kwak, D.H.; Choi, H.H.; Anthony, J.E.; Kim, S.; Chae, H.; Hwang, J.; Lee, S.; Park, H.J.; Kim, B.-G.; Lee, W.H. Optimization of gate-Bias stability and gas-Sensing properties of triethylsilylethynyl anthradithiophene micro-Strip field-EFFECT transistors by incorporating insulating polymer. Org. Electron. 2020, 85, 105878. [Google Scholar] [CrossRef]
  16. Smith, J.; Hamilton, R.; McCulloch, I.; Stingelin-Stutzmann, N.; Heeney, M.; Bradley, D.D.; Anthopoulos, T.D. Solution-Processed organic transistors based on semiconducting blends. J. Mater. Chem. 2010, 20, 2562–2574. [Google Scholar] [CrossRef]
  17. Lee, W.H.; Park, Y.D. Organic semiconductor/insulator polymer blends for high-Performance organic transistors. Polymers 2014, 6, 1057–1073. [Google Scholar] [CrossRef] [Green Version]
  18. Riera-Galindo, S.; Leonardi, F.; Pfattner, R.; Mas-Torrent, M. Organic semiconductor/polymer blend films for organic field-Effect transistors. Adv. Mater. Technol. 2019, 4, 1900104. [Google Scholar] [CrossRef]
  19. Lu, J.; Liu, D.; Zhou, J.; Chu, Y.; Chen, Y.; Wu, X.; Huang, J. Porous organic field-Effect transistors for enhanced chemical sensing performances. Adv. Funct. Mater. 2017, 27, 1700018. [Google Scholar] [CrossRef]
  20. Kang, B.; Jang, M.; Chung, Y.; Kim, H.; Kwak, S.K.; Oh, J.H.; Cho, K. Enhancing 2D growth of organic semiconductor thin films with macroporous structures via a small-Molecule heterointerface. Nat. Commun. 2014, 5, 4752. [Google Scholar] [CrossRef] [Green Version]
  21. Rivnay, J.; Jimison, L.H.; Northrup, J.E.; Toney, M.F.; Noriega, R.; Lu, S.; Marks, T.J.; Facchetti, A.; Salleo, A. Large modulation of carrier transport by grain-boundary molecular packing and microstructure in organic thin films. Nat. Mater. 2009, 8, 952–958. [Google Scholar] [CrossRef]
  22. Virkar, A.A.; Mannsfeld, S.; Bao, Z.; Stingelin, N. Organic semiconductor growth and morphology considerations for organic thin-Film transistors. Adv. Mater. 2010, 22, 3857–3875. [Google Scholar] [CrossRef]
  23. Dai, J.; Ogbeide, O.; Macadam, N.; Sun, Q.; Yu, W.; Li, Y.; Su, B.-L.; Hasan, T.; Huang, X.; Huang, W. Printed gas sensors. Chem. Soc. Rev. 2020, 49, 1756–1789. [Google Scholar] [CrossRef]
  24. Kim, K.; Bae, J.; Noh, S.H.; Jang, J.; Kim, S.H.; Park, C.E. Direct writing and aligning of small-molecule organic semiconductor crystals via “dragging mode” electrohydrodynamic jet printing for flexible organic field-Effect transistor arrays. J. Phys. Chem. Lett. 2017, 8, 5492–5500. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, C.; Chen, P.; Hu, W. Organic field-Effect transistor-based gas sensors. Chem. Soc. Rev. 2015, 44, 2087–2107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Zhang, S.; Zhao, Y.; Du, X.; Chu, Y.; Zhang, S.; Huang, J. Gas sensors based on nano/microstructured organic field-Effect transistors. Small 2019, 15, 1805196. [Google Scholar] [CrossRef] [PubMed]
  27. Yu, S.H.; Cho, J.; Sim, K.M.; Ha, J.U.; Chung, D.S. Morphology-Driven high-Performance polymer transistor-based ammonia gas sensor. ACS Appl. Mater. Interfaces 2016, 8, 6570–6576. [Google Scholar] [CrossRef] [PubMed]
  28. Fan, H.; Han, S.; Song, Z.; Yu, J.; Katz, H.E. Organic field-Effect transistor gas sensor based on GO/PMMA hybrid dielectric for the enhancement of sensitivity and selectivity to ammonia. Org. Electron. 2019, 67, 247–252. [Google Scholar] [CrossRef]
  29. Kumar, A.; Meunier-Prest, R.; Bouvet, M. Organic heterojunction devices based on phthalocyanines: A new approach to gas chemosensing. Sensors 2020, 20, 4700. [Google Scholar] [CrossRef]
  30. Song, R.; Wang, Z.; Zhou, X.; Huang, L.; Chi, L. Gas-Sensing performance and operation mechanism of organic π-conjugated materials. ChemPlusChem 2019, 84, 1222–1234. [Google Scholar] [CrossRef]
  31. Angione, M.D.; Cotrone, S.; Magliulo, M.; Mallardi, A.; Altamura, D.; Giannini, C.; Cioffi, N.; Sabbatini, L.; Fratini, E.; Baglioni, P. Interfacial electronic effects in functional biolayers integrated into organic field-Effect transistors. Proc. Natl. Acad. Sci. USA 2012, 109, 6429–6434. [Google Scholar] [CrossRef] [Green Version]
  32. Lee, Y.H.; Jang, M.; Lee, M.Y.; Kweon, O.Y.; Oh, J.H. Flexible field-Effect transistor-Type sensors based on conjugated molecules. Chem 2017, 3, 724–763. [Google Scholar] [CrossRef] [Green Version]
  33. Andringa, A.M.; Meijboom, J.R.; Smits, E.C.; Mathijssen, S.G.; Blom, P.W.; De Leeuw, D.M. Gate-Bias controlled charge trapping as a mechanism for NO2 detection with Field-Effect transistors. Adv. Funct. Mater. 2011, 21, 100–107. [Google Scholar] [CrossRef] [Green Version]
  34. Paska, Y.; Haick, H. Interactive effect of hysteresis and surface chemistry on gated silicon nanowire gas sensors. ACS Appl. Mater. Interfaces 2012, 4, 2604–2617. [Google Scholar] [CrossRef]
  35. Wu, S.; Wang, G.; Xue, Z.; Ge, F.; Zhang, G.; Lu, H.; Qiu, L. Organic field-Effect transistors with macroporous semiconductor films as high-performance humidity sensors. ACS Appl. Mater. Interfaces 2017, 9, 14974–14982. [Google Scholar] [CrossRef]
  36. Jang, M.; Kim, S.K.; Lee, J.; Ji, S.; Song, W.; Myung, S.; Lim, J.; Lee, S.S.; Jung, H.-K.; Lee, J. Amplifying gas sensor performance by embedding a cellulose-Based buffer layer in organic transistors. J. Mater. Chem. C 2019, 7, 14504–14510. [Google Scholar] [CrossRef]
  37. Lee, J.H.; Seo, Y.; Park, Y.D.; Anthony, J.E.; Kwak, D.H.; Lim, J.A.; Ko, S.; Jang, H.W.; Cho, K.; Lee, W.H. Effect of crystallization modes in TIPS-Pentacene/insulating polymer blends on the gas sensing properties of organic field-Effect transistors. Sci. Rep. 2019, 9, 21. [Google Scholar] [CrossRef] [Green Version]
  38. Kang, B.; Lee, W.H.; Cho, K. Recent advances in organic transistor printing processes. ACS Appl. Mater. Interfaces 2013, 5, 2302–2315. [Google Scholar] [CrossRef]
  39. Kwon, Y.-J.; Park, Y.D.; Lee, W.H. Inkjet-Printed organic transistors based on organic semiconductor/insulating polymer blends. Materials 2016, 9, 650. [Google Scholar] [CrossRef] [Green Version]
  40. Ling, M.M.; Bao, Z. Thin film deposition, patterning, and printing in organic thin film transistors. Chem. Mater. 2004, 16, 4824–4840. [Google Scholar] [CrossRef]
  41. Yan, H.; Chen, Z.; Zheng, Y.; Newman, C.; Quinn, J.R.; Dötz, F.; Kastler, M.; Facchetti, A. A high-Mobility electron-Transporting polymer for printed transistors. Nature 2009, 457, 679–686. [Google Scholar] [CrossRef]
  42. Seekaew, Y.; Lokavee, S.; Phokharatkul, D.; Wisitsoraat, A.; Kerdcharoen, T.; Wongchoosuk, C. Low-Cost and flexible printed graphene–PEDOT: PSS gas sensor for ammonia detection. Org. Electron. 2014, 15, 2971–2981. [Google Scholar] [CrossRef]
  43. Yin, X.; Yang, J.; Wang, H. Vertical Phase Separation Structure for High-Performance Organic Thin-Film Transistors: Mechanism, Optimization Strategy, and Large-Area Fabrication toward Flexible and Stretchable Electronics. Adv. Funct. Mater. 2022, 32, 2202071. [Google Scholar] [CrossRef]
  44. Chou, L.-H.; Na, Y.; Park, C.-H.; Park, M.S.; Osaka, I.; Kim, F.S.; Liu, C.-L. Semiconducting small molecule/polymer blends for organic transistors. Polymer 2020, 191, 122208. [Google Scholar] [CrossRef]
  45. Janasz, L.; Borkowski, M.; Blom, P.W.; Marszalek, T.; Pisula, W. Organic semiconductor/insulator blends for elastic field-Effect transistors and sensors. Adv. Funct. Mater. 2022, 32, 2105456. [Google Scholar] [CrossRef]
  46. Miyata, R.; Inoue, S.; Nakajima, K.; Hasegawa, T. Insulating Polymer Blend Organic Thin-Film Transistors Based on Bilayer-Type Alkylated Benzothieno [3, 2-b] naphtho [2, 3-b] thiophene. ACS Appl. Mater. Interfaces 2022, 14, 17719–17726. [Google Scholar] [CrossRef]
  47. Shen, T.; Zhou, H.; Xin, J.; Fan, Q.; Yang, Z.; Wang, J.; Mei, T.; Wang, X.; Wang, N.; Li, J. Controllable microstructure of polymer-small molecule blend thin films for high-Performance organic field-Effect transistors. Appl. Surf. Sci. 2019, 498, 143822. [Google Scholar] [CrossRef]
  48. Wang, W.; Lu, B.; Deng, W.; Zhang, X.; Lu, Z.; Wu, D.; Jie, J.; Zhang, X. Controlled 2D growth of organic semiconductor crystals by suppressing “coffee-Ring” effect. Nano Res. 2020, 13, 2478–2484. [Google Scholar] [CrossRef]
  49. Lee, S.B.; Lee, S.; Kim, D.G.; Kim, S.H.; Kang, B.; Cho, K. Solutal-Marangoni-Flow-Mediated Growth of Patterned Highly Crystalline Organic Semiconductor Thin Film Via Gap-Controlled Bar Coating. Adv. Funct. Mater. 2021, 31, 2100196. [Google Scholar] [CrossRef]
  50. Anthony, J.E. The larger acenes: Versatile organic semiconductors. Angew. Chem. Int. Ed. 2008, 47, 452–483. [Google Scholar] [CrossRef] [PubMed]
  51. Müller, M.; Ahrens, L.; Brosius, V.; Freudenberg, J.; Bunz, U.H. Unusual stabilization of larger acenes and heteroacenes. J. Mater. Chem. C 2019, 7, 14011–14034. [Google Scholar] [CrossRef]
  52. Natsume, Y.; Minakata, T.; Aoyagi, T. Pentacene thin film transistors fabricated by solution process with directional crystal growth. Org. Electron. 2009, 10, 107–114. [Google Scholar] [CrossRef]
  53. Tao, C.-L.; Zhang, X.-H.; Zhang, F.-J.; Liu, Y.-Y.; Zhang, H.-L. Solution processed pentacene thin films and their structural properties. Mater. Sci. Eng. B 2007, 140, 1–4. [Google Scholar] [CrossRef]
  54. Li, Y.; Wu, Y.; Liu, P.; Prostran, Z.; Gardner, S.; Ong, B.S. Stable solution-Processed high-Mobility substituted pentacene semiconductors. Chem. Mater. 2007, 19, 418–423. [Google Scholar] [CrossRef]
  55. Anthony, J.E.; Brooks, J.S.; Eaton, D.L.; Parkin, S.R. Functionalized pentacene: Improved electronic properties from control of solid-State order. J. Am. Chem. Soc. 2001, 123, 9482–9483. [Google Scholar] [CrossRef]
  56. Diao, Y.; Shaw, L.; Bao, Z.; Mannsfeld, S.C. Morphology control strategies for solution-Processed organic semiconductor thin films. Energy Environ. Sci. 2014, 7, 2145–2159. [Google Scholar] [CrossRef] [Green Version]
  57. Giri, G.; Verploegen, E.; Mannsfeld, S.C.; Atahan-Evrenk, S.; Kim, D.H.; Lee, S.Y.; Becerril, H.A.; Aspuru-Guzik, A.; Toney, M.F.; Bao, Z. Tuning charge transport in solution-sheared organic semiconductors using lattice strain. Nature 2011, 480, 504–508. [Google Scholar] [CrossRef]
  58. Temiño, I.; Del Pozo, F.G.; Ajayakumar, M.; Galindo, S.; Puigdollers, J.; Mas-Torrent, M. A rapid, low-Cost, and scalable technique for printing state-Of-The-Art organic field-Effect transistors. Adv. Mater. Technol. 2016, 1, 1600090. [Google Scholar] [CrossRef] [Green Version]
  59. Payne, M.M.; Parkin, S.R.; Anthony, J.E.; Kuo, C.-C.; Jackson, T.N. Organic Field-Effect Transistors from Solution-Deposited Functionalized Acenes with Mobilities as High as 1 cm2/V·s. J. Am. Chem. Soc. 2005, 127, 4986–4987. [Google Scholar] [CrossRef]
  60. Dickey, K.C.; Anthony, J.E.; Loo, Y.L. Improving organic thin-Film transistor performance through solvent-Vapor annealing of solution-processable triethylsilylethynyl anthradithiophene. Adv. Mater. 2006, 18, 1721–1726. [Google Scholar] [CrossRef]
  61. Lee, W.H.; Kim, D.H.; Cho, J.H.; Jang, Y.; Lim, J.A.; Kwak, D.; Cho, K. Change of molecular ordering in soluble acenes via solvent annealing and its effect on field-effect mobility. Appl. Phys. Lett. 2007, 91, 092105. [Google Scholar] [CrossRef]
  62. Lee, W.H.; Lim, J.A.; Kwak, D.; Cho, J.H.; Lee, H.S.; Choi, H.H.; Cho, K. Semiconductor-Dielectric blends: A facile all solution route to flexible all-Organic transistors. Adv. Mater. 2009, 21, 4243–4248. [Google Scholar] [CrossRef]
  63. Subramanian, S.; Park, S.K.; Parkin, S.R.; Podzorov, V.; Jackson, T.N.; Anthony, J.E. Chromophore fluorination enhances crystallization and stability of soluble anthradithiophene semiconductors. J. Am. Chem. Soc. 2008, 130, 2706–2707. [Google Scholar] [CrossRef]
  64. Hunter, S.; Chen, J.; Anthopoulos, T.D. Microstructural Control of Charge Transport in Organic Blend Thin-Film Transistors. Adv. Funct. Mater. 2014, 24, 5969–5976. [Google Scholar] [CrossRef]
  65. Salzillo, T.; Montes, N.; Pfattner, R.; Mas-Torrent, M. Selection of the two enantiotropic polymorphs of diF-TES-ADT in solution sheared thin film transistors. J. Mater. Chem. C 2020, 8, 15361–15367. [Google Scholar] [CrossRef]
  66. Deng, W.; Xiao, Y.; Lu, B.; Zhang, L.; Xia, Y.; Zhu, C.; Zhang, X.; Guo, J.; Zhang, X.; Jie, J. Water-Surface Drag Coating: A New Route Toward High-Quality Conjugated Small-Molecule Thin Films with Enhanced Charge Transport Properties. Adv. Mater. 2021, 33, 2005915. [Google Scholar] [CrossRef]
  67. Takimiya, K.; Osaka, I.; Mori, T.; Nakano, M. Organic semiconductors based on [1] benzothieno [3, 2-b][1] benzothiophene substructure. Acc. Chem. Res. 2014, 47, 1493–1502. [Google Scholar] [CrossRef]
  68. Yuan, Y.; Giri, G.; Ayzner, A.L.; Zoombelt, A.P.; Mannsfeld, S.C.; Chen, J.; Nordlund, D.; Toney, M.F.; Huang, J.; Bao, Z. Ultra-high mobility transparent organic thin film transistors grown by an off-centre spin-coating method. Nat. Commun. 2014, 5, 3005. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Zhao, Y.; Fan, X.; Feng, J.; Wang, X.; Wu, Y.; Su, B.; Jiang, L. Regulated Dewetting for Patterning Organic Single Crystals with Pure Crystallographic Orientation toward High Performance Field-Effect Transistors. Adv. Funct. Mater. 2018, 28, 1800470. [Google Scholar] [CrossRef]
  70. Wu, M.; Hou, S.; Yu, X.; Yu, J. Recent progress in chemical gas sensors based on organic thin film transistors. J. Mater. Chem. C 2020, 8, 13482–13500. [Google Scholar] [CrossRef]
  71. Lv, A.; Pan, Y.; Chi, L. Gas sensors based on polymer field-Effect transistors. Sensors 2017, 17, 213. [Google Scholar] [CrossRef] [Green Version]
  72. Di, C.-A.; Liu, Y.; Yu, G.; Zhu, D. Interface engineering: An effective approach toward high-Performance organic field-effect transistors. Acc. Chem. Res. 2009, 42, 1573–1583. [Google Scholar] [CrossRef]
  73. Wang, Z.; Huang, L.; Zhu, X.; Zhou, X.; Chi, L. An ultrasensitive organic semiconductor NO2 sensor based on crystalline TIPS-Pentacene films. Adv. Mater. 2017, 29, 1703192. [Google Scholar] [CrossRef]
  74. Lee, B.H.; Kim, S.; Lee, S.Y. Ammonia Gas Sensing Properties of 6, 13-Bis (tri-Isopropylsilyethynyl) Pentacene Based Field-Effect Transistor. Trans. Electr. Electron. Mater. 2022, 23, 182–186. [Google Scholar] [CrossRef]
  75. Shao, B.; Liu, Y.; Zhuang, X.; Hou, S.; Han, S.; Yu, X.; Yu, J. Crystallinity and grain boundary control of TIPS-Pentacene in organic thin-film transistors for the ultra-High sensitive detection of NO2. J. Mater. Chem. C 2019, 7, 10196–10202. [Google Scholar] [CrossRef]
  76. Seo, Y.; Lee, J.H.; Anthony, J.E.; Nguyen, K.V.; Kim, Y.H.; Jang, H.W.; Ko, S.; Cho, Y.; Lee, W.H. Effects of Grain Boundary Density on the Gas Sensing Properties of Triethylsilylethynyl-Anthradithiophene Field-Effect Transistors. Adv. Mater. Interfaces 2018, 5, 1701399. [Google Scholar] [CrossRef]
  77. Nguyen, K.V.; Payne, M.M.; Anthony, J.E.; Lee, J.H.; Song, E.; Kang, B.; Cho, K.; Lee, W.H. Grain boundary induced bias instability in soluble acene-based thin-film transistors. Sci. Rep. 2016, 6, 33224. [Google Scholar] [CrossRef]
  78. Chen, S.; Li, Z.; Qiao, Y.; Song, Y. Solution-Processed organic semiconductor crystals for field-Effect transistors: From crystallization mechanism towards morphology control. J. Mater. Chem. C 2021, 9, 1126–1149. [Google Scholar] [CrossRef]
  79. Hou, S.; Zhuang, X.; Yang, Z.; Yu, J. Effect of vertical annealing on the nitrogen dioxide response of organic thin film transistors. Nanomaterials 2018, 8, 203. [Google Scholar] [CrossRef] [Green Version]
  80. Hou, S.; Zhuang, X.; Fan, H.; Yu, J. Grain boundary control of organic semiconductors via solvent vapor annealing for high-Sensitivity NO2 detection. Sensors 2021, 21, 226. [Google Scholar] [CrossRef]
  81. Li, L.; Gao, P.; Baumgarten, M.; Müllen, K.; Lu, N.; Fuchs, H.; Chi, L. High performance field-Effect ammonia sensors based on a structured ultrathin organic semiconductor film. Adv. Mater. 2013, 25, 3419–3425. [Google Scholar] [CrossRef]
  82. Wang, B.; Ding, J.; Zhu, T.; Huang, W.; Cui, Z.; Chen, J.; Huang, L.; Chi, L. Fast patterning of oriented organic microstripes for field-Effect ammonia gas sensors. Nanoscale 2016, 8, 3954–3961. [Google Scholar] [CrossRef]
  83. Kwak, D.H.; Seo, Y.; Anthony, J.E.; Kim, S.; Hur, J.; Chae, H.; Park, H.J.; Kim, B.G.; Lee, E.; Ko, S. Enhanced gas sensing performance of organic field-Effect transistors by modulating the dimensions of triethylsilylethynyl-anthradithiophene microcrystal arrays. Adv. Mater. Interfaces 2020, 7, 1901696. [Google Scholar] [CrossRef]
  84. Kim, K.; Jang, M.; Lee, M.; An, T.K.; Anthony, J.E.; Kim, S.H.; Yang, H.; Park, C.E. Unified film patterning and annealing of an organic semiconductor with micro-Grooved wet stamps. J. Mater. Chem. C 2016, 4, 6996–7003. [Google Scholar] [CrossRef]
  85. Nguyen, K.V.; Lee, J.H.; Lee, S.C.; Ku, G.M.; Lee, W.H. Simultaneous control of molecular orientation and patterning of small-Molecule organic semiconductors for organic transistors. Org. Electron. 2017, 41, 107–113. [Google Scholar] [CrossRef]
  86. Lee, W.H.; Kwak, D.; Anthony, J.E.; Lee, H.S.; Choi, H.H.; Kim, D.H.; Lee, S.G.; Cho, K. The influence of the solvent evaporation rate on the phase separation and electrical performances of soluble acene-Polymer blend semiconductors. Adv. Funct. Mater. 2012, 22, 267–281. [Google Scholar] [CrossRef]
  87. Zhang, X.; Wang, B.; Huang, L.; Huang, W.; Wang, Z.; Zhu, W.; Chen, Y.; Mao, Y.; Facchetti, A.; Marks, T.J. Breath figure–Derived porous semiconducting films for organic electronics. Sci. Adv. 2020, 6, eaaz1042. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Lee, W.H.; Choi, H.H.; Kim, D.H.; Cho, K. 25th anniversary article: Microstructure dependent bias stability of organic transistors. Adv. Mater. 2014, 26, 1660–1680. [Google Scholar] [CrossRef] [PubMed]
  89. Yu, X.; Zhou, N.; Han, S.; Lin, H.; Buchholz, D.B.; Yu, J.; Chang, R.P.; Marks, T.J.; Facchetti, A. Flexible spray-Coated TIPS-pentacene organic thin-Film transistors as ammonia gas sensors. J. Mater. Chem. C 2013, 1, 6532–6535. [Google Scholar] [CrossRef]
  90. Feng, L.; Tang, W.; Zhao, J.; Yang, R.; Hu, W.; Li, Q.; Wang, R.; Guo, X. Unencapsulated air-stable organic field effect transistor by all solution processes for low power vapor sensing. Sci. Rep. 2016, 6, 20671. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Lee, W.H.; Kim, D.H.; Jang, Y.; Cho, J.H.; Hwang, M.; Park, Y.D.; Kim, Y.H.; Han, J.I.; Cho, K. Solution-Processable pentacene microcrystal arrays for high performance organic field-effect transistors. Appl. Phys. Lett. 2007, 90, 132106. [Google Scholar] [CrossRef]
Figure 1. Chemical structures of prototypical soluble acenes: (a) 6,13-bis(triisopropylsilylethynyl)pentacene (TIPS-pentacene), (b) 5,11-bis(triethylsilylethynyl)anthradithiophene (TES-ADT), (c) 2,8-difluoro-5,11-bis(triethylsilylethynyl)anthradithiophene (diF-TES-ADT), (d) 2,7-dihexyl-dithieno[2,3-d;2′,3′-d′]benzo[1,2-b;4,5-b′]dithiophene, and (e) 2,7-dioctylbenzothieno[2,3-b]benzothiophene (C8-BTBT).
Figure 1. Chemical structures of prototypical soluble acenes: (a) 6,13-bis(triisopropylsilylethynyl)pentacene (TIPS-pentacene), (b) 5,11-bis(triethylsilylethynyl)anthradithiophene (TES-ADT), (c) 2,8-difluoro-5,11-bis(triethylsilylethynyl)anthradithiophene (diF-TES-ADT), (d) 2,7-dihexyl-dithieno[2,3-d;2′,3′-d′]benzo[1,2-b;4,5-b′]dithiophene, and (e) 2,7-dioctylbenzothieno[2,3-b]benzothiophene (C8-BTBT).
Nanomaterials 12 02564 g001
Figure 2. Gas sensing mechanism of organic-effect transistor gas sensors. OSC represents organic semiconductor.
Figure 2. Gas sensing mechanism of organic-effect transistor gas sensors. OSC represents organic semiconductor.
Nanomaterials 12 02564 g002
Figure 3. (a) Device configuration of TIPS-pentacene FET sensors, (b) sensor setup system, (c) dynamic response of the TIPS-pentacene/para-sexiphenyl (p-6P) film upon periodic exposure to NO2 (1–5 ppm), and (d) responsivity (R) and sensitivity (S) to different gases (NO2, SO2, wet air, and NH3) [74]. Copyright 2017 Wiley.
Figure 3. (a) Device configuration of TIPS-pentacene FET sensors, (b) sensor setup system, (c) dynamic response of the TIPS-pentacene/para-sexiphenyl (p-6P) film upon periodic exposure to NO2 (1–5 ppm), and (d) responsivity (R) and sensitivity (S) to different gases (NO2, SO2, wet air, and NH3) [74]. Copyright 2017 Wiley.
Nanomaterials 12 02564 g003
Figure 4. (a) The device structure and sensing mechanism of TIPS-pentacene sensors showing diffusion of NO2 through the grain boundaries, (b) the change in the transfer curves for the o-xylene, toluene, chlorobenzene (CB), and 1,2-dichlorobenzene (1,2-DCB)-processed TIPS-pentacene FETs upon NO2 exposure at various concentrations, (c) atomic force microscopy images of TIPS-pentacene films processed with different solvents [75]. Copyright 2019 Royal Society of Chemistry. (d) Polarized optical microscopy images of the solvent vapor annealed TES-ADT films whose solutions were mixed for 5 min or 12 h, and (e) response curves and summarized performance of TES-ADT FETs upon sequential NO2 and N2 exposure [76]. Copyright 2017 Wiley.
Figure 4. (a) The device structure and sensing mechanism of TIPS-pentacene sensors showing diffusion of NO2 through the grain boundaries, (b) the change in the transfer curves for the o-xylene, toluene, chlorobenzene (CB), and 1,2-dichlorobenzene (1,2-DCB)-processed TIPS-pentacene FETs upon NO2 exposure at various concentrations, (c) atomic force microscopy images of TIPS-pentacene films processed with different solvents [75]. Copyright 2019 Royal Society of Chemistry. (d) Polarized optical microscopy images of the solvent vapor annealed TES-ADT films whose solutions were mixed for 5 min or 12 h, and (e) response curves and summarized performance of TES-ADT FETs upon sequential NO2 and N2 exposure [76]. Copyright 2017 Wiley.
Nanomaterials 12 02564 g004
Figure 5. (a) Scanning electron microscopy image of C6-BTBDT microstripes, (b) sensing mechanism of C6-BTBDT with NH3, (c) sensitivity, response time, and recovery time of C6-BTBDT FET sensors according to the target gas molecules [81]. Sensitivity is defined as INH3-off/INH3-on. Copyright 2013 Wiley. (d) Fabrication procedure and scanning electron microscopy images of C6-BTBDT microstripes on Au source/drain electrodes [82]. Copyright 2016 Royal Society of Chemistry.
Figure 5. (a) Scanning electron microscopy image of C6-BTBDT microstripes, (b) sensing mechanism of C6-BTBDT with NH3, (c) sensitivity, response time, and recovery time of C6-BTBDT FET sensors according to the target gas molecules [81]. Sensitivity is defined as INH3-off/INH3-on. Copyright 2013 Wiley. (d) Fabrication procedure and scanning electron microscopy images of C6-BTBDT microstripes on Au source/drain electrodes [82]. Copyright 2016 Royal Society of Chemistry.
Nanomaterials 12 02564 g005
Figure 6. (a) Fabrication procedure of TES-ADT microstripes using a polydimethylsiloxane (PDMS) mold and the types of fabricated microstripes: (A) rectangular patterns of 200 µm width, (B) line pattern of 65 µm width, and (C) line patterns of 25 µm width. The white arrows indicate the grain boundary. (b) gas-sensing characteristics and summarized performance according to the types of microstripes, (c) comparison curve of the sensitivity versus pathway length (sum of grain boundary length and edge length), and (d) schematic showing gas diffusion in TES-ADT microstripes [83]. Copyright 2020 Wiley.
Figure 6. (a) Fabrication procedure of TES-ADT microstripes using a polydimethylsiloxane (PDMS) mold and the types of fabricated microstripes: (A) rectangular patterns of 200 µm width, (B) line pattern of 65 µm width, and (C) line patterns of 25 µm width. The white arrows indicate the grain boundary. (b) gas-sensing characteristics and summarized performance according to the types of microstripes, (c) comparison curve of the sensitivity versus pathway length (sum of grain boundary length and edge length), and (d) schematic showing gas diffusion in TES-ADT microstripes [83]. Copyright 2020 Wiley.
Nanomaterials 12 02564 g006
Figure 7. Schematic showing phase-separation behaviors of TIPS-pentacene/polymer blends during the evaporation of the residual solvent [86]. Copyright 2012 Wiley. Typically, a vertically phase-separated structure consisting of TIPS-pentacene at the top and polymer at the bottom is formed on SiO2/Si.
Figure 7. Schematic showing phase-separation behaviors of TIPS-pentacene/polymer blends during the evaporation of the residual solvent [86]. Copyright 2012 Wiley. Typically, a vertically phase-separated structure consisting of TIPS-pentacene at the top and polymer at the bottom is formed on SiO2/Si.
Nanomaterials 12 02564 g007
Figure 8. Spin-coating time (5 s or 50 s) governs the 1D versus 2D growth of TIPS-pentacene on a polymer, affecting the performance of FETs and gas sensors [37]. Copyright 2018 Wiley. (a) Mechanism and polarized optical microscopy images of TIPS-pentacene crystals and (b) NO2 sensing performance and schematics showing the gas-sensing mechanism [37]. Copyright 2019 Nature Publishing Group.
Figure 8. Spin-coating time (5 s or 50 s) governs the 1D versus 2D growth of TIPS-pentacene on a polymer, affecting the performance of FETs and gas sensors [37]. Copyright 2018 Wiley. (a) Mechanism and polarized optical microscopy images of TIPS-pentacene crystals and (b) NO2 sensing performance and schematics showing the gas-sensing mechanism [37]. Copyright 2019 Nature Publishing Group.
Nanomaterials 12 02564 g008
Figure 9. (a) Fabrication procedure for a porous organic semiconductor film, (b) atomic force microscopy image of a porous C8-BTBT/polystyrene (PS) film, and (c) sensitivities of FETs-based gas sensors based on dense/porous C8-BTBT/PS and poly[N,N′-bis(2-octyldodecyl)-naphthalene-1,4,5,8-bis(dicarboximide)-2,6-diyl]-alt-5,5′-(2,2′-bithio-phene) (N2200)/PS films toward different NH3 concentrations (ppm) [87]. Sensitivity is defined as S = [(IGasI0)/I0]/[NH3] × 100%. Copyright 2020 AAAS.
Figure 9. (a) Fabrication procedure for a porous organic semiconductor film, (b) atomic force microscopy image of a porous C8-BTBT/polystyrene (PS) film, and (c) sensitivities of FETs-based gas sensors based on dense/porous C8-BTBT/PS and poly[N,N′-bis(2-octyldodecyl)-naphthalene-1,4,5,8-bis(dicarboximide)-2,6-diyl]-alt-5,5′-(2,2′-bithio-phene) (N2200)/PS films toward different NH3 concentrations (ppm) [87]. Sensitivity is defined as S = [(IGasI0)/I0]/[NH3] × 100%. Copyright 2020 AAAS.
Nanomaterials 12 02564 g009
Figure 10. (a) Fabrication procedure of TES-ADT/poly(methyl methacrylate) (PMMA) microstripes using a PDMS mold, (b) polarized optical microscopy images of the fabricated microstripes from TES-ADT and TES-ADT/PMMA blends, (c) gas-sensing characteristics toward NO2 according to the types of microstripes, and (d) gate-bias stabilities at continuous bias conditions (VGS = −20 V, VDS = −10 V) [15]. Copyright 2020 Elsevier.
Figure 10. (a) Fabrication procedure of TES-ADT/poly(methyl methacrylate) (PMMA) microstripes using a PDMS mold, (b) polarized optical microscopy images of the fabricated microstripes from TES-ADT and TES-ADT/PMMA blends, (c) gas-sensing characteristics toward NO2 according to the types of microstripes, and (d) gate-bias stabilities at continuous bias conditions (VGS = −20 V, VDS = −10 V) [15]. Copyright 2020 Elsevier.
Nanomaterials 12 02564 g010
Figure 11. (a) Schematic showing the spray coating of TIPS-pentacene on a PMMA/substrate, chemical structures of the used materials, and gas sensor structure of the FETs, (b) changes in the output curves before (N2) and after target gas (NH3, 100 ppm) exposure, (c) camera image of a flexible TIPS-pentacene gas sensor, and (d) transfer curves of a flexible gas sensor before (N2) and after inserting NH3 with different concentrations [89]. Copyright 2013 Royal Society of Chemistry.
Figure 11. (a) Schematic showing the spray coating of TIPS-pentacene on a PMMA/substrate, chemical structures of the used materials, and gas sensor structure of the FETs, (b) changes in the output curves before (N2) and after target gas (NH3, 100 ppm) exposure, (c) camera image of a flexible TIPS-pentacene gas sensor, and (d) transfer curves of a flexible gas sensor before (N2) and after inserting NH3 with different concentrations [89]. Copyright 2013 Royal Society of Chemistry.
Nanomaterials 12 02564 g011
Figure 12. (a) Device structure of a flexible FET gas sensor based on a TIPS-pentacene/PS blend, (b) optical microscopy image of a TIPS-pentacene/PS blend film, (c) camera image of a flexible FET gas sensor, and (d) relative change of Vout with time upon NH3 inlet/outlet with different concentrations [90]. Copyright 2016 Nature Publishing Group.
Figure 12. (a) Device structure of a flexible FET gas sensor based on a TIPS-pentacene/PS blend, (b) optical microscopy image of a TIPS-pentacene/PS blend film, (c) camera image of a flexible FET gas sensor, and (d) relative change of Vout with time upon NH3 inlet/outlet with different concentrations [90]. Copyright 2016 Nature Publishing Group.
Nanomaterials 12 02564 g012
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lee, J.H.; Chun, J.H.; Chung, H.-J.; Lee, W.H. Microstructural Control of Soluble Acene Crystals for Field-Effect Transistor Gas Sensors. Nanomaterials 2022, 12, 2564. https://doi.org/10.3390/nano12152564

AMA Style

Lee JH, Chun JH, Chung H-J, Lee WH. Microstructural Control of Soluble Acene Crystals for Field-Effect Transistor Gas Sensors. Nanomaterials. 2022; 12(15):2564. https://doi.org/10.3390/nano12152564

Chicago/Turabian Style

Lee, Jung Hun, Jeong Hwan Chun, Hyun-Jong Chung, and Wi Hyoung Lee. 2022. "Microstructural Control of Soluble Acene Crystals for Field-Effect Transistor Gas Sensors" Nanomaterials 12, no. 15: 2564. https://doi.org/10.3390/nano12152564

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