Next Article in Journal
Synergistic Effects of Radical Distributions of Soluble and Insoluble Polymers within Electrospun Nanofibers for an Extending Release of Ferulic Acid
Previous Article in Journal
Study on the Diffusion Characteristics of Polymer Grouting Materials Applied for Crack Filling in Underground Mines Based on Numerical Simulation and Experimental Methods
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Patterning of Organic Semiconductors Leads to Functional Integration: From Unit Device to Integrated Electronics

1
Department of Semiconductor Engineering, Gachon University, Seongnam 13120, Republic of Korea
2
Department of Electronic Engineering, Gachon University, Seongnam 13120, Republic of Korea
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Polymers 2024, 16(18), 2613; https://doi.org/10.3390/polym16182613
Submission received: 6 August 2024 / Revised: 31 August 2024 / Accepted: 13 September 2024 / Published: 15 September 2024
(This article belongs to the Special Issue Polymer-Based Smart Materials: Preparation and Applications)

Abstract

:
The use of organic semiconductors in electronic devices, including transistors, sensors, and memories, unlocks innovative possibilities such as streamlined fabrication processes, enhanced mechanical flexibility, and potential new applications. Nevertheless, the increasing technical demand for patterning organic semiconductors requires greater integration and functional implementation. This paper overviews recent efforts to pattern organic semiconductors compatible with electronic devices. The review categorizes the contributions of organic semiconductor patterning approaches, such as surface-grafting polymers, capillary force lithography, wettability, evaporation, and diffusion in organic semiconductor-based transistors and sensors, offering a timely perspective on unconventional approaches to enable the patterning of organic semiconductors with a strong focus on the advantages of organic semiconductor utilization. In addition, this review explores the opportunities and challenges of organic semiconductor-based integration, emphasizing the issues related to patterning and interconnection.

1. Introduction

Organic semiconductors (OSCs) combine the electronic properties of semiconductors and organic compounds [1,2,3,4,5]. Efforts regarding OSCs are actively being made because of their flexibility and even stretchable properties [6,7,8,9]. Furthermore, they provide a low-cost process and easy control of physical properties using thin film deposition techniques [10,11,12,13], making them highly advantageous for various applications. OSCs are used in a wide range of devices, including light-emitting diodes [14,15,16], photodetectors [17,18,19,20], and thin film transistors [21,22,23,24]. In addition to the conventional functions mentioned above, emerging functional devices made from OSCs are currently being investigated, and representative examples include neuromorphic devices [25,26,27,28], physically unclonable functions [29,30,31,32,33], logic devices [34,35,36], and multi-functional sensors [37,38,39,40]. The cost-effective and room-temperature compatibility of solution-based processes can reduce the cost of existing electronic products significantly and enable disposable [41,42,43,44] large-area electronic applications [44,45,46,47]. OSCs typically exist as thin films in devices. Hence, researchers have developed various thin film preparation techniques based on solution processes or vacuum deposition. On the other hand, although commercially successful silicon-based semiconductors are used in industry, OSCs still face integration challenges. To progress toward the commercialization of organic semiconductors, it is essential to improve their sensitivity to environmental stresses such as chemical solvents [48], heat [49], and UV exposure [50] during the patterning process. This improvement is particularly important for large-area integration, which is essential for developing practical applications that require precise patterning processes for OSCs.
Patterning OSCs is challenging because they are incompatible with photoresist-based photolithography processes. Most OSCs are vulnerable to physical and chemical damage from solvents, making it difficult to use conventional photoresists and achieve compatibility with etching and other existing patterning processes. Accordingly, considerable efforts have been made to develop alternative methods for patterning OSCs. Techniques that consider the vulnerability of OSCs by avoiding conventional photoresist and etching processes have been developed. Patterning OSCs is being reported by controlling the chemical and physical energy between organic semiconductor materials and substrates [51,52,53,54]. This patterning technique not only avoids the need for photoresist and etching processes but also improves the crystallinity and crystal orientation of the organic semiconductor [55,56,57,58,59,60], reducing defects and enhancing charge transport [61,62,63,64,65]. These advances have significantly contributed to the development of OSCs from the unit device to the integration of multiple devices, extending their application to fields where the physical properties of organic materials change due to external energy, such as in-time temperature integrator devices [66,67,68] and fluorescence sensors [69,70,71].
In this context, this paper presents a focused review of various approaches to pattern OSCs, emphasizing the control of the chemical and physical energy between organic semiconductor materials and substrates. This review provides a comprehensive summary of principles underlying the induction of patterns in OSCs, including surface energy control of substrates, chemical reactions of materials, and interactions among solvents. Furthermore, it provides specific examples of controlling crystallinity in OSCs. Furthermore, application examples are provided for electronic components, such as transistors and sensors, utilizing patterned OSCs. This review contributes to the understanding and advancement of patterning techniques critical for integrating OSCs into practical electronic devices.

2. Patterning Methods of Organic Materials

2.1. Surface-Grafting Polymers

Surface-grafting polymers are produced using a technique that induces covalent bonding between a substrate and an organic material through chemical treatment, allowing the organic material to grow [72,73,74,75]. This method enables localized chemical treatments, facilitating the patterning of organic materials [76]. The methodologies for producing surface-grafting polymers can be classified into “grafting to” and “grafting from” methods. “Grafting to” methods involve the terminal group of a preformed polymer binding to a chemically modified substrate surface, resulting in the polymer chain grafting onto the substrate surface [77,78]. “Grafting from” methods involve polymer chains being grafted via surface-initiated polymerization reactions of monomers with an initiator chemically bound to the substrate surface [79,80]. This approach allows simple surface energy control, which is useful for patterning through partial chemical reactions. Narupai et al. reported a strategy for applying a precursor solution containing a mixture of a monomer and photoredox catalyst onto a surface-initiated silicon wafer to pattern the polymer brushes of 2-(dimethylamino)ethyl methacrylate (DMAEMA), followed by covering with a glass cover and exposure to light (Figure 1a) [54]. The initiator was undecyl-2-bromo-2-methylpropanoate, with a visible LED with a 405 nm wavelength applied to activate the photocatalyst. The glass cover acts as a barrier to oxygen, facilitating polymer growth in an inert atmosphere within a glovebox or a chamber designed for specific environments. This strategy allows the straightforward patterning of organic materials without expensive equipment. Intricate organic patterns with a resolution of 20 × 200 μm can be achieved by extending the glass cover to serve as a photomask (Figure 1b). Interestingly, p(DMAEMA) patterned on a 20 × 200 μm scale exhibited uniform polymer patterning regardless of the presence of a patterned polymer in the underlying layer (X: substrate, Y: patterned polymer), as shown in Figure 1c,d. Under ambient conditions, large-scale p(DMAEMA) patterning was achieved on a 4-inch silicon wafer functionalized with the initiator by simultaneously using coverslips with the shapes of the letters ‘M’, ‘R’, and ‘L’ (Figure 1e–h). Following the proposed strategy, 200 μL of a monomer precursor solution was dropped, and individual letter cover slips were placed. Under 405 nm LED illumination, polymer patterns in the shapes of the letters were formed, demonstrating spatial control over the patterning process. Based on these results, various iridium-based monomers that emit under 254 nm illumination were grafted simultaneously, producing multicolored patterned polymer brushes.
The surface-grafting polymer method can improve the crystal quality and morphology of OSCs by controlling the chain length of the grafted polymers. Improvements in the crystallinity of inkjet-printed small organic semiconductor films and their influence on device performance by controlling the chain length of grafted polystyrene (PS) was reported by Ge et al. [57]. Figure 1i shows the method for synthesizing PS brushes, where the PS chain lengths were variably synthesized by controlling polymerization times at 100 °C. PS grafting onto SiO2/Si substrates enhances compatibility with the organic solvent tetralin, resulting in strong pinning of the three-phase contact line (Figure 1j). Between the PS brush and organic semiconductor ink, strong pinning of the three-phase contact line and heterogeneous nucleation enhance the crystallization of the organic semiconductor by restricting the outward migration of organic molecules. As a result, organic molecules crystallize from the edges toward the center, preventing the “coffee ring” effect during crystallization. Based on this mechanism, Figure 1k presents a schematic diagram of the crystal growth of OSCs as a function of the PS brush chain length. As the grafted polymer chain length becomes longer, the surface energy of the PS brush increases, causing the printed ink to spread over a wider diameter. This phenomenon enhances the crystallization rate at the receding contact line as the solvent evaporates. Using a PS brush with a chain length of 5.14 mm, 6,13-Bis(triisopropylsilylethynyl)pentacene (TIPS-pentacene) was printed to achieve a mobility of 0.35 ± 0.23 cm2·V−1·S−1, demonstrating the fabrication of organic thin film transistors (Figure 1l).

2.2. Capillary Force Lithography

Capillary force lithography has been studied extensively to achieve patterning and high-quality crystal formation in OSCs [81,82,83,84]. Capillary force lithography is a method for patterning materials based on the phenomenon occurring where a liquid contacts and ascends along the walls of capillary tubes. The procedure involves depositing a polymer with a glass transition temperature (Tg) onto a substrate, followed by applying an elastomeric mold, such as polydimethylsiloxane (PDMS), on top. As the polymer on the substrate reaches temperatures above its Tg, it undergoes a change in physical properties, forming an isotropic liquid that can spontaneously spread over the wetted surfaces. This results in the filling of voids in the mold due to capillary forces, causing localized patterning [85]. Deng et al. reported high-resolution patterned single-crystal organic TFT arrays based on capillary platforms fabricated by photolithography [55]. Figure 2a presents a schematic diagram of the capillary force-based patterning process for 2,7-dioctyl [1]benzothieno[3,2-b][1]benzothiophene (C8-BTBT). The hydrophobic CYTOP mold, fabricated using etching and photolithography techniques, revealed localized hydrophilic SiO2 surfaces, inducing a significant wetting contrast that drives the patterning of organic materials. C8-BTBT was dispersed in dichloromethane and deposited onto the substrate in film form as a deep coating. The film was heated above the Tg of C8-BTBT (T = 120 °C). At this temperature, the C8-BTBT film transitions from a solid to a liquid state, allowing it to flow into the voids of the CYTOP mold and form a patterned 1D crystal array (Figure 2b). Patterning was achieved in a structure where capillary forces opposed gravity, demonstrating the dominance of capillary forces in the process. Figure 2c shows the mechanism of the proposed process. The voids of the CYTOP mold are filled with air (Figure 2c(i)). Upon heating above its Tg, the liquefied C8-BTBT flows downward along the hydrophilic plasma-treated CYTOP sidewalls (Figure 2c(ii)). As the liquid C8-BTBT flows along the sidewalls, it wets the hydrophilic SiO2 surface and forms droplets (Figure 2c(iii)). These droplets transform into upward concave meniscus shapes as the SiO2 surface is gradually filled with liquid C8-BTBT (Figure 2c(iv,v)). The liquid C8-BTBT on CYTOP moves into the voids of the mold because of capillary forces (Figure 2c(vi,vii)), forming a uniformly distributed C8-BTBT liquid over time (Figure 2c(viii)). OSCs patterned using capillary force lithography were fabricated into organic thin film transistors (OTFTs) arranged in a 13 × 13 array (Figure 2d). The single-crystal C8-BTBT in the unit device exhibited a well-organized form without defects such as voids (Figure 2e,f). The fabricated TFTs achieved a mobility of 5.78 cm2·V−1·S−1 and an on/off ratio of ~106 (Figure 2g).
One of the techniques widely used in organic semiconductor patterning is lithographically controlled wetting (LCW), which utilizes capillary forces driven by surface tension in the formed capillaries [86,87,88]. In 2018, Zhao et al. developed organic field-effect transistors (OFETs) by producing one-dimensional crystalline C8-BTBT patterns using capillary bridges fabricated through photolithography and ion etching [60]. Capillary forces were induced through the LCW technique using micropillar-structure bridges (capillary bridges) with controlled wettability differences between the top and sidewalls, allowing the organic semiconductor solution to form unidirectionally aligned nanowires at the tops of the capillary bridges. The sidewalls of the fabricated hydrophobic template were modified to make them hydrophobic using heptadecafluorodecyltrimethoxysilane (FAS) molecules. The organic semiconductor was patterned by dropping a precursor solution onto the asymmetrically wettable topographical capillary bridges, which continuously covered the flat substrate (Figure 2h). The liquid between the pillars and the substrate of the capillary bridges exhibited capillary action with an upward concave meniscus as the solution evaporated (Figure 2i). As the solvent evaporated, the concentration of the liquid increased, with capillary flow from the asymmetrically wettable pillars directed toward the substrate, resulting in the nucleation and growth of the organic semiconductor on the target substrate. A single-crystal C8-BTBT belt array OFET was fabricated using this phenomenon. The performance enhancement of this device was confirmed through a comparison with OFETs based on spin-coated C8-BTBT thin films (Figure 2j). At the same gate voltage (VGS), the patterned single-crystal C8-BTBT belt array OFET achieved a higher drain current (IDS) than the thin film C8-BTBT OFET, with the hole mobility increasing more than 79-fold. The performance-validated C8-BTBT belt array was transferred onto a PET substrate for application as a pressure sensor (Figure 2k,l). The C8-BTBT belt array used as a flexible channel exhibited stable current variations when pressure was applied to the device (Figure 2m).

2.3. Wettability

The development of techniques for local control of surface wettability of substrates has allowed the widespread use of these methods in various patterning fields [89,90,91,92]. Wettability is generally classified into hydrophilicity and hydrophobicity. Hydrophilic surfaces exhibit a high affinity for water, resulting in a contact angle of less than 90°. Hydrophobic surfaces, however, have a low affinity for water and a contact angle of 90° or more, tending to repel water. Through local control of the surface wettability of the substrate through external chemical or physical treatments, the OSC solution can form patterned films with specific shapes as a result of growing along the adjusted surface energy regions. Hydrophilic surfaces are used in thin film coating industries to deposit aqueous liquids uniformly, while hydrophobic surfaces are used to restrict deposition areas and control pattern sizes in inkjet printing industries [93,94,95,96]. The ability to control the wettability of substrates provides a simple and flexible method to modify surfaces into desired patterns, leading to broad application possibilities in the lab-on-a-chip field. Fang et al. presented a method for patterning C8-BTBT liquid crystalline (LC) thin films in high-performance organic integrated circuits based on the synergistic combination of inkjet printing and melting treatment (IJP-MP) [51]. This approach uses wetting/non-wetting patterned surfaces to achieve precise spatial control of the LC solution during inkjet printing, which allows the production of high-quality C8-BTBT LC crystals and high-resolution patterns. On the other hand, in the conventional inkjet printing process, as the solution ejected from the nozzle onto the substrate dries, the solute tends to move toward the edges, causing the coffee-ring effect [97]. This can lead to lower crystal quality and poor crystal orientation [94,97]. Figure 3a presents a schematic diagram of the process of IJP-MP for patterning C8-BTBT LC films. Initially, a C8-BTBT solution ink (≈200 pL) was printed selectively onto a mold constructed from a hydrophobic photoresist. The frameworks within the mold were composed of divinyltetramethyldisiloxane bis(benzocyclobutene) (BCB), which possesses hydrophilic properties. As a result, the deposited ink droplets were confined to the hydrophilic areas and did not spread into the hydrophobic patterned photoresist.
As the ink gradually evaporated, ring-shaped crystalline aggregates formed at the edges of the frameworks due to the well-known “coffee ring” effect. Subsequently, the sample was heated above the melting point of C8-BTBT (T = 102 °C) in ambient air, causing the aggregated crystals to transform into an isotropic liquid phase, which spread spontaneously to form a thin liquid film within the framework. The sample was then slowly cooled to 95 °C at a rate of 1 °C min−1, allowing a phase transition from the isotropic liquid phase to the crystalline phase. Finally, the C8-BTBT LC films formed within each framework, resulting in a high-resolution patterned structure. Figure 3b presents an optical microscopy image of an OFET array fabricated with patterned C8-BTBT LC films. The patterned C8-BTBT LC film formed a uniform thin film without voids (Figure 3c). Figure 3d plots the overlapping transfer curves of 49 OFETs measured in the saturation region. All devices operated correctly, achieving 100% yield with similar transfer curves, showing that the patterned OSC layer effectively minimized crosstalk between devices. The carrier mobility distribution of the 7 × 7 OFET array had a uniform color range (Figure 3e), confirming the high-quality patterning of C8-BTBT LC films achieved using the IJP-MP method.
In 2021, Lee et al. introduced a novel method for depositing OSC single-crystal thin films using a gap-controlled bar-coating method combined with a wetting patterning strategy and soluble Marangoni flow [58]. This simple bar-coating approach forms arrays of single-crystal OSC thin film patterns at speeds of millimeters per second [58,98,99,100]. Figure 3f presents a schematic diagram of the bar-coating system for processing patterned OSC thin films. The SiO2/Si surface was patterned using octadecyltrichlorosilane (ODTS) to induce the selective wetting of a C8-BTBT/poly[bis(4-phenyl)(2,4,6-trimethyl-phenyl)amine] (PTAA)/toluene ink. The patterned areas were exposed to ultraviolet/ozone (UV/O3) for an appropriate time. The gap between the substrate and the coating bar was maintained at 10 μm, into which the ink was placed. As the bar moved, the solvent (i.e., toluene mixture) began to evaporate along the meniscus line formed in the bar–substrate gap, and the solutes (C8-BTBT and PTAA) were supersaturated along the solution contact line. Supersaturation increased the nucleation and crystal growth of C8-BTBT molecules. The active ink contains solvent additives (1 vol%) that can influence the flow at the contact lines because of their different physical properties compared to the primary solvent. The concentration of additives can vary according to the location, resulting in a surface tension gradient in the solution. This surface tension gradient, induced by the mixed solvent, generates Marangoni flow along the meniscus line [101,102,103]. This solutal Marangoni flow can improve film uniformity and promote the formation of a high-crystallinity patterned OSC film formation by enhancing vertical phase separation [104,105].
A significant difference in surface energy between the patterned and non-patterned regions of the substrate is detrimental to crystal growth. The solution does not adhere selectively to the printed patterns when the surface energy difference between the wetting and dewetting regions is small (Δγ ≤ 21.1 mN m−1) because of the weak differences in wettability. In contrast, a large surface energy difference between the wetting and dewetting regions (Δγ ≥ 31.5 mN·m−1) results in additional nucleation along the pattern walls, yielding polycrystalline C8-BTBT growth (Figure 3g). This occurs because the solution spreads from the dewetting region to the wetting region, forming a small contact angle (θc ≤ 10°). The small θc indicates rapid solvent evaporation and supersaturation at the pattern edges, leading to additional (undesirable) nucleation sites and randomly oriented polycrystalline growth. With an appropriate difference in surface energy (Δγ = 21.1 mN·m−1), the solution can still maintain a high θc at the pattern edges. These results extend the understanding of the fundamental fluid dynamics and crystallization processes of OSC molecules in the gap-controlled bar-coating system.
As another example, a simple and efficient selective surface modification method using atmospheric pressure RF plasma (apRFp) and electrohydrodynamic (EHD) jet printing was reported by Lee et al. [106]. apRFp can discharge at atmospheric pressure and room temperature, allowing a simple and cost-effective modification process. This method allows easy control of surface wettability by selecting appropriate gases and processing conditions [107,108,109]. The EHD jet printing technology overcomes limitations by controlling the electric field between the nozzle and substrate to emit droplets or jets of various volumes [110,111,112]. Thus, EHD technology can be used to fabricate small patterns ranging from the microscale to the nanoscale using solutions of almost any viscosity.
Figure 3h presents the process of locally patterning the hydrophilicity of a substrate using apRFp and EHD jet printing. Radicals from aldehyde groups generated by apRFp attach to the substrate surface, forming a superhydrophobic layer. EHD jet printing locally prints polar alcohols, such as butyl alcohol, onto substrates with modified hydrophilic surfaces. In this process, radicals from aldehyde groups on the substrate surface chemically react with polar alcohols and are removed, making the surface hydrophilic. Areas with different wettabilities can be formed on the same substrate by selectively patterning hydrophilic regions on the surface. Templates produced using this technique can have poly(3,4-ethylenedioxythiophene)-poly(styrenesulfonate) ink, Luria–Bertani broth containing Salmonella bacteria, or polystyrene fluorescent nanoparticle suspension solutions deposited onto the hydrophilic regions by immersing them in the solutions or applying the solutions dropwise. Polystyrene nanoparticles were coated onto the template, and the emitted fluorescence confirmed that the materials had been coated uniformly onto the hydrophilic regions (Figure 3i). EHD jet printing is a promising technology that can reduce undesirable defects, costs, and complex processing steps associated with conventional photolithography and overcome the resolution limitations of traditional inkjet printing techniques. Considering the excellent versatility of this patterning process and the high quality of patterned films, it has great potential for high-performance organic integrated circuits and OFET arrays.
Figure 3. (a) Schematic diagram showing the patterning process for C8-BTBT LC films using the IJP-MP method. (b) Polarized optical microscopy image of a 7 × 7 OFET array patterned with C8-BTBT LC film. (c) Polarized optical microscopy images of a typical OFET device. (d) Transfer characteristics measured from 49 OFETs within the 7 × 7 device array. (e) Spatial distribution of saturation mobilities across the 49 OFET devices [51]. Copyright © 2021, John Wiley and Sons. (f) Schematic diagram showing the fabrication of patterned OSC crystals on a wettability-controlled substrate using the bar-coating method. The inset details the pattern design details and an enlarged view of the film crystallization processes. (g) Cross-sectional and top views of the meniscus and contact line at the interface between the wetting (white) and dewetting (light blue) regions with high (Δγ ≥ 31.5 mN·m−1) and appropriate (Δγ = 21.1 mN·m−1) surface tension differences [58]. Copyright © 2021, John Wiley and Sons. (h) Polarized optical microscopy images of coated films on substrates with large and small surface tension differences. Preparation of a superhydrophobic surface on a silicon wafer using atmospheric RF plasma, followed by the formation of a selective hydrophilic area using an EHD jet printing system. (i) fluorescent image of polystyrene nanoparticles after deposition (scale bar: 50 μm) [106]. Copyright © 2017, Elsevier.
Figure 3. (a) Schematic diagram showing the patterning process for C8-BTBT LC films using the IJP-MP method. (b) Polarized optical microscopy image of a 7 × 7 OFET array patterned with C8-BTBT LC film. (c) Polarized optical microscopy images of a typical OFET device. (d) Transfer characteristics measured from 49 OFETs within the 7 × 7 device array. (e) Spatial distribution of saturation mobilities across the 49 OFET devices [51]. Copyright © 2021, John Wiley and Sons. (f) Schematic diagram showing the fabrication of patterned OSC crystals on a wettability-controlled substrate using the bar-coating method. The inset details the pattern design details and an enlarged view of the film crystallization processes. (g) Cross-sectional and top views of the meniscus and contact line at the interface between the wetting (white) and dewetting (light blue) regions with high (Δγ ≥ 31.5 mN·m−1) and appropriate (Δγ = 21.1 mN·m−1) surface tension differences [58]. Copyright © 2021, John Wiley and Sons. (h) Polarized optical microscopy images of coated films on substrates with large and small surface tension differences. Preparation of a superhydrophobic surface on a silicon wafer using atmospheric RF plasma, followed by the formation of a selective hydrophilic area using an EHD jet printing system. (i) fluorescent image of polystyrene nanoparticles after deposition (scale bar: 50 μm) [106]. Copyright © 2017, Elsevier.
Polymers 16 02613 g003

2.4. Evaporation Assistant

Organic semiconductors can be patterned by controlling the flow within evaporating droplets to regulate pattern formation. The evaporation-assisted technique involves mixing a polymer with a solvent that has high solubility for the polymer and a nonsolvent that does not dissolve the polymer. Due to the difference in evaporation rates, the nonsolvent remains, causing the polymer to precipitate and, as the nonsolvent condenses and evaporates, a honeycomb pattern is formed. In addition, they produce concentric ring patterns of organic nanowires by controlling the evaporation of dye solution droplets [113,114,115]. For example, Cheng et al. suggested a large-scale and low-cost method to fabricate ordered films with honeycomb patterns using common commercialized polymers, such as polystyrene (PS), polysulfone (PSF), and poly(ether sulfone) (PES), under normal indoor conditions [116]. As the solvent evaporates, the nonsolvent forms droplet templates on the film surface, resulting in large-scale honeycomb-patterned films. Figure 4a shows the process and mechanism of surface pattern formation. A solvent/nonsolvent mixture was prepared using methylene chloride (solvent) and propylene glycol (nonsolvent) at various ratios, to which a specified amount of polymer was added and dissolved to obtain a casting film solution. Methylene chloride has high volatility, while propylene glycol has low volatility. These characteristics can lead to phase separation in solution. Phase separation occurs as the polymer solution is cast onto the glass substrate and the solvent evaporates, resulting in the nucleation and growth of nonsolvent droplets on the film surface. Due to its surface tension, the nonsolvent propylene glycol forms uniform droplet templates on the polymer surface. As the nonsolvent droplets are removed through continuous drying, a honeycomb pattern emerges on the film surface. This method enabled the large-scale fabrication of honeycomb pattern films (Figure 4b). Figure 4c shows the honeycomb structure across the entire surface of the polymer film in scanning electron microscopy images. Furthermore, pattern size was controlled the using various commercial polymers such as PS, PSF, and PES (Figure 4d). This patterning technique allows the production of honeycomb-shaped films in typical ambient conditions without stabilizers or humidity control. This technique also enables the modification of the surface patterns based on the polymer materials and nonsolvent content.
As another example, a novel approach to preparing regular concentric rings of aligned organic nanowires by simply evaporating the solvent in constrained geometries was demonstrated by Wang et al. [117]. The density, length, and periodicity of the nanowire arrays can be adjusted by controlling the evaporation rate. Adjusting the initial concentration allows control over the density and spacing of the concentric arrangement of organic nanowires. This easy technique can also fabricate large-scale organic semiconductor devices with nanowire configurations. The evaporation-induced self-assembly (EISA) method relies on the fact that a drop of colloidal solution always leaves ring-like deposits around its perimeter [115,118,119,120,121]. For example, Lin et al. designed a simple method to control droplet evaporation in constrained geometries, leaving well-organized gradient concentric ring patterns [122,123]. Integrating the EISA method with concentric ring patterning achieved the simultaneous self-assembly, alignment, and patterning of organic semiconductor nanowires in a single step. This study revealed the straightforward approach to fabricating large-scale concentric arrays of nanowires by solvent evaporation in constrained geometries. N,N′-dimethylquinacridone (DMQA) was chosen as the nonvolatile solute, and concentric ring patterns of DMQA nanowires were prepared from chloroform solutions at concentrations of 0.2, 0.1, and 0.05 mmol·L−1. Approximately 150 μL of a DMQA solution was dropped into the gap between the slides and a lens, with a nitrogen gas stream aiding solvent removal. The evaporation process was conducted at room temperature in a fume hood to control the solvent evaporation direction and rate (Figure 4e). After drying the solution on the substrate, concentric rings with aligned nanowires formed on both surfaces (Figure 4f). Optical microscopy and scanning electron microscopy (Figure 4g,h) revealed the prepared concentric rings at 0.10 mmol·L−1, with remarkable regularity over large areas spanning hundreds of micrometers. Each ring consisted of numerous DMQA nanowires. This easy evaporation approach was applied to fabricate well-aligned organic nanowires from homogenous solutions, significantly reducing material consumption and production costs for nanoscale devices. Furthermore, well-organized concentric nanowire rings can form from several other organic compounds, such as 2,4-Bis[4-(N,N-dimethylamino)phenyl] squaraine (SQ) and perylene tetracarboxylic diimide (PTCDI) through evaporation [124]. The self-organized patterns of molecular OSCs over large areas have potential applications in various optoelectronic devices, including OFETs and biosensors.

2.5. Diffusion

Diffusion-based organic semiconductor patterning techniques utilize the selective diffusion properties of organic semiconductor materials in solvents. The diffusion of polymers on solid surfaces has broad implications in various chemical processes, including surface-based biosensors [125,126,127], polymer separations [128], lubrication [129,130,131,132], heterogeneous catalysis [133,134,135], and surface-mediated supramolecular self-assembly [136]. Patterns are formed in organic semiconductors using masks or templates to direct the solvent to specific areas of the organic semiconductor, inducing selective diffusion. The organic semiconductor is patterned only in specific regions by diffusing the solvent along the mask pattern into the organic semiconductor material. High-resolution patterns can be achieved by controlling the diffusion rate and time. This method is suitable for the simple and rapid micro-patterning of organic semiconductors. Wang et al. used polystyrene-b-poly(methyl methacrylate) (PS-b-PMMA) thin films as templates to fabricate an array of silica nanopillars, illustrating the manufacturing process schematically (Figure 5a) [137]. The PS-b-PMMA solution was spin-coated onto a fused-silica surface functionalized with a random PS-r-PMMA copolymer with a PS:PMMA monomer ratio of 64:36. A thin film of block copolymers was self-assembled through thermal annealing, exhibiting PMMA cylindrical microdomains arranged vertically on the surface. Subsequently, the PMMA blocks were removed by UV irradiation and acetic acid washing (nonsolvent for PS but a good solvent for PMMA). The remaining PS template was further treated with oxygen plasma to remove any residual PMMA. A linear PDMS solution was spin-coated onto these PS template films and annealed at 80 °C under vacuum to allow PDMS diffusion into the nanopores. The oxygen plasma treatment decomposed the organic material on the surface, and the PDMS in the nanopores was transformed into silica. Additional etching with CF4 plasma was conducted to remove the remaining PDMS. Figure 5b presents representative images of the patterned surface observed by scanning electron microscopy. The columnar pattern surface influenced polymer motion and lateral diffusion in various ways.
For the fine patterning of OSCs, in 2020, Perevedentsev et al. proposed a versatile approach using the local diffusion of functional small molecules through solution-processed “molecular gate” layers [138]. The molecular gate controls the diffusion of functional small molecules into the target film under regulated stimuli. Functional small molecules diffuse using membranes that control permeability, leading to pattern formation. Poly(sodium 4-styrenesulfonate) (pNaSS) was used as the molecular gate material (Figure 5c), with individual processing steps shown in Figure 5d,e. The poly(9,9-dioctylfluorene) (PFO) film spin-coated on a glass substrate exhibits a deep blue PL under UV illumination (Figure 5f). After spin-coating pNaSS in an aqueous solution, a uniform film was formed, with the photoluminescence (PL) characteristics of PFO unchanged. The presence of the pNaSS molecular gate protects the PFO film during the spin-coating of linoleic acid (LA), used as a functional ”solid solvent”. The controlled diffusion of LA is activated by heating it slightly above its melting point (Tm = 44 °C) or by exposure to solvent vapor. This leads to LA diffusion into the PFO film, locally modifying the PFO chain structure and inducing β-phase formation and subsequent PL changes. Although the glassy PFO emits dark blue fluorescence, the β-phase PFO emitted a brighter blue fluorescence. Exposing the solvent vapor through a nozzle induces localized changes in the chain structure through LA–polymer interactions. The residual LA donor layers and pNaSS gates were removed by spin-off using solvents such as acetone and water. The uniform PL emission color from the sample regions where LA diffused through the gate contrasts with the less uniform color where LA was deposited directly onto PFO, indicating that gate diffusion is a highly controllable process. Laser patterning of the β-phase chains in PFO film was performed in the configuration shown in Figure 5g. Heating an indium tin oxide substrate with a laser controls LA diffusion into the PFO film, with the diffusion being controlled by the position and parameters of the laser. Patterned β-phase lines were produced by scanning the sample with constant writing speed v and varying power P at successive laser focal planes. PL microscopy images showed that laser power, which controls local temperature rise and consequent functional small molecule diffusion, is a key parameter in patterning features (Figure 5h). Molecular gate-based patterning, maintaining film planarity by donor integration, facilitates direct integration into multilayer structures. It offers the spatial resolution of photolithography and the versatility of printing techniques. Various semiconductor polymers can be spatially patterned by selecting appropriate donor compounds, chain structure and orientation, and material composition.
In this review, we revisited various OSC patterning techniques through the control of chemical and physical energy between the substrate and organic semiconductor materials. Table 1 provides a detailed summary of the OSC patterning techniques in terms of patterning methods, organic semiconductor materials, pattern resolution, and electronic device types and performance parameters.

3. Conclusions and Outlook

The development of advanced patterning techniques for OSCs has led to significant progress in organic electronics, reflecting a growing understanding of material behaviors and processing methods. Various approaches to patterning OSCs, including surface-grafting polymers, capillary force lithography, wettability, evaporation assistant, and diffusion-based methods, have collectively enhanced the capability to produce high-resolution, high-performance organic electronic devices by controlling the physical and chemical energies. These techniques have demonstrated their ability to control the morphology and alignment of OSCs, improving the performance and reliability of the devices.
Surface-grafting polymers have enabled intricate patterning and enhanced functionality of organic materials. Capillary force lithography has provided a means for high-resolution patterning and single-crystal formation, contributing to the fabrication of advanced organic thin film transistors. Wettability has proven effective in controlling the deposition and organization of organic semiconductor films, leading to uniform, high-performance devices. Evaporation-assisted methods have facilitated the production of highly ordered surface patterns, while diffusion-based techniques have offered a simple and efficient approach for micro-patterning and the integration of functional molecules.
In particular, a delicate process technology capable of patterning OSCs without physical and chemical damage is essential, and efforts are needed to achieve this. More specifically, the following aspects should be referenced and developed: (1) whether OSCs can be patterned independently of the underlying substrate; (2) the feature size of each patterning technology; (3) whether the process is solvent-free or etchant-free; (4) whether large-area patterning is possible (up to wafer size limits); (5) the cost of the patterning process; and (6) whether the chemical bonds or composition of OSCs remain unchanged before and after patterning. By consistently checking these criteria and dedicating efforts to developing OSC patterning technology, it is expected that polymer-based unit electronic devices can further advance into integrated systems.
Overall, the integration and optimization of these patterning methods have significantly advanced the field of organic electronics, paving the way for developing more efficient, scalable, and versatile organic semiconductor devices. The continued development of these technologies is expected to accelerate advances in organic electronics, offering new possibilities across a wide range of applications, from flexible electronic devices to next-generation semiconductor technologies. Furthermore, the combination of advantages such as cost-effectiveness, flexibility, deformability, and light weight in organic semiconductors, along with improved crystallinity achieved through the reviewed patterning techniques, is expected to provide significant competitiveness in the organic semiconductor industry [4]. Finally, with growing environmental concerns, the introduction of biodegradable materials, such as polyvinyl alcohol, will open up opportunities for novel patterning techniques that generate less waste and pollution.

Author Contributions

Conceptualization, H.Y.; formal analysis, W.C. and Y.E.K.; investigation, W.C. and Y.E.K.; resources, H.Y.; writing—original draft preparation, W.C. and Y.E.K.; writing—review and editing, H.Y.; supervision, H.Y.; funding acquisition, H.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by a 2023 research grant from Gachon University (GCU-202400670001), and the National Research Foundation of Korea (NRF), funded by the Korean Government (MSIT) under Grant (RS-2024-00433166). This work was funded in part by the Technology Innovation Program (1415181712, RS-2022-00144300, Interface Technology of 3D Stacked Heterogeneous System for SCM-based Process-in-Memory) funded by the Ministry of Trade, Industry and Energy (MOTIE, Republic of Korea).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Ahmad, S. Organic semiconductors for device applications: Current trends and future prospects. J. Polym. Eng. 2014, 34, 279–338. [Google Scholar] [CrossRef]
  2. Franchi, D.; Amara, Z. Applications of sensitized semiconductors as heterogeneous visible-light photocatalysts in organic synthesis. ACS Sustain. Chem. Eng. 2020, 8, 15405–15429. [Google Scholar] [CrossRef]
  3. Ruiz, C.; García-Frutos, E.M.; Hennrich, G.; Gómez-Lor, B. Organic semiconductors toward electronic devices: High mobility and easy processability. J. Phys. Chem. Lett. 2012, 3, 1428–1436. [Google Scholar] [CrossRef] [PubMed]
  4. Wang, C.; Dong, H.; Jiang, L.; Hu, W. Organic semiconductor crystals. Chem. Soc. Rev. 2018, 47, 422–500. [Google Scholar] [CrossRef]
  5. Wang, S.; Peng, L.; Sun, H.; Huang, W. The future of solution processing toward organic semiconductor devices: A substrate and integration perspective. J. Mater. Chem. C 2022, 10, 12468–12486. [Google Scholar] [CrossRef]
  6. Qian, Y.; Zhang, X.; Xie, L.; Qi, D.; Chandran, B.K.; Chen, X.; Huang, W. Stretchable organic semiconductor devices. Adv. Mater. 2016, 28, 9243–9265. [Google Scholar] [CrossRef]
  7. Gui, K.; Mutkins, K.; Schwenn, P.E.; Krueger, K.B.; Pivrikas, A.; Wolfer, P.; Stutzmann, N.S.; Burn, P.L.; Meredith, P. A flexible n-type organic semiconductor for optoelectronics. J. Mater. Chem. 2012, 22, 1800–1806. [Google Scholar] [CrossRef]
  8. Gupta, S.K.; Jha, P.; Singh, A.; Chehimi, M.M.; Aswal, D.K. Flexible organic semiconductor thin films. J. Mater. Chem. C 2015, 3, 8468–8479. [Google Scholar] [CrossRef]
  9. Zschieschang, U.; Ante, F.; Yamamoto, T.; Takimiya, K.; Kuwabara, H.; Ikeda, M.; Sekitani, T.; Someya, T.; Kern, K.; Klauk, H. Flexible low-voltage organic transistors and circuits based on a high-mobility organic semiconductor with good air stability. Adv. Mater. 2010, 22, 982–985. [Google Scholar] [CrossRef]
  10. Bi, P.; Ren, J.; Zhang, S.; Wang, J.; Hou, J. PTV-based p-type organic semiconductors: Candidates for low-cost photovoltaic donors with simple synthetic routes. Polymer 2020, 209, 122900. [Google Scholar] [CrossRef]
  11. Duan, S.; Gao, X.; Wang, Y.; Yang, F.; Chen, M.; Zhang, X.; Ren, X.; Hu, W. Scalable fabrication of highly crystalline organic semiconductor thin film by channel-restricted screen printing toward the low-cost fabrication of high-performance transistor arrays. Adv. Mater. 2019, 31, e1807975. [Google Scholar] [CrossRef] [PubMed]
  12. Li, X.; Pan, F.; Sun, C.; Zhang, M.; Wang, Z.; Du, J.; Wang, J.; Xiao, M.; Xue, L.; Zhang, Z.-G.; et al. Simplified synthetic routes for low cost and high photovoltaic performance n-type organic semiconductor acceptors. Nat. Commun. 2019, 10, 519. [Google Scholar] [CrossRef] [PubMed]
  13. 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] [PubMed]
  14. Patel, N.K.; Cina, S.; Burroughes, J.H. High-efficiency organic light-emitting diodes. IEEE J. Sel. Top. Quantum Electron. 2002, 8, 346–361. [Google Scholar] [CrossRef]
  15. Tan, X.; Dou, D.; Chua, L.-L.; Png, R.-Q.; Congrave, D.G.; Bronstein, H.; Baumgarten, M.; Li, Y.; Blom, P.W.M.; Wetzelaer, G.-J.A.H. Inverted device architecture for high efficiency single-layer organic light-emitting diodes with imbalanced charge transport. Nat. Commun. 2024, 15, 4107. [Google Scholar] [CrossRef]
  16. Zheng, D.-G.; Lee, H.-D.; Lee, G.W.; Shin, D.-S.; Kim, J.; Shim, J.-I.; Lin, Z.; Lee, T.-W.; Kim, D.H. Investigation into charge carrier dynamics in organic light-emitting diodes. Nano Res. Energy 2024, 3, e9120109. [Google Scholar] [CrossRef]
  17. Wang, Z.; Tang, Y.; Han, J.; Zhu, F. Visible-blind near-infrared organic photodetectors. Appl. Phys. Lett. 2024, 124, 021103. [Google Scholar] [CrossRef]
  18. Yang, M.; Yin, B.; Hu, G.; Cao, Y.; Lu, S.; Chen, Y.; He, Y.; Yang, X.; Huang, B.; Li, J.; et al. Sensitive short-wavelength infrared photodetection with a quinoidal ultralow band-gap n-type organic semiconductor. Chem 2024, 10, 1425–1444. [Google Scholar] [CrossRef]
  19. Yin, B.; Zhou, X.; Li, Y.; Hu, G.; Wei, W.; Yang, M.; Jeong, S.; Deng, W.; Wu, B.; Cao, Y.; et al. Sensitive Organic Photodetectors With Spectral Response up to 1.3 µm Using a Quinoidal Molecular Semiconductor. Adv. Mater. 2024, 36, e2310811. [Google Scholar] [CrossRef]
  20. Zhang, Y.; Qin, Z.; Gao, H.; Wang, T.; Gao, C.; Zhang, X.; Hu, W.; Dong, H. Highly-Polarized Solar-Blind Ultraviolet Organic Photodetectors. Adv. Mater. 2024, 36, 2404309. [Google Scholar] [CrossRef]
  21. Shekar, B.C.; Lee, J.; Rhee, S.-W. Organic thin film transistors: Materials, processes and devices. Korean J. Chem. Eng. 2004, 21, 267–285. [Google Scholar] [CrossRef]
  22. Jia, X.; Fuentes-Hernandez, C.; Wang, C.-Y.; Park, Y.; Kippelen, B. Stable organic thin-film transistors. Sci. Adv. 2018, 4, eaao1705. [Google Scholar] [CrossRef] [PubMed]
  23. Klauk, H. Organic thin-film transistors. Chem. Soc. Rev. 2010, 39, 2643–2666. [Google Scholar] [CrossRef] [PubMed]
  24. Newman, C.R.; Frisbie, C.D.; Filho, D.A.d.S.; Brédas, J.-L.; Ewbank, P.C.; Mann, K.R. Introduction to organic thin film transistors and design of n-channel organic semiconductors. Chem. Mater. 2004, 16, 4436–4451. [Google Scholar] [CrossRef]
  25. Jeon, Y.; Lee, G.; Kim, Y.J.; Jang, B.C.; Yoo, H. Dual Synapses and Security Devices from Ternary C60-Pentacene-TiO2-x Nanorods Heterostructures. Adv. Funct. Mater. 2024, 2409578. [Google Scholar] [CrossRef]
  26. Sun, J.; Fu, Y.; Wan, Q. Organic synaptic devices for neuromorphic systems. J. Phys. D Appl. Phys. 2018, 51, 314004. [Google Scholar] [CrossRef]
  27. Tuchman, Y.; Mangoma, T.N.; Gkoupidenis, P.; van de Burgt, Y.; John, R.A.; Mathews, N.; Shaheen, S.E.; Daly, R.; Malliaras, G.G.; Salleo, A. Organic neuromorphic devices: Past, present, and future challenges. MRS Bull. 2020, 45, 619–630. [Google Scholar] [CrossRef]
  28. Xie, Z.; Zhuge, C.; Zhao, Y.; Xiao, W.; Fu, Y.; Yang, D.; Zhang, S.; Li, Y.; Wang, Q.; Wang, Y.; et al. All-solid-state vertical three-terminal n-type organic synaptic devices for neuromorphic computing. Adv. Funct. Mater. 2022, 32, 2107314. [Google Scholar] [CrossRef]
  29. Kayaci, N.; Ozdemir, R.; Kalay, M.; Kiremitler, N.B.; Usta, H.; Onses, M.S. Organic light-emitting physically unclonable functions. Adv. Funct. Mater. 2022, 32, 2108675. [Google Scholar] [CrossRef]
  30. Kim, D.; Im, S.; Kim, D.; Lee, H.; Choi, C.; Cho, J.H.; Ju, H.; Lim, J.A. Reconfigurable electronic physically unclonable functions based on organic thin-film transistors with multiscale polycrystalline entropy for highly secure cryptography primitives. Adv. Funct. Mater. 2023, 33, 2210367. [Google Scholar] [CrossRef]
  31. Kuribara, K.; Watanabe, Y.; Takei, A.; Uemura, S.; Yoshida, M. Robustness of organic physically unclonable function with buskeeper circuit for flexible security devices. Jpn. J. Appl. Phys. 2022, 61, SE1016. [Google Scholar] [CrossRef]
  32. Lee, S.; Jang, B.C.; Kim, M.; Lim, S.H.; Ko, E.; Kim, H.H.; Yoo, H. Machine Learning Attacks-Resistant Security by Mixed-Assembled Layers-Inserted Graphene Physically Unclonable Function. Adv. Sci. 2023, 10, 2302604. [Google Scholar] [CrossRef] [PubMed]
  33. Shin, J.; Ko, R.; Park, T.; Kim, Y.J.; Jang, B.C.; Yoo, H. Multi-Dimensional Physically Unclonable Functions: Optoelectronic Variation-Induced Multi-Key Generation from Small Molecule PN Heterostructures. Adv. Funct. Mater. 2024, 34, 2314949. [Google Scholar] [CrossRef]
  34. Erokhin, V.; Howard, G.D.; Adamatzky, A. Organic memristor devices for logic elements with memory. Int. J. Bifurc. Chaos 2012, 22, 1250283. [Google Scholar] [CrossRef]
  35. Panigrahi, D.; Hayakawa, R.; Honma, K.; Kanai, K.; Wakayama, Y. Organic heterojunction transistors for mechanically flexible multivalued logic circuits. Appl. Phys. Express 2021, 14, 081004. [Google Scholar] [CrossRef]
  36. Park, T.; Han, Y.; Lee, S.; Kim, Y.H.; Yoo, H. Wavelength-Dependent Multistate Programmability and Optoelectronic Log-ic-in-Memory Operation from the Narrow Bandgap pNDI-SVS Floating Gate. Nano Lett. 2024, 24, 9544–9552. [Google Scholar] [CrossRef]
  37. Di, C.A.; Zhang, F.; Zhu, D. Multi-functional integration of organic field-effect transistors (OFETs): Advances and perspectives. Adv. Mater. 2013, 25, 313–330. [Google Scholar] [CrossRef]
  38. Song, M.; Seo, J.; Kim, H.; Kim, Y. Ultrasensitive multi-functional flexible sensors based on organic field-effect transistors with polymer-dispersed liquid crystal sensing layers. Sci. Rep. 2017, 7, 2630. [Google Scholar] [CrossRef]
  39. Zhang, Y.; Wang, Y.; Gao, C.; Ni, Z.; Zhang, X.; Hu, W.; Dong, H. Recent advances in n-type and ambipolar organic semi-conductors and their multi-functional applications. Chem. Soc. Rev. 2023, 52, 1331–1381. [Google Scholar] [CrossRef]
  40. Zhang, Y.; Yan, B. A 2-in-1 multi-functional sensor for efficient epinephrine detection based on a cucurbit [7] uril functionalized lanthanide metal–organic framework and its intelligent application in a molecular robot. J. Mater. Chem. C 2022, 10, 9326–9333. [Google Scholar] [CrossRef]
  41. Diao, Y.; Shaw, L.; Bao, Z.; Mannsfeld, S.C.B. Morphology control strategies for solution-processed organic semiconductor thin films. Energy Environ. Sci. 2014, 7, 2145–2159. [Google Scholar] [CrossRef]
  42. Kolesov, V.A.; Fuentes-Hernandez, C.; Chou, W.-F.; Aizawa, N.; Larrain, F.A.; Wang, M.; Perrotta, A.; Choi, S.; Graham, S.; Bazan, G.C.; et al. Solution-based electrical doping of semiconducting polymer films over a limited depth. Nat. Mater. 2017, 16, 474–480. [Google Scholar] [CrossRef] [PubMed]
  43. Qi, Y.; Mohapatra, S.K.; Kim, S.B.; Barlow, S.; Marder, S.R.; Kahn, A. Solution doping of organic semiconductors using air-stable n-dopants. Appl. Phys. Lett. 2012, 100, 083305. [Google Scholar] [CrossRef]
  44. Shaw, L.; Bao, Z. The Large-Area, Solution-Based Deposition of Single-Crystal Organic Semiconductors. Isr. J. Chem. 2014, 54, 496–512. [Google Scholar] [CrossRef]
  45. Azarova, N.A.; Owen, J.W.; McLellan, C.A.; Grimminger, M.A.; Chapman, E.K.; Anthony, J.E.; Jurchescu, O.D. Fabrication of organic thin-film transistors by spray-deposition for low-cost, large-area electronics. Org. Electron. 2010, 11, 1960–1965. [Google Scholar] [CrossRef]
  46. Dimitrakopoulos, C.D.; Malenfant, P.R. Organic thin film transistors for large area electronics. Adv. Mater. 2002, 14, 99–117. [Google Scholar] [CrossRef]
  47. Søndergaard, R.R.; Hösel, M.; Krebs, F.C. Roll-to-Roll fabrication of large area functional organic materials. J. Polym. Sci. Part B Polym. Phys. 2013, 51, 16–34. [Google Scholar] [CrossRef]
  48. Zhang, X.; Jie, J.; Deng, W.; Shang, Q.; Wang, J.; Wang, H.; Chen, X.; Zhang, X. Alignment and patterning of ordered small-molecule organic semiconductor micro-/nanocrystals for device applications. Adv. Mater. 2016, 28, 2475–2503. [Google Scholar] [CrossRef]
  49. Ma, Z.; Zhao, B.; Gao, H.; Gong, Y.; Yu, R.; Tan, Z. Recent advances of crosslinkable organic semiconductors in achieving solution-processed and stable optoelectronic devices. J. Mater. Chem. A 2022, 10, 18542–18576. [Google Scholar] [CrossRef]
  50. Zhao, X.; Wang, S.; Ni, Y.; Tong, Y.; Tang, Q.; Liu, Y. High-performance full-photolithographic top-contact conformable organic transistors for soft electronics. Adv. Sci. 2021, 8, 2004050. [Google Scholar] [CrossRef]
  51. Fang, X.; Shi, J.; Zhang, X.; Ren, X.; Lu, B.; Deng, W.; Jie, J.; Zhang, X. Patterning liquid crystalline organic semiconductors via inkjet printing for high-performance transistor arrays and circuits. Adv. Funct. Mater. 2021, 31, 2100237. [Google Scholar] [CrossRef]
  52. Jiang, X.; Jiang, L. One-dimensional arrays of sensing materials based on wettability interface dewetting process. Acc. Mater. Res. 2020, 1, 53–62. [Google Scholar] [CrossRef]
  53. Kwon, H.-J.; Li, X.; Hong, J.; Park, C.E.; Jeong, Y.J.; Moon, H.C.; Kim, S.H. Non-lithographic direct patterning of carbon nanomaterial electrodes via electrohydrodynamic-printed wettability patterns by polymer brush for fabrication of organic field-effect transistor. Appl. Surf. Sci. 2020, 515, 145989. [Google Scholar] [CrossRef]
  54. Narupai, B.; Page, Z.A.; Treat, N.J.; McGrath, A.J.; Pester, C.W.; Discekici, E.H.; Dolinski, N.D.; Meyers, G.F.; de Alaniz, J.R.; Hawker, C.J. Simultaneous preparation of multiple polymer brushes under ambient conditions using microliter volumes. Angew. Chem. Int. Ed. 2018, 57, 13433–13438. [Google Scholar] [CrossRef]
  55. Deng, W.; Lv, Y.; Zhang, X.; Fang, X.; Lu, B.; Lu, Z.; Jie, J. High-resolution patterning of organic semiconductor single crystal arrays for high-integration organic field-effect transistors. Mater. Today Proc. 2020, 40, 82–90. [Google Scholar] [CrossRef]
  56. Fu, B.; Yang, F.; Sun, L.; Zhao, Q.; Ji, D.; Sun, Y.; Zhang, X.; Hu, W. Challenging bendable organic single crystal and transistor arrays with high mobility and durability toward flexible electronics. Adv. Mater. 2022, 34, e2203330. [Google Scholar] [CrossRef]
  57. Ge, F.; Wang, X.; Zhang, Y.; Song, E.; Zhang, G.; Lu, H.; Cho, K.; Qiu, L. Modulating the surface via polymer brush for high-performance inkjet-printed organic thin-film transistors. Adv. Electron. Mater. 2017, 3, 1600402. [Google Scholar] [CrossRef]
  58. 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]
  59. Yuan, J.; Liu, D.; Zhao, H.; Lin, B.; Zhou, X.; Naveed, H.B.; Zhao, C.; Zhou, K.; Tang, Z.; Chen, F.; et al. Patterned blade coating strategy enables the enhanced device reproducibility and optimized morphology of organic solar cells. Adv. Energy Mater. 2021, 11, 2100098. [Google Scholar] [CrossRef]
  60. 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]
  61. He, Z.; Zhang, Z.; Asare-Yeboah, K.; Bi, S.; Chen, J.; Li, D. Crystal growth of small-molecule organic semiconductors with nucleation additive. Curr. Appl. Phys. 2021, 21, 107–115. [Google Scholar] [CrossRef]
  62. Higashino, T.; Inoue, S.; Arai, S.; Matsui, H.; Toda, N.; Horiuchi, S.; Azumi, R.; Hasegawa, T. Architecting layered crystalline organic semiconductors based on unsymmetric π-extended thienoacenes. Chem. Mater. 2021, 33, 7379–7385. [Google Scholar] [CrossRef]
  63. Horowitz, G.; Hajlaoui, M.E. Mobility in polycrystalline oligothiophene field-effect transistors dependent on grain size. Adv. Mater. 2000, 12, 1046–1050. [Google Scholar] [CrossRef]
  64. Jurchescu, O.D.; Baas, J.; Palstra, T.T.M. Effect of impurities on the mobility of single crystal pentacene. Appl. Phys. Lett. 2004, 84, 3061–3063. [Google Scholar] [CrossRef]
  65. Minari, T.; Nemoto, T.; Isoda, S. Fabrication and characterization of single-grain organic field-effect transistor of pentacene. J. Appl. Phys. 2004, 96, 769–772. [Google Scholar] [CrossRef]
  66. Belmonte, A.; Pilz da Cunha, M.; Nickmans, K.; Schenning, A.P. Brush-paintable, temperature and light responsive triple shape-memory photonic coatings based on micrometer-sized cholesteric liquid crystal polymer particles. Adv. Opt. Mater. 2020, 8, 2000054. [Google Scholar] [CrossRef]
  67. Cavallini, M.; Melucci, M. Organic Materials for Time–Temperature Integrator Devices. ACS Appl. Mater. Interfaces 2015, 7, 16897–16906. [Google Scholar] [CrossRef]
  68. Gentili, D.; Durso, M.; Bettini, C.; Manet, I.; Gazzano, M.; Capelli, R.; Muccini, M.; Melucci, M.; Cavallini, M. A time-temperature integrator based on fluorescent and polymorphic compounds. Sci. Rep. 2013, 3, 2581. [Google Scholar] [CrossRef]
  69. Kim, H.; Choi, S.-K.; Ahn, J.; Yu, H.; Min, K.; Hong, C.; Shin, I.-S.; Lee, S.; Lee, H.; Im, H.; et al. Kaleidoscopic fluorescent arrays for machine-learning-based point-of-care chemical sensing. Sens. Actuators B Chem. 2021, 329, 129248. [Google Scholar] [CrossRef]
  70. Kim, J.-M.; Lee, Y.B.; Yang, D.H.; Lee, J.-S.; Lee, G.S.; Ahn, D.J. A polydiacetylene-based fluorescent sensor chip. J. Am. Chem. Soc. 2005, 127, 17580–17581. [Google Scholar] [CrossRef]
  71. Melucci, M.; Zambianchi, M.; Favaretto, L.; Palermo, V.; Treossi, E.; Montalti, M.; Bonacchi, S.; Cavallini, M. Multicolor, large-area fluorescence sensing through oligothiophene-self-assembled monolayers. Chem. Commun. 2011, 47, 1689–1691. [Google Scholar] [CrossRef] [PubMed]
  72. Michalek, L.; Barner, L.; Barner-Kowollik, C. Polymer on top: Current limits and future perspectives of quantitatively evaluating surface grafting. Adv. Mater. 2018, 30, e1706321. [Google Scholar] [CrossRef] [PubMed]
  73. Michalek, L.; Mundsinger, K.; Barner-Kowollik, C.; Barner, L. The long and the short of polymer grafting. Polym. Chem. 2019, 10, 54–59. [Google Scholar] [CrossRef]
  74. Roy, D.; Semsarilar, M.; Guthrie, J.T.; Perrier, S. Cellulose modification by polymer grafting: A review. Chem. Soc. Rev. 2009, 38, 2046–2064. [Google Scholar] [CrossRef] [PubMed]
  75. Vega-Hernández, M.; Cano-Díaz, G.S.; Vivaldo-Lima, E.; Rosas-Aburto, A.; Hernández-Luna, M.G.; Martinez, A.; Palacios-Alquisira, J.; Mohammadi, Y.; Penlidis, A. A review on the synthesis, characterization, and modeling of polymer grafting. Processes 2021, 9, 375. [Google Scholar] [CrossRef]
  76. Li, Y.; Zuilhof, H. Photochemical grafting and patterning of organic monolayers on indium tin oxide substrates. Langmuir 2012, 28, 5350–5359. [Google Scholar] [CrossRef]
  77. Ebata, K.; Furukawa, K.; Matsumoto, N. Synthesis and characterization of end-grafted polysilane on a substrate surface. J. Am. Chem. Soc. 1998, 120, 7367–7368. [Google Scholar] [CrossRef]
  78. Koutsos, V.; van der Vegte, E.W.; Pelletier, E.; Stamouli, A.; Hadziioannou, G. Structure of chemically end-grafted polymer chains studied by scanning force microscopy in bad-solvent conditions. Macromolecules 1997, 30, 4719–4726. [Google Scholar] [CrossRef]
  79. Ito, Y.; Ochiai, Y.; Park, Y.S.; Imanishi, Y. pH-sensitive gating by conformational change of a polypeptide brush grafted onto a porous polymer membrane. J. Am. Chem. Soc. 1997, 119, 1619–1623. [Google Scholar] [CrossRef]
  80. Jordan, R.; Ulman, A. Surface initiated living cationic polymerization of 2-oxazolines. J. Am. Chem. Soc. 1998, 120, 243–247. [Google Scholar] [CrossRef]
  81. Feng, J.; Jiang, X.; Yan, X.; Wu, Y.; Su, B.; Fu, H.; Yao, J.; Jiang, L. “Capillary-Bridge Lithography” for Patterning Organic Crystals toward Mode-Tunable Microlaser Arrays. Adv. Mater. 2017, 29, 1603652. [Google Scholar] [CrossRef] [PubMed]
  82. Jo, P.S.; Vailionis, A.; Park, Y.M.; Salleo, A. Scalable fabrication of strongly textured organic semiconductor micropatterns by capillary force lithography. Adv. Mater. 2012, 24, 3269–3274. [Google Scholar] [CrossRef] [PubMed]
  83. Kwon, H.-J.; Kim, K.; An, T.K.; Kim, S.H.; Park, C.E. Effect of lateral confinement on crystallization behavior of a small-molecule semiconductor during capillary force lithography for use in high-performance OFETs. J. Ind. Eng. Chem. 2019, 75, 187–193. [Google Scholar] [CrossRef]
  84. Park, Y.; Park, J.; Cho, S.; Sung, M.M. Large-area single-crystal organic patterned thin films by vertically confined lateral crystal growth via capillary force lithography. Appl. Surf. Sci. 2019, 494, 1023–1029. [Google Scholar] [CrossRef]
  85. Ho, D.; Zou, J.; Zdyrko, B.; Iyer, K.S.; Luzinov, I. Capillary force lithography: The versatility of this facile approach in de-veloping nanoscale applications. Nanoscale 2015, 7, 401–414. [Google Scholar] [CrossRef]
  86. Cavallini, M.; Albonetti, C.; Biscarini, F. Nanopatterning soluble multifunctional materials by unconventional wet lithography. Adv. Mater. 2009, 21, 1043–1053. [Google Scholar] [CrossRef]
  87. Cavallini, M.; Biscarini, F. Nanostructuring conjugated materials by lithographically controlled wetting. Nano Lett. 2003, 3, 1269–1271. [Google Scholar] [CrossRef]
  88. Cavallini, M.; Gentili, D.; Greco, P.; Valle, F.; Biscarini, F. Micro- and nanopatterning by lithographically controlled wetting. Nat. Protoc. 2012, 7, 1668–1676. [Google Scholar] [CrossRef]
  89. Huang, S.; Song, J.; Lu, Y.; Chen, F.; Zheng, H.; Yang, X.; Liu, X.; Sun, J.; Carmalt, C.J.; Parkin, I.P.; et al. Underwater spontaneous pumpless transportation of nonpolar organic liquids on extreme wettability patterns. ACS Appl. Mater. Interfaces 2016, 8, 2942–2949. [Google Scholar] [CrossRef]
  90. Lee, H.S.; Kwak, D.; Lee, W.H.; Cho, J.H.; Cho, K. Self-organization characteristics of soluble pentacene on wettability-controlled patterned substrate for organic field-effect transistors. J. Phys. Chem. C 2010, 114, 2329–2333. [Google Scholar] [CrossRef]
  91. Liu, S.; Wang, W.M.; Mannsfeld, S.C.B.; Locklin, J.; Erk, P.; Gomez, M.; Richter, F.; Bao, Z. Solution-assisted assembly of organic semiconducting single crystals on surfaces with patterned wettability. Langmuir 2007, 23, 7428–7432. [Google Scholar] [CrossRef] [PubMed]
  92. Svarnas, P.; Yang, L.; Munz, M.; Edwards, A.J.; Shard, A.G.; Bradley, J.W. Highly-selective wettability on organic light-emitting-diodes patterns by sequential low-power plasmas. J. Appl. Phys. 2010, 107, 103313. [Google Scholar] [CrossRef]
  93. Bae, J.; Park, S.; Jung, H.; Ko, E.-H.; Kymissis, I.; Kim, C.-H. Polymer field-effect transistors with inkjet-printed silver electrodes: From device fabrication to circuit simulation. J. Mater. Sci. Mater. Electron. 2024, 35, 185. [Google Scholar] [CrossRef]
  94. Kwak, D.; Lim, J.A.; Kang, B.; Lee, W.H.; Cho, K. Self-organization of inkjet-printed organic semiconductor films prepared in inkjet-etched microwells. Adv. Funct. Mater. 2013, 23, 5224–5231. [Google Scholar] [CrossRef]
  95. 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]
  96. Ren, X.; Qiu, F.; Deng, W.; Fang, X.; Wu, Y.; Yu, S.; Liu, X.; Grigorian, S.; Shi, J.; Jie, J.; et al. Topology-Mediated Molecule Nucleation Anchoring Enables Inkjet Printing of Organic Semiconducting Single Crystals for High-Performance Printed Electronics. ACS Nano 2023, 17, 25175–25184. [Google Scholar] [CrossRef]
  97. 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]
  98. Lee, S.B.; Kang, B.; Kim, D.; Park, C.; Kim, S.; Lee, M.; Lee, W.B.; Cho, K. Motion-programmed bar-coating method with controlled gap for high-speed scalable preparation of highly crystalline organic semiconductor thin films. ACS Appl. Mater. Interfaces 2019, 11, 47153–47161. [Google Scholar] [CrossRef]
  99. Sung, S.; Lee, W.-J.; Payne, M.M.; Anthony, J.E.; Kim, C.-H.; Yoon, M.-H. Large-area printed low-voltage organic thin film transistors via minimal-solution bar-coating. J. Mater. Chem. C 2020, 8, 15112–15118. [Google Scholar] [CrossRef]
  100. Zhang, Z.; Peng, B.; Ji, X.; Pei, K.; Chan, P.K.L. Marangoni-effect-assisted bar-coating method for high-quality organic crystals with compressive and tensile strains. Adv. Funct. Mater. 2017, 27, 1703443. [Google Scholar] [CrossRef]
  101. Liu, Z.; Li, Z.; Cai, Z.; Qiao, Y.; Yang, Y.; Chen, S.; Ma, X.; Li, H.; Meng, Q.; Zhang, F.; et al. Vapor-induced marangoni coating for organic functional films. J. Mater. Chem. C 2021, 9, 17518–17525. [Google Scholar] [CrossRef]
  102. Wang, H.; Liu, S.; Li, H.; Li, M.; Wu, X.; Zhang, S.; Ye, L.; Hu, X.; Chen, Y. Green Printing for Scalable Organic Photovoltaic Modules by Controlling the Gradient Marangoni Flow. Adv. Mater. 2024, 36, e2313098. [Google Scholar] [CrossRef] [PubMed]
  103. Zhao, H.; Wang, Z.; Dong, G.; Duan, L. Fabrication of highly oriented large-scale TIPS pentacene crystals and transistors by the Marangoni effect-controlled growth method. Phys. Chem. Chem. Phys. 2015, 17, 6274–6279. [Google Scholar] [CrossRef] [PubMed]
  104. Kang, B.; Ge, F.; Qiu, L.; Cho, K. Effective Use of Electrically Insulating Units in Organic Semiconductor Thin Films for High-Performance Organic Transistors. Adv. Electron. Mater. 2017, 3, 1600240. [Google Scholar] [CrossRef]
  105. Shin, N.; Kang, J.; Richter, L.J.; Prabhu, V.M.; Kline, R.J.; Fischer, D.A.; DeLongchamp, D.M.; Toney, M.F.; Satija, S.K.; Gundlach, D.J.; et al. Vertically Segregated Structure and Properties of Small Molecule–Polymer Blend Semiconductors for Organic Thin-Film Transistors. Adv. Funct. Mater. 2013, 23, 366–376. [Google Scholar] [CrossRef]
  106. Lee, J.; Hwang, S.; Cho, D.-H.; Hong, J.; Shin, J.H.; Byun, D. RF plasma based selective modification of hydrophilic regions on super hydrophobic surface. Appl. Surf. Sci. 2017, 394, 543–553. [Google Scholar] [CrossRef]
  107. Goessl, A.; Garrison, M.D.; Lhoest, J.-B.; Hoffman, A.S. Plasma lithography—Thin-film patterning of polymeric biomaterials by RF plasma polymerization I: Surface preparation and analysis. J. Biomater. Sci. Polym. Ed. 2001, 12, 721–738. [Google Scholar] [CrossRef]
  108. Steudel, S.; Myny, K.; De Vusser, S.; Genoe, J.; Heremans, P. Patterning of organic thin film transistors by oxygen plasma etch. Appl. Phys. Lett. 2006, 89, 183503. [Google Scholar] [CrossRef]
  109. Visconti, P.; Turco, C.; Rinaldi, R.; Cingolani, R. Nanopatterning of organic and inorganic materials by holographic li-thography and plasma etching. Microelectron. Eng. 2000, 53, 391–394. [Google Scholar] [CrossRef]
  110. Ali, S.; Bae, J.; Lee, C.H. Organic diode with high rectification ratio made of electrohydrodynamic printed organic layers. Electron. Mater. Lett. 2016, 12, 270–275. [Google Scholar] [CrossRef]
  111. Li, X.; Go, M.; Lim, S.; An, T.K.; Jeong, Y.J.; Kim, S.H. Electrohydrodynamic (EHD) jet printing of carbon-black composites for solution-processed organic field-effect transistors. Org. Electron. 2019, 73, 279–285. [Google Scholar] [CrossRef]
  112. Sun, Z.; Li, B.; Zhou, C.; Luo, K.; Wu, J. Organic phase change materials melting with electrohydrodynamics (EHD): Numerical simulation and experimental validation. Int. J. Heat Mass Transf. 2023, 217, 124646. [Google Scholar] [CrossRef]
  113. Fukuhira, Y.; Yabu, H.; Ijiro, K.; Shimomura, M. Interfacial tension governs the formation of self-organized honey-comb-patterned polymer films. Soft Matter 2009, 5, 2037–2041. [Google Scholar] [CrossRef]
  114. Li, L.; Zhong, Y.; Ma, C.; Li, J.; Chen, C.; Zhang, A.; Tang, D.; Xie, S.; Ma, Z. Honeycomb-patterned hybrid films and their template applications via a tunable amphiphilic block polymer/inorganic precursor system. Chem. Mater. 2009, 21, 4977–4983. [Google Scholar] [CrossRef]
  115. Yabu, H.; Tanaka, M.; Ijiro, K.; Shimomura, M. Preparation of honeycomb-patterned polyimide films by self-organization. Langmuir 2003, 19, 6297–6300. [Google Scholar] [CrossRef]
  116. Cheng, Q.; Lin, L.; Deng, X.; Zheng, T.; Wang, Q.; Gao, Y.; Zhai, X.; Yang, J.; Ma, W.; Li, X.; et al. Large-scale and low-cost preparation of ordered honeycomb-patterned film by solvent evaporation-induced phase separation method. Ind. Eng. Chem. Res. 2021, 60, 898–907. [Google Scholar] [CrossRef]
  117. Wang, Z.; Bao, R.; Zhang, X.; Ou, X.; Lee, C.S.; Chang, J.C.; Zhang, X. One-Step Self-Assembly, Alignment, and Patterning of Organic Semiconductor Nanowires by Controlled Evaporation of Confined Microfluids. Angew. Chem. Int. Ed. 2011, 12, 2811–2815. [Google Scholar] [CrossRef]
  118. Abdul Wahab, M.; Sellinger, A. Photoluminescent surfactants: Mesostructured organosilica nanocomposites via evaporation-induced self assembly (EISA). Chem. Lett. 2006, 35, 1240–1241. [Google Scholar] [CrossRef]
  119. Chou, Y.-H.; Lee, W.-Y.; Chen, W.-C. Self-assembled nanowires of organic n-type semiconductor for nonvolatile transistor memory devices. Adv. Funct. Mater. 2012, 22, 4352–4359. [Google Scholar] [CrossRef]
  120. Fan, X.; Yu, T.; Wang, Y.; Zheng, J.; Gao, L.; Li, Z.; Ye, J.; Zou, Z. Role of phosphorus in synthesis of phosphated mesoporous TiO2 photocatalytic materials by EISA method. Appl. Surf. Sci. 2008, 254, 5191–5198. [Google Scholar] [CrossRef]
  121. Ki, H.S.; Yeum, J.H.; Choe, S.; Kim, J.H.; Cheong, I.W. Fabrication of transparent conductive carbon nano-tubes/polyurethane-urea composite films by solvent evaporation-induced self-assembly (EISA). Compos. Sci. Technol. 2009, 69, 645–650. [Google Scholar] [CrossRef]
  122. Hong, S.W.; Byun, M.; Lin, Z. Robust Self-Assembly of Highly Ordered Complex Structures by Controlled Evaporation of Confined Microfluids. Angew. Chem. Int. Ed. 2009, 48, 512–516. [Google Scholar] [CrossRef] [PubMed]
  123. Hong, S.W.; Xu, J.; Lin, Z. Template-assisted formation of gradient concentric gold rings. Nano Lett. 2006, 6, 2949–2954. [Google Scholar] [CrossRef] [PubMed]
  124. Balakrishnan, K.; Datar, A.; Oitker, R.; Chen, H.; Zuo, J.; Zang, L. Nanobelt self-assembly from an organic n-type semiconductor: Propoxyethyl-PTCDI. J. Am. Chem. Soc. 2005, 127, 10496–10497. [Google Scholar] [CrossRef]
  125. Abolfath-Beygi, M.; Krishnamurthy, V.; Cornell, B. Multiple surface-based biosensors for enhanced molecular detection in fluid flow systems. IEEE Sens. J. 2013, 13, 1265–1273. [Google Scholar] [CrossRef]
  126. Hansen, R.; Bruus, H.; Callisen, T.H.; Hassager, O. Transient convection, diffusion, and adsorption in surface-based biosensors. Langmuir 2012, 28, 7557–7563. [Google Scholar] [CrossRef]
  127. Malhotra, B.D.; Chaubey, A.; Singh, S. Prospects of conducting polymers in biosensors. Anal. Chim. Acta 2006, 578, 59–74. [Google Scholar] [CrossRef]
  128. Radke, W. Polymer separations by liquid interaction chromatography: Principles–prospects–limitations. J. Chromatogr. A 2014, 1335, 62–79. [Google Scholar] [CrossRef]
  129. Cohen, S.C.; Tabor, D. The friction and lubrication of polymers. Proc. R. Soc. Lond. Ser. A Math. Phys. Sci. 1966, 291, 186–207. [Google Scholar]
  130. Fusaro, R. Self-lubricating polymer composites and polymer transfer film lubrication for space applications. Tribol. Int. 1990, 23, 105–122. [Google Scholar] [CrossRef]
  131. Liu, G.; Feng, Y.; Zhao, N.; Chen, Z.; Shi, J.; Zhou, F. Polymer-based lubricating materials for functional hydration lubrication. J. Chem. Eng. 2022, 429, 132324. [Google Scholar] [CrossRef]
  132. Raviv, U.; Giasson, S.; Kampf, N.; Gohy, J.-F.; Jérôme, R.; Klein, J. Lubrication by charged polymers. Nature 2003, 425, 163–165. [Google Scholar] [CrossRef] [PubMed]
  133. Collman, J.P.; Kosydar, K.M.; Bressan, M.; Lamanna, W.; Garrett, T. Polymer-bound substrates: A method to distinguish between homogeneous and heterogeneous catalysis. J. Am. Chem. Soc. 1984, 106, 2569–2579. [Google Scholar] [CrossRef]
  134. Kaur, P.; Hupp, J.T.; Nguyen, S.T. Porous organic polymers in catalysis: Opportunities and challenges. ACS Catal. 2011, 1, 819–835. [Google Scholar] [CrossRef]
  135. Zhou, Y.; Zhan, Z. Conjugated microporous polymers for heterogeneous catalysis. Chem. Asian J. 2018, 13, 9–19. [Google Scholar] [CrossRef]
  136. Yang, B.; Adams, D.J.; Marlow, M.; Zelzer, M. Surface-mediated supramolecular self-assembly of protein, peptide, and nucleoside derivatives: From surface design to the underlying mechanism and tailored functions. Langmuir 2018, 34, 15109–15125. [Google Scholar] [CrossRef]
  137. Wang, D.; He, C.; Stoykovich, M.P.; Schwartz, D.K. Nanoscale topography influences polymer surface diffusion. ACS Nano 2015, 9, 1656–1664. [Google Scholar] [CrossRef]
  138. Perevedentsev, A.; Campoy-Quiles, M. Rapid and high-resolution patterning of microstructure and composition in organic semiconductors using ‘molecular gates’. Nat. Commun. 2020, 11, 3610. [Google Scholar] [CrossRef]
Figure 1. (a) Schematic diagram of visible light-mediated p(DMAEMA) brush growth under ambient conditions. (b) Optical image of photomask-guided sequential p(DMAEMA) growth on a wafer-based substrate patterned with underlying polymers (X: substrate, Y: patterned polymer). Atomic force microscopy topography images and height profiles for patterned p(DMAEMA), at (c) X region (substrate) and (d) Y region (patterned polymer). The p(DMAEMA) brush synthesis process under ambient conditions using M, R, and L letter glass covers: (e) initial chemically treated silicon wafer, (f) polymer brush growth process through M, R, and L letter covers, (g) p(DMAEMA) brushes patterned with the letters M, R, and L. (h) Enlarged optical image of p(DMAEMA) brushes patterned in the shape of MRL [54]. Copyright © 2018, John Wiley and Sons. (i) Schematic diagram of the PS-brush synthesis process. (j) Schematic diagram of the evaporation process for organic semiconductor crystal growth between the PS brushes and organic semiconductor ink. (k) Mechanism of organic semiconductor crystal growth depending on the chain length of PS brushes. (l) Schematic diagram of the device and transfer curve of the TIPS-pentacene-based TFT with optimized crystallinity [57]. Copyright © 2016, John Wiley and Sons.
Figure 1. (a) Schematic diagram of visible light-mediated p(DMAEMA) brush growth under ambient conditions. (b) Optical image of photomask-guided sequential p(DMAEMA) growth on a wafer-based substrate patterned with underlying polymers (X: substrate, Y: patterned polymer). Atomic force microscopy topography images and height profiles for patterned p(DMAEMA), at (c) X region (substrate) and (d) Y region (patterned polymer). The p(DMAEMA) brush synthesis process under ambient conditions using M, R, and L letter glass covers: (e) initial chemically treated silicon wafer, (f) polymer brush growth process through M, R, and L letter covers, (g) p(DMAEMA) brushes patterned with the letters M, R, and L. (h) Enlarged optical image of p(DMAEMA) brushes patterned in the shape of MRL [54]. Copyright © 2018, John Wiley and Sons. (i) Schematic diagram of the PS-brush synthesis process. (j) Schematic diagram of the evaporation process for organic semiconductor crystal growth between the PS brushes and organic semiconductor ink. (k) Mechanism of organic semiconductor crystal growth depending on the chain length of PS brushes. (l) Schematic diagram of the device and transfer curve of the TIPS-pentacene-based TFT with optimized crystallinity [57]. Copyright © 2016, John Wiley and Sons.
Polymers 16 02613 g001
Figure 2. Capillary force-based (a) schematic diagram of the C8-BTBT patterning process, (b) sequential optical microscopy images of the formation process of C8-BTBT crystals (i: solid; ii and iii: phase transition; iv: microchannel formation; v: cooling), and (c) COMSOL multiphysics simulation of liquid C8-BTBT molecule flow (i: C8-BTBT deposition; ii: C8-BTBT liquid moves along the sidewall of CYTOP; iii: C8-BTBT liquid starts to fill the bottoms; iv and v: as the C8-BTBT liquid fills the droplets form an upward concave meniscus; vi and vii: C8-BTBT liquid fills the voids due to capillary forces; viii: C8-BTBT microchannel formation). (d) Schematic diagram of a 13 × 13 array of OTFTs based on single-crystal C8-BTBT. At the unit device scale, (e,f) scanning electron microscopy images of a single-crystal C8-BTBT-based OFET and (g) transfer curve (red line: logarithmic scale; purple line: linear scale; dashed line: leakage current) [55]. Copyright © 2020, Elsevier. 1D single-crystal C8-BTBT array synthesis: (h) schematic diagram and (i) mechanism. (j) Transfer curves of OFETs based on single-crystal C8-BTBT morphology (left OFET: 1D single-crystalline belt arrays; right OFET: thin film; red line: logarithmic scale; blue line: linear scale). 1D single-crystal C8-BTBT-based pressure sensor: (k) schematic diagram, (l) photograph of the device, and (m) pressure–current response curves [60]. Copyright © 2018, John Wiley and Sons.
Figure 2. Capillary force-based (a) schematic diagram of the C8-BTBT patterning process, (b) sequential optical microscopy images of the formation process of C8-BTBT crystals (i: solid; ii and iii: phase transition; iv: microchannel formation; v: cooling), and (c) COMSOL multiphysics simulation of liquid C8-BTBT molecule flow (i: C8-BTBT deposition; ii: C8-BTBT liquid moves along the sidewall of CYTOP; iii: C8-BTBT liquid starts to fill the bottoms; iv and v: as the C8-BTBT liquid fills the droplets form an upward concave meniscus; vi and vii: C8-BTBT liquid fills the voids due to capillary forces; viii: C8-BTBT microchannel formation). (d) Schematic diagram of a 13 × 13 array of OTFTs based on single-crystal C8-BTBT. At the unit device scale, (e,f) scanning electron microscopy images of a single-crystal C8-BTBT-based OFET and (g) transfer curve (red line: logarithmic scale; purple line: linear scale; dashed line: leakage current) [55]. Copyright © 2020, Elsevier. 1D single-crystal C8-BTBT array synthesis: (h) schematic diagram and (i) mechanism. (j) Transfer curves of OFETs based on single-crystal C8-BTBT morphology (left OFET: 1D single-crystalline belt arrays; right OFET: thin film; red line: logarithmic scale; blue line: linear scale). 1D single-crystal C8-BTBT-based pressure sensor: (k) schematic diagram, (l) photograph of the device, and (m) pressure–current response curves [60]. Copyright © 2018, John Wiley and Sons.
Polymers 16 02613 g002
Figure 4. (a) Schematic diagram of the formation process of honeycomb-patterned films. (b) Large-area honeycomb film for digital photography. (c) Scanning electron microscopy image of the honeycomb-patterned film. (d) Scanning electron microscopy images of the surfaces of patterned films made from different polymers, along with the molecular structural formulas of these polymers: PS, PSF, and PES [116]. Copyright © 2021, American Chemical Society. (e) Illustration of the sphere-on-flat evaporation setup. (f) Illustration of the concentric rings of as-prepared organic nanowires. (g) Optical micrographs and (h) scanning electron microscopy images of concentric rings of DMQA nanowires formed during solvent evaporation [117]. Copyright © 2011, John Wiley and Sons.
Figure 4. (a) Schematic diagram of the formation process of honeycomb-patterned films. (b) Large-area honeycomb film for digital photography. (c) Scanning electron microscopy image of the honeycomb-patterned film. (d) Scanning electron microscopy images of the surfaces of patterned films made from different polymers, along with the molecular structural formulas of these polymers: PS, PSF, and PES [116]. Copyright © 2021, American Chemical Society. (e) Illustration of the sphere-on-flat evaporation setup. (f) Illustration of the concentric rings of as-prepared organic nanowires. (g) Optical micrographs and (h) scanning electron microscopy images of concentric rings of DMQA nanowires formed during solvent evaporation [117]. Copyright © 2011, John Wiley and Sons.
Polymers 16 02613 g004
Figure 5. (a) Schematic diagram of the preparation procedure for an array of silica pillars. (b) Representative scanning electron microscopy image of the patterned silica pillar array on the surface [137]. Copyright © 2015, American Chemical Society. (c) Chemical structures of pNaSS. Illustration of the molecular gate concept through PL switching in PFO films, depicted for (d) thermal and (e) solvent vapor-based stimuli. The top rows show photographs of the films under UV light at various processing steps, with schematic illustrations in the bottom rows. (f) Chemical structures of PFO (R = C8H17). (g) Schematic diagram of the laser-patterning process. (h) PL intensity at 438 nm showing a spectral selection of β-phase emission. In addition, a Raman intensity ratio image (right panel) of the line patterned using P = 30 mW was obtained, indicating the local β-phase fraction as estimated from the ratio of Raman intensities at 1257 and 1606 cm−1 [138]. Copyright © 2020, Springer Nature.
Figure 5. (a) Schematic diagram of the preparation procedure for an array of silica pillars. (b) Representative scanning electron microscopy image of the patterned silica pillar array on the surface [137]. Copyright © 2015, American Chemical Society. (c) Chemical structures of pNaSS. Illustration of the molecular gate concept through PL switching in PFO films, depicted for (d) thermal and (e) solvent vapor-based stimuli. The top rows show photographs of the films under UV light at various processing steps, with schematic illustrations in the bottom rows. (f) Chemical structures of PFO (R = C8H17). (g) Schematic diagram of the laser-patterning process. (h) PL intensity at 438 nm showing a spectral selection of β-phase emission. In addition, a Raman intensity ratio image (right panel) of the line patterned using P = 30 mW was obtained, indicating the local β-phase fraction as estimated from the ratio of Raman intensities at 1257 and 1606 cm−1 [138]. Copyright © 2020, Springer Nature.
Polymers 16 02613 g005
Table 1. Summarization of OSC patterning techniques.
Table 1. Summarization of OSC patterning techniques.
Patterning MethodOrganic MaterialPattern
Resolution
Device TypeDevice
Performance
[Ref.]
Surface-grafting polymersp(DMAEMA)20 μmN/AN/A[54]
Surface-grafting polymersTIPS-pentacene46.4–105 μmOTFTCarrier mobility = 1.2 cm2·V−1·s−1[57]
Capillary force lithographyC8-BTBT
1D single crystal
~3 μmOFETCarrier mobility = 5.7 cm2·V−1·s−1[55]
Capillary force lithographyC8-BTBT
1D single crystal
~3 μmOFETCarrier mobility = 8.7 cm2·V−1·s−1[60]
WettabilityC8-BTBT
single crystal
100 μmOFETCarrier mobility = 9.3 cm2·V−1·s−1[51]
WettabilityC8-BTBT/PTAA
single crystal
300–400 μmOTFTCarrier mobility = 20.6 cm2·V−1·s−1[58]
WettabilityPS nanoparticle30 μmN/AN/A[106]
LB medium 130 μmN/AN/A
PEDOT:PSS ink 2200 μmN/AN/A
Evaporation AssistantPS4–6 μmN/AN/A[116]
PSF 34–6 μmN/AN/A
PES 44–6 μmN/AN/A
Evaporation AssistantDMQA
nanowire
30–40 μmMSM 5The slope of the I–V curve = 3.1 × 10−12 S[117]
DiffusionPS0.024 μmN/AN/A[137]
DiffusionPFO4–25 μmN/AN/A[138]
1 Luria–Bertani broth. 2 poly(3,4-ethylenedioxythiophene)-poly(styrenesulfonate) (PEDOT:PSS). 3 polysulfone (PSF). 4 polyethersulfone (PES). 5 metal–semiconductor–metal (MSM).
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

Choi, W.; Kim, Y.E.; Yoo, H. Patterning of Organic Semiconductors Leads to Functional Integration: From Unit Device to Integrated Electronics. Polymers 2024, 16, 2613. https://doi.org/10.3390/polym16182613

AMA Style

Choi W, Kim YE, Yoo H. Patterning of Organic Semiconductors Leads to Functional Integration: From Unit Device to Integrated Electronics. Polymers. 2024; 16(18):2613. https://doi.org/10.3390/polym16182613

Chicago/Turabian Style

Choi, Wangmyung, Yeo Eun Kim, and Hocheon Yoo. 2024. "Patterning of Organic Semiconductors Leads to Functional Integration: From Unit Device to Integrated Electronics" Polymers 16, no. 18: 2613. https://doi.org/10.3390/polym16182613

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