Next Article in Journal
Investigating a Detection Method for Viruses and Pathogens Using а Dual-Microcantilever Sensor
Previous Article in Journal
Operator-Based Triboelectric Nanogenerator Power Management and Output Voltage Control
Previous Article in Special Issue
Clarifying the Dominant Role of Crystallinity and Molecular Orientation in Differently Processed Thin Films of Regioregular Poly(3-hexylthiophene)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Polymer Dielectric-Based Emerging Devices: Advancements in Memory, Field-Effect Transistor, and Nanogenerator Technologies

by
Wangmyung Choi
1,†,
Junhwan Choi
2,†,
Yongbin Han
1,3,
Hocheon Yoo
1,3,* and
Hong-Joon Yoon
1,3,*
1
Department of Semiconductor Engineering, Gachon University, Seongnam 13120, Republic of Korea
2
Department of Chemical Engineering, Dankook University, Yongin 16890, Republic of Korea
3
Department of Electronic Engineering, Gachon University, Seongnam 13120, Republic of Korea
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Micromachines 2024, 15(9), 1115; https://doi.org/10.3390/mi15091115 (registering DOI)
Submission received: 6 August 2024 / Revised: 28 August 2024 / Accepted: 29 August 2024 / Published: 31 August 2024
(This article belongs to the Special Issue Organic Semiconductors and Devices, 2nd Edition)

Abstract

:
Polymer dielectric materials have recently attracted attention for their versatile applications in emerging electronic devices such as memory, field-effect transistors (FETs), and triboelectric nanogenerators (TENGs). This review highlights the advances in polymer dielectric materials and their integration into these devices, emphasizing their unique electrical, mechanical, and thermal properties that enable high performance and flexibility. By exploring their roles in self-sustaining technologies (e.g., artificial intelligence (AI) and Internet of Everything (IoE)), this review emphasizes the importance of polymer dielectric materials in enabling low-power, flexible, and sustainable electronic devices. The discussion covers design strategies to improve the dielectric constant, charge trapping, and overall device stability. Specific challenges, such as optimizing electrical properties, ensuring process scalability, and enhancing environmental stability, are also addressed. In addition, the review explores the synergistic integration of memory devices, FETs, and TENGs, focusing on their potential in flexible and wearable electronics, self-powered systems, and sustainable technologies. This review provides a comprehensive overview of the current state and prospects of polymer dielectric-based devices in advanced electronic applications by examining recent research breakthroughs and identifying future opportunities.

1. Introduction

Polymer dielectric materials have become increasingly important in electronic devices because of their unique electrical, mechanical, and thermal properties [1,2,3,4]. Unlike traditional inorganic dielectrics, polymers offer flexibility, lightweight characteristics, and ease of processing, making them ideal for various applications. The ability to modify the chemical structure of polymers allows for the adjustment of their dielectric properties, enabling the design of materials with high dielectric constants, low dielectric loss, and excellent insulating performance [5,6,7,8]. These attributes are critical for developing high-performance electronic devices that meet the demands of modern technology.
The versatility of polymer dielectric materials has led to their application in various emerging electronic devices, including memory devices [9,10,11], field-effect transistors (FETs) [12,13,14,15], and triboelectric nanogenerators (TENGs) [16,17,18]. In memory devices, polymer dielectrics can improve crystallinity through molecular structure design, thereby providing stable switching characteristics in memristors [19]. Additionally, polymer dielectrics have the ability to trap electrostatic charges, resulting in a quasi-permanent electric field, making them useful as charge-trapping layers [20]. By controlling the composition of the material, the characteristics of the device can be adjusted [21,22]. For FETs, polymer dielectrics induce the field-effect when voltage is applied to a gate electrode to accumulate charge carriers while blocking leakage current. Also, polymer dielectric layers provide the interface with semiconducting layers, significantly affecting charge carrier movement in the channel. Thus, polymers in FETs determine the power consumption, operational stability, and reliability of the devices [23,24,25]. TENGs, which convert mechanical energy into electrical energy, benefit from the high charge-trapping and transfer capabilities of polymer dielectrics [26,27], leading to improved energy-harvesting efficiency. Polymer dielectrics in TENGs improve electrical output by increasing the charge generation and storage capacity, primarily due to their high dielectric constant and effective surface properties. This leads to higher surface charge density, which directly relates to improved performance [26]. Additionally, their mechanical flexibility and durability allow TENGs to maintain consistent performance over cycles, making them ideal for applications that require repeated mechanical stress. The integration of these devices into flexible and wearable electronics highlights the importance of polymer dielectrics in advancing next-generation technologies [28,29].
Polymer dielectrics have superior electrical and mechanical properties to other dielectric materials, making them a material of choice for numerous applications [30,31,32,33]. Their high dielectric strength and processability allow for the fabrication of devices with complex geometries and large surface areas. Through appropriate molecular design, polymer dielectric materials are able to exhibit intrinsic flexibility and stretchability, making them key components in flexible and stretchable devices [34,35,36,37,38,39,40]. Their high dielectric strength, combined with excellent flexibility and processability, allows for the fabrication of devices with complex geometries and large surface areas. In addition, polymers can be processed using various methods, including solution casting [41,42,43], spin coating [44,45,46,47], and printing [47,48,49], facilitating large-scale manufacturing. These properties enhance the performance and reliability of electronic devices and open new possibilities for innovative applications in flexible, stretchable, and wearable electronics.
This paper reviews the recent advances in polymer dielectric materials and discusses their applications in memory devices, FETs, and TENGs. The focus is on the unique properties that make these materials suitable for these applications, the challenges faced in their implementation, and the potential future developments that can further enhance their performance. The paper provides a comprehensive overview of their role in advancing modern electronic technologies by examining the synergies and integration possibilities of these polymer-based devices.

2. Polymer Dielectric-Based Electronics

2.1. Memory Devices

Memory is an electronic device that stores information through changes in the programming and erasing states. They can provide different output currents up to several orders of magnitude depending on the programming (Vprog) and erasing voltages (Veras) [44,50]. Therefore, the performance of a memory device is determined by the difference in the current–voltage characteristics between the programming and erasing states, known as the memory window, and the number of states that can be produced within this range. In addition, a fast memory speed is generally desirable for high-speed computing, which is determined by the extent of the memory window that can be achieved based on the duration for which Vprog and Veras are applied. Memory devices can be categorized as volatile or non-volatile based on the charge retention capability of the stored data. Another important characteristic for evaluating memory devices is endurance [51]. This metric assesses whether the memory can be operated reliably in response to repeated programming and erasing operations.
Memory devices based on polymer dielectrics have been developed where the key performance indicators for memory devices have been investigated, including low operating voltage, long retention time, and stable cycle-to-cycle endurance, even after multiple cycles [52,53]. These characteristics can vary according to the specific memory structures and material combinations. In particular, using polymer insulators allows for inherent mechanical flexibility, enabling the development of flexible or stretchable memory devices [54,55]. This paper revisits the latest results on memory devices using polymer dielectric materials as active layers. The representative memory performance and their structures and combinations are examined.

2.2. Field-Effect Transistors (FETs)

A FET is an essential component in modern electronics, acting as a switch or amplifier by differentiating the on- and off-states. Hence, increasing the on-current while maintaining the reduced off-current is crucial to maximizing their functions. The gate dielectric layer plays a vital role in FET operation, which blocks unwanted leakage current between the gate and source/drain electrodes in the off-state, accumulating charge carriers at the interface between the gate dielectric layer and the semiconductor in the on-state. In the FET operation, the drain current (ID) can be expressed using the following equations:
I D , lin = W L μ lin C i V G V T V D 2 V D
I D , sat = W 2 L μ sat C i V G V T 2
where ID,lin and ID,sat are the drain currents in the linear and saturation regime, respectively, and W and L are channel width and length, respectively. μlin and μsat are the linear and saturation mobility, respectively, and Ci is the capacitance per unit area of the dielectric layer. VG, VT, and VD are the gate voltage, threshold voltage, and drain voltage, respectively. The current output of FETs is proportional to the Ci value of the gate dielectric layer, which is determined using the following equation:
C i = ε 0 k d
where ε0, k, and d are the vacuum permittivity, dielectric constant, and thickness of the gate dielectric layer, respectively. Therefore, it is highly desirable to increase the dielectric constant and decrease the thickness of the dielectric layer to maximize ID under low operating voltages. An appropriate dielectric strength and low leakage current should be secured for the reliable operation of FETs. Exploiting advantages, such as cost-effectiveness, mechanical deformability, and tunability in their chemistry, polymer dielectric materials have been used widely to achieve reliable operation and optimize the electrical characteristics of emerging FET devices [56,57,58,59].

2.3. Triboelectric Nanogenerators (TENGs)

Since the research on TENGs was first reported in 2012 [16], the research field of TENGs has been highlighted as one of the promising energy-harvesting technologies. Triboelectricity is generated by contact charging and electrostatic induction between triboelectric materials with different dielectric properties. The dielectric constant is the central parameter that describes how efficiently the active material is polarized when subjected to an electric field and is calculated using Equation (4) [60]:
k = C × d ε 0 × A
where k, C, d, and A are the dielectric constant, capacitance, thickness of the active layer, and area of the active layer, respectively. In the metal−insulator system, the surface charge transfer between the two layers can be induced as follows:
σ = W E 0 e 1 / k t + 1 z ε 0 + 1 / N s E ¯ e 2
where σ, W, E0, t, z, e, and N s E ¯ represent the surface charge density on the dielectric material, the metal’s work function, the effective work function, the thickness of the insulator, the gap between the metal and the insulator layer, the elementary charge, and the average surface density of states in the dielectric layer, respectively [61]. Equation (5) suggests that a thinner dielectric layer combined with a larger ε/t ratio can lead to an increased surface charge density. Consequently, the permittivity of the bulk insulator may play the most crucial role in influencing contact electrification among the various factors affecting surface charge density. The permittivity of the bulk insulator can be adjusted by incorporating a high-k material layer or by embedding high-k particles. Additionally, high-k materials can act as charge-trapping layers, enhancing surface charge retention. This section explores strategies to enhance contact electrification by modifying the insulator’s permittivity.

3. Recent Advances in Electronic Devices

3.1. Memory Devices

3.1.1. Polymer Dielectric-Based Memory Devices

Memory devices based on polymer dielectrics are being developed. The use of polymer insulators provides inherent mechanical flexibility, enabling the formation of flexible or stretchable memory devices [54,55]. The most frequently attempted memory devices based on polymer insulators are two-terminal memristor structures with a sandwich configuration [62,63], where the polymer insulator is placed in the middle. A memristor operates on the memory principle where the polymer insulator typically contributes by preventing a leakage current between the two electrodes, resulting in a high resistance state (HRS). On the other hand, when metal filaments form, the current increases sharply, leading to a low resistance state (LRS). The charge trap device is another structure for memory devices based on polymer dielectric materials [20,64]. An appropriate design of polymer dielectric can enhance the charge trap memory devices, where the charge trapping in a floating gate is facilitated within low Vprog and Veras. Moreover, trapping charges within the insulator to adjust the VT of the transistor or using a polymer dielectric that induces surface charge traps as a storage layer enables charge trap-based memory operation without the need for a floating gate. This section divides the discussion into memristor-based and charge trap-based polymer memory devices. The focus is on highlighting recent developments and exploring the polymer materials, the device structures incorporating them, and the resulting memory device characteristics.

3.1.2. Application of Polymer Dielectric Materials for Conductive Filament Memristors

Polymer insulator-based memristors have attracted attention because of the low cost, simple processing, and flexibility of polymers [65,66,67,68]. As environmental concerns are increasing, biodegradable materials are also in demand [69,70]. For example, Oh et al. used poly(vinyl alcohol) (PVA) to develop an eco-friendly and flexible memristor (Figure 1a) [71]. As the molecular weight of PVA decreases, the memristor performance is improved because it allows for the increased free volume necessary for ion migration and metallization [72]. The optimized PVA-based memristor device remained stable for 104 s without current degradation (Figure 1b). A device was fabricated on an ITO-coated polyethylene naphthalate (PEN) substrate using the excellent mechanical flexibility of PVA. The device showed stable operation under tensile and compressive stresses with a bending radius of 5 mm (Figure 1c). The high solubility of PVA in water enables the development of biodegradable electronic devices, as shown in Figure 1d [73,74]. The memristor device dissolves rapidly in distilled water at room temperature (within 30 s). These results suggest the potential of facilitating flexible electronic device production and addressing the issue of electronic waste [75].
Many reported memristors based on polymer materials often show poor thermal tolerance, which can degrade the device’s performance [76]. Poly(triphenylamine) (PTPA) and PVA have been reported to form robust polymer-based memristors because of their thermal stability [77,78]. Li et al. implemented a memristor using poly[2-methoxy-5-(3′,7′-dimethyloctyloxy)-1,4-phenylenevinylene] (MDMO-PPV), showing robust performance in both extremely low- and high-temperature environments [79]. The MDMO-PPV thin film was deposited on an ITO-patterned glass substrate using the spin-coating technique (Figure 1e), achieving more than 95% yield with consistent performance across an 8 × 8 array. The MDMO-PPV-based memristor devices demonstrated real-time measurements across an ultra-wide temperature range (173–473 K), exhibiting reliable operation from 77 to 573 K in cryogenic and high-temperature environments (Figure 1f). The memristor characteristics remained stable even at temperatures as low as 77 K for 72 h.
One notable issue in solution-processed polymer memristors is performance degradation due to impurities or residues [80]. This can be resolved using a solvent-free initiated chemical vapor deposition (iCVD) technique. Kim et al. reported memristor devices based on poly(1,3,5-trivinyl-1,3,5-trimethyl cyclotrisiloxane) (pV3D3) using iCVD, showing excellent stability [51]. pV3D3 exhibits high thermal and chemical stability because of its high crosslinking density resulting from three crosslinking sites [80,81]. These characteristics can enhance the durability of the device. Indeed, the memristor devices based on pV3D3 maintained robust electrical properties after 105 s and 105 endurance cycles (Figure 1g). They also exhibited reliable retention characteristics, even at 85 °C. Crossbar-array memristor devices fabricated in a 32 × 32 matrix were manufactured using the solvent-free iCVD technique (Figure 1h). This method ensures that all electronic devices operate properly without impurities or residues from solution processes.
Studies have attempted to improve the performance of polymer-based memristor devices by exploring the introduction of interfacial load layers containing active metal clusters and constructing active layers with oriented molecular chain arrangements to direct filament growth [67,82]. Zhang et al. enhanced the performance of memristors by adding the metal salt AlClO4 to polyethylenimine (PEI) [83]. Compared to pure PEI-based memristors, devices containing AgClO4 allowed for a large number of silver (Ag) ions in the PEI layer to move more quickly to the platinum (Pt) electrode, contributing to the formation of Ag filaments. This phenomenon reduced the operating voltage of the device and enhanced its energy consumption efficiency (Figure 1i). Furthermore, adding AgClO4 improved the non-volatile performance of the memristor (Figure 1j), and its switching characteristics remained unchanged, even after exposure to air for 30 days (Figure 1k).
In this chapter, we revisited the deposition methods according to the polymer dielectric-based material classification. Table 1 represents a detailed summary of the polymer dielectric-based memristor devices in terms of polymer dielectric materials, deposition methods, device performance, environmental stability, and biodegradability parameters.
Figure 1. (a) Schematic diagram and photograph of PVA-based memristors. The memory retention characteristics of memristors investigated for (b) rigid substrate-based and (c) flexible substrate-based devices. (d) The photograph of the biodegradation process of a PVA-based memristor device at 300 K [71]. Copyright © 2023, John Wiley and Sons. (e) The schematic diagram of a memristor device based on MDMO-PPV, and (f) its memory retention characteristics at temperatures ranging from 77 to 573 K [79]. Copyright © 2023, John Wiley and Sons. (g) The memory retention characteristics of a pV3D3-based memristor device over time and endurance cycles, and (h) optical microscopy image of a 32 × 32 array of memristors [51]. Copyright © 2024, John Wiley and Sons. (i) Variation in the SET voltage with the presence of metal salt (AgClO4) in PEI-based memristors. (j) The memory retention characteristics of PEI-AgClO4-based memristor. (k) The resistance switching curve of PEI-AgClO4-based memristor after 30 days of exposure to air [83]. Copyright © 2022, American Chemical Society.
Figure 1. (a) Schematic diagram and photograph of PVA-based memristors. The memory retention characteristics of memristors investigated for (b) rigid substrate-based and (c) flexible substrate-based devices. (d) The photograph of the biodegradation process of a PVA-based memristor device at 300 K [71]. Copyright © 2023, John Wiley and Sons. (e) The schematic diagram of a memristor device based on MDMO-PPV, and (f) its memory retention characteristics at temperatures ranging from 77 to 573 K [79]. Copyright © 2023, John Wiley and Sons. (g) The memory retention characteristics of a pV3D3-based memristor device over time and endurance cycles, and (h) optical microscopy image of a 32 × 32 array of memristors [51]. Copyright © 2024, John Wiley and Sons. (i) Variation in the SET voltage with the presence of metal salt (AgClO4) in PEI-based memristors. (j) The memory retention characteristics of PEI-AgClO4-based memristor. (k) The resistance switching curve of PEI-AgClO4-based memristor after 30 days of exposure to air [83]. Copyright © 2022, American Chemical Society.
Micromachines 15 01115 g001

3.1.3. Application of Polymer Dielectric Materials for Charge Trap Memory Transistors

Appropriate dielectric materials and thin thicknesses are required for floating gate-based transistor devices with a low operating voltage and stable endurance. The insulating layer preventing a leakage current at the control gate should be thin with a high dielectric constant to ensure low operating voltage. In contrast, the tunneling dielectric layer between the floating gate and the channel should use materials with a low dielectric constant to maintain charge trapping. Yang et al. developed MoS2-based low-voltage non-volatile memory devices using a high-k poly(2-cyanoethyl acrylate-co-diethylene glycol divinyl ether) [p(CEA-co-DEGDVE)] copolymer (pC1D1, k~6.2) that was synthesized from 2-cyanoethyl acrylate (CEA) and di(ethylene glycol) divinyl ether (DEGDVE), and a low dielectric constant pV3D3 polymer (k~2.2) [50]. The high energy barrier between MoS2 and pV3D3 enabled programming and erasing through tunneling, while gold nanoparticles (Au NPs) captured the charges from the deep quantum well (Figure 2a). The shallow lowest unoccupied molecular orbital (LUMO) level of pC1D1 prevented charges from moving to the control gate. In this energy configuration, the thin thickness and high dielectric constant of pC1D1 resulted in a low operating voltage (±13 V) and robust memory operation (Figure 2b). The low dielectric constant of pV3D3 helps prevent the de-trapping of the charges captured by the Au NPs, maintaining good performance (<105 s).
A polymer electret is a dielectric material that can be used as a charge-trapping layer without a floating gate [87]. Chen et al. examined the characteristics of non-volatile memory transistors using poly(9,9-dioctylfluorene) (PFO) as the charge-trapping layer [88]. Using PFO as a charge-trapping layer in memory transistors resulted in semiconductor holes being trapped in the PFO layer during the electrical erasing process, shifting the VT of the memory transistor in the direction of negative gate voltage (erasing). When PFO was exposed to light that is capable of activating it, PFO generated photo-generated electron–hole pairs. During this process, the photo-generated electrons recombine with the holes trapped in the PFO, neutralizing them, while the photo-generated holes transport to the organic semiconductor layer, maintaining the initial state (Figure 2c,d).
Additional materials are often added to polymer dielectrics [89,90] or block polymers [91,92], where two or more polymers are regularly linked within a single polymer chain to enhance the performance of memory transistors. This approach exhibits an operational mechanism similar to floating gate-based memory transistors. For example, Hsu et al. introduced a block copolymer of poly(9,9-di-n-hexyl-2,7-fluorene) and poly(δ-decanolactone) (PF-b-PDL) as the charge-trapping layer in memory transistors [21]. The PF chain served to trap holes, while the PDL chain acted as an insulating polymer serving as the dielectric layer and maintained the trapped charges in PF (Figure 2e). In the PF-b-PDL structure, as the PDL arm increased, the memory window increased from 63 V to 102 V (Figure 1f), and the memory retention characteristics were improved, maintaining a memory ratio of 3.5 × 104. These attributes stem from the increased branching of PDL stabilizing the charges stored in PF. Zhang et al. developed memory transistors using phenyl-C61-butyric acid methyl ester (PCBM) and poly(2-vinyl naphthalene) (PVN), with the ability to detect and store light selectively based on the PCBM blending content (Figure 2g) [22]. The roles of PVN and PCBM in the device varied according to the blending content of PCBM. In a pure PVN device (Device A), the electrons generated by exposure to ultraviolet light were trapped in the PVN, enabling memory operation. In the device with 2.5% PCBM (Device B), when exposed to blue light, the electrons photoexcited in the PCBM were transported to and trapped in the PVN, with PCBM playing a supporting role. In the device with 30% PCBM (Device C), the electrons generated by red light were trapped in the PCBM, which functioned as a charge-trapping layer, while the PVN prevented electron detrapping from the PCBM. Each device demonstrated a correlation between the light exposure energy and the surface density of the trapping layer through variations in VT with the exposure time. All the devices showed an increase in the surface density of the trapping layer as the light exposure energy increased, as shown in Figure 2h. The calculated sensitivity values were 2.3 × 1011 (Device A), 0.3 × 1011 (Device B), and 0.6 × 1011 mJ−1 (Device C).
To clearly present the characteristics of charge trap-based memory devices according to the types of polymer dielectrics and device types, we provided a detailed summary focusing on the structure and type of charge traps, as well as the device performance parameters (Table 2).
Figure 2. (a) Operating mechanism of MoS2-based memory for programming and erasing operations, and (b) memory retention characteristics under ±13 V with a pulse width of three seconds for operating gate pulses [50]. Copyright © 2019, John Wiley and Sons. (c) The schematic diagram of the photo-induced writing mechanism in pre-erased charge-trap polymer layer-based memory devices. (d) The writing/erasing characteristics of PFO charge-trap layer-based memory devices under various wavelengths [88]. Copyright © 2021, American Chemical Society. (e) The operating mechanism of PF-b-PDL-based memory for programming and erasing operations, and (f) the writing/erasing characteristics of memory devices based on the number of PDL arms in the PF-b-PDL structure [21]. Copyright © 2021, American Chemical Society. (g) The schematic diagram of the structure of PCBM/PVN blend-based memory devices and the selective photoresponse of the device according to the PCBM blending ratio. (h) The surface density of trapped charge versus light exposure energy per unit area curves of PCBM/PVN blend-based memory devices with different PCBM blending ratios [22]. Copyright © 2019, American Chemical Society.
Figure 2. (a) Operating mechanism of MoS2-based memory for programming and erasing operations, and (b) memory retention characteristics under ±13 V with a pulse width of three seconds for operating gate pulses [50]. Copyright © 2019, John Wiley and Sons. (c) The schematic diagram of the photo-induced writing mechanism in pre-erased charge-trap polymer layer-based memory devices. (d) The writing/erasing characteristics of PFO charge-trap layer-based memory devices under various wavelengths [88]. Copyright © 2021, American Chemical Society. (e) The operating mechanism of PF-b-PDL-based memory for programming and erasing operations, and (f) the writing/erasing characteristics of memory devices based on the number of PDL arms in the PF-b-PDL structure [21]. Copyright © 2021, American Chemical Society. (g) The schematic diagram of the structure of PCBM/PVN blend-based memory devices and the selective photoresponse of the device according to the PCBM blending ratio. (h) The surface density of trapped charge versus light exposure energy per unit area curves of PCBM/PVN blend-based memory devices with different PCBM blending ratios [22]. Copyright © 2019, American Chemical Society.
Micromachines 15 01115 g002

3.2. Field-Effect Transistors (FETs)

3.2.1. Overview of Polymer Dielectric Materials for FET Devices

Various polymer gate dielectric materials have been studied extensively to accommodate the important requirements in FETs, including high carrier mobility (μ), low operating voltage, and device stability. High-k polymer dielectric materials have been designed to improve Ci by enhancing the polarization of the polymer materials [36,97]. Polyvinylidene fluoride (PVDF)-based ferroelectric polymers are one of the representative high-k polymer dielectric materials, which can also exhibit unique electrical properties, including negative capacitance (NC) behavior [98,99]. Reducing the thickness of the gate dielectric layer is another efficient way to increase the Ci value, and crosslinking has been explored to maintain the proper dielectric strength while decreasing the thickness of the dielectric layer [100]. Crosslinking is also important in improving mechanical deformability, which allows for designing stretchable polymer dielectric materials [101,102]. In addition to the gate dielectric layer, polymer dielectric materials have been introduced as doping and barrier layers to improve the charge transport characteristics and environmental stability, respectively [103,104].
Polymer dielectric materials have also been used to develop organic–inorganic hybrid dielectric layers, allowing for the synergistic effects of both organic and inorganic materials [105]. In layered hybrid dielectrics, the polymer dielectric layer can effectively passivate the surface trap sites of the inorganic dielectric layers and improve the mechanical flexibility of FET devices [106,107]. Furthermore, hybrid dielectric materials based on organic–inorganic mixtures have been developed, where inorganic nanoparticles or inorganic segments are incorporated in polymer matrices [108,109]. The following section highlights the recent research achievements in polymer-based dielectric materials for FET applications.

3.2.2. Polymer-Based Dielectric Materials for FET Devices

Polymer dielectric materials containing polar functionalities, such as cyano (−C≡N), hydroxyl (−OH), and epoxy groups, have been used to improve the dielectric constants [23,97,110,111,112,113]. For example, Yoo et al. developed high-k polyurea gate dielectric materials where the polar urea bonds were responsible for a high dielectric constant [94]. Polyurea dielectric materials were synthesized from different diisocyanate and diamine monomers. The effects of hydrogen bonding on dielectric properties were investigated systematically by analyzing the dielectric properties of polyurea with different chemical structures. Without the alkyl substituent chains in the polymer backbone, a high dielectric constant (k~5.8) was obtained by strengthening the hydrogen bonding interactions (Figure 3a). In addition, the leakage current density (J) was decreased because of the reduced free volume, resulting in low J even with 60 nm thickness (Figure 3b). Dinaphtho[2,3-b:2′,3′-f]thieno[3,2-b]thiophene (DNTT)-based organic FETs were fabricated using the polyurea dielectric layer, which showed low operating voltages (<3 V). The charge carrier μ was improved from 0.069 cm2·V−1·s−1 (untreated) to 1.390 cm2·V−1·s−1 (with the surface treatment layer) with the metal oxide-assisted surface treatment layer. Table 3 summarizes the key parameters of the FET devices based on polymer dielectric materials that are the focus of this paper.
Crosslinking is a powerful strategy to improve the dielectric strength of polymer dielectric materials by densifying the polymer matrices [2,5,80,114,115,116,117,118,119,120,121,122,123,124,125,126]. A previous study reported a reduced leakage current after a crosslinking structure was introduced in polymer dielectric layers [127,128]. Furthermore, Choi et al. investigated the effect of the chain length in the crosslinkers [5]. A monomer containing polar cyano group, CEA, was copolymerized with DEGDVE and 1,4-butanediol divinyl ether (BDDVE). These crosslinkers have similar chemical structures but different chain lengths. High-purity polymer films and homogenous mixing between CEA and crosslinker were achieved using the iCVD process to fabricate polymer dielectric layers. The decrease in dielectric constant according to the addition of a crosslinker was alleviated in poly(2-cyanoethyl acrylate-co-1,4-butanediol divinyl ether) [p(CEA-co-BDDVE)] compared with p(CEA-co-DEGDVE) (Figure 3c). Hence, with a similar crosslinker mole fraction, the copolymer dielectric with a short-chain crosslinker can exhibit a higher dielectric constant. 2,7-Dioctyl[1]benzothieno[3,2-b][1]benzothiophene (C8-BTBT)-based FETs were fabricated using p(CEA-co-BDDVE) with the optimal chemical composition, which showed a low operating voltage of less than 5 V and a high carrier level μ (>3.0 cm2·V−1·s−1). Moreover, the organic FETs were fabricated on a plastic substrate, maintaining their electrical characteristics even under the applied tensile strain up to ~2.4%.
Recently, intrinsically stretchable FET devices with highly deformable electronics, such as stretchable displays and electronic skins, have emerged [34,126,127,128,129,130]. For the reliable operation of stretchable FETs, intrinsically stretchable polymer dielectric materials have been widely researched, and crosslinking also plays a significant role in designing such materials [35,101,102,131,132,133,134,135,136,137,138,139]. Kang et al. developed a stretchable polymer dielectric composed of dicarboxy-terminated poly(acrylonitrile-co-butadiene) (CTBN) as a soft matrix and 1,6-bis(trichlorosilyl)hexane (C6) as a hard segment (Figure 3d) [101]. The leakage current was decreased as a portion of the C6 crosslinking agent increased. In particular, a metal–insulator–metal (MIM) device with CTBN:7wt.%-C6 maintained J of ~10−8 A·cm−2 even under an applied strain of 30%. Poly(3-hexylthiophene) nanowires (P3HT NWs)/polydimethylsiloxane (PDMS) blend organic semiconductor-based stretchable FETs were fabricated using a CTBN:7wt.%-C6 dielectric layer (Figure 3e). Although μ decreased with the applied strain, the stretchable FETs retained a gate leakage current of ~10−9 A, enabling FET devices with distinguishable on- and off-states.
In addition to the gate dielectric layer, polymer dielectric materials have been used as an additional layer on top of FETs to enhance the charge transport characteristics and the environmental stability of FET devices [104,140,141,142,143,144,145,146,147,148]. An electron-rich polyethylenimine (PEI) is a widely used n-doping polymer in metal oxide and 2-dimensional (2D) semiconductor FETs [149,150,151,152,153]. Hong et al. introduced truncated-branched PEI on top of 2D molybdenum disulfide (MoS2) semiconductor FET devices (Figure 3f) [150]. Compared to pristine 2D MoS2 FETs (~3.7 V), the PEI-doped device showed VT = ~0.72 V, and this negative shift of VT supported the strong n-doping effect of PEI (Figure 3g). Moreover, the on-current was increased from ~3.3 μA to ~4.8 μA, which could be explained by the reduced Schottky barrier and facilitated electron injection from the source to the channel layer with the PEI doping layer. In addition, 2D MoS2 FET with a PEI doping layer was used as a photodetector, which exhibited enhanced photo-responsivity compared to the pristine device. When light is irradiated, some of the holes in the photo-generated electron–hole pairs are trapped in a new energy level formed by PEI, leading to band-bending in the energy band and enhancing the photoresponsivity.
One of the noticeable issues in FETs with emerging semiconductor materials, such as organic and 2D semiconductors, is their susceptibility to ambient air, and hydrophobic polymers can be used as barrier layers [104,140,141,143,146,148]. Park et al. introduced a barrier polymer on top of nanostructured 2D MoS2 FETs [104]. The nanostructured 2D materials can exhibit unique electrical and optical properties by periodically removing the layer. On the other hand, they showed some disadvantages related to the lack of electrical stability resulting from the edge exposures. These issues were addressed by depositing poly(1H,1H,2H,2H-perfluorodecyl methacrylate) (pPFDMA) on nanoporous 2D MoS2 FETs by the iCVD process, which only exhibited a marginal change in the electrical characteristics of the FET devices owing to the solvent-free nature of the iCVD process. The hydrophobic pPFDMA effectively passivated nanoporous 2D MoS2 FETs and prevented moisture penetration. As a result, the ambient stability was improved remarkably in the pPFDMA-passivated nanoporous 2D MoS2 FETs compared to the pristine, non-passivated ones (Figure 3h). As extensively described above, the charge accumulation capability and dielectric reliability were significantly improved by enhancing the dielectric constant and/or introducing a crosslinking network. In addition to the gate dielectric layer, polymer materials can be used as functional adlayers to optimize the electrical characteristics and elevate the ambient stability of FET devices. Another effective strategy is utilizing polyelectrolytes and their composites to boost the charge transport performance and lower the power consumption of FETs, which also paves the way to develop FET-based sensors and neuromorphic devices [154,155,156,157,158,159]. Additionally, introducing ferroelectric polymers and incorporating inorganic components allow for further enhancing the performance of FETs or imparting unique behaviors, as described in the following sections.
Table 3. Summary of polymer gate dielectric-based FETs.
Table 3. Summary of polymer gate dielectric-based FETs.
Dielectric
Material
SemiconductorOperating Voltage
(V)
Current on/off RatioMobility
(cm2·V−1·s−1)
Subthreshold Swing
(mV·dec−1.)
Applied Strain
(%)
[Ref.]
PolyureaDNTT3>1051.393702[97]
P(CEA-co-BDDVE)C8-BTBT5>1063.3987.42.5[5]
CTBN-C6P3HT NWs/PDMS50>1040.0031N/A34[101]
pentacene50>1060.12N/AN/A
PVDF-TrFE
(vertically poled)
TIPS-pentacene4>1040.08200N/A[160]
PVDF-TrFE
(textured poled)
TIPS-pentacene4>1051350N/A
4FCOC 1C10-DNTT30>1071.30N/AN/A[161]
PTCDI-C13 230>1060.13N/AN/A
PVDF-TrFE2D MoS240>105N/A24.2N/A[162]
HfO2/PVDF-TrFE2D MoS22.5>105N/A42.5N/A[163]
PMMA/Al2O32D InSe8>1071055300N/A[107]
P(CEA-co-DEGDVE)/Al2O3C8-BTBT3>1062.4868.41.2[164]
CYTOP/HfO2-Al2O3TIPS-pentacene/PTAA10>1050.8700N/A[165]
diF-TES-ADT 3/PTAA10>1051.4320N/A
PSAND-4pentacene3>1030.40N/AN/A[106]
F16CuPC 42>1010.020N/AN/A
DPP-DTT3>1030.17N/A1.2
PDIF-CN2 52>1020.18N/AN/A
BaTiO3/PDMSCarbon nanotube30>1024N/A50[166]
1 cyclo-olefin copolymer (COC) with fluorophenyl azide (FPA) crosslinker (4FCOC). 2 N,N’-Ditridecyl-3,4,9,10-perylenetetracarboxylic diimide (PTCDI-C13). 3 2,8-difluoro-5,11-bis(triethylsilylethynyl) anthradithiophene (diF-TES-ADT). 4 1,2,3,4,8,9,10,11,15,16,17,18,22,23,-24,25-hexadecafluoro-29H,31H-phthalocyanine (F16CuPc). 5 N,N′-1H,1Hperfluorobutyl dicyanoperylenecarboxydiimide (PDIF-CN2).
Figure 3. (a) Ci−frequency (f) and (b) J−electric field (E) curves of polyurea with different thickness [97]. Copyright © 2018, American Chemical Society. (c) The change in dielectric constant according to the mole fraction of the crosslinker of p(CEA-co-DEGDVE) and p(CEA-co-BDDVE) [5]. Copyright © 2020, John Wiley and Sons. (d) The schematic diagram and the chemical structure of stretchable polymer dielectric. (e) The schematic diagram and photograph of P3HT NWs/PDMS-based stretchable organic FETs [101]. Copyright © 2018, American Chemical Society. (f) The schematic diagram of PEI-doped 2D MoS2 FET and (g) its transfer curve compared to the pristine device [150]. Copyright © 2017, John Wiley and Sons. (h) The change in VT according to the storage time of pPFDMA-passivated and non-passivated nanoporous 2D MoS2 FETs [104]. Copyright © 2022, Springer Nature.
Figure 3. (a) Ci−frequency (f) and (b) J−electric field (E) curves of polyurea with different thickness [97]. Copyright © 2018, American Chemical Society. (c) The change in dielectric constant according to the mole fraction of the crosslinker of p(CEA-co-DEGDVE) and p(CEA-co-BDDVE) [5]. Copyright © 2020, John Wiley and Sons. (d) The schematic diagram and the chemical structure of stretchable polymer dielectric. (e) The schematic diagram and photograph of P3HT NWs/PDMS-based stretchable organic FETs [101]. Copyright © 2018, American Chemical Society. (f) The schematic diagram of PEI-doped 2D MoS2 FET and (g) its transfer curve compared to the pristine device [150]. Copyright © 2017, John Wiley and Sons. (h) The change in VT according to the storage time of pPFDMA-passivated and non-passivated nanoporous 2D MoS2 FETs [104]. Copyright © 2022, Springer Nature.
Micromachines 15 01115 g003

3.2.3. Ferroelectric Polymers for FET Devices

PVDF-based ferroelectric polymers, such as poly(vinylidene fluoride trifluorethylene) (PVDF-TrFE) and poly(vinylidene fluoride-trifluoroethylenechlorofloroethylene) (PVDF-TrFE-CFE), have been investigated widely in FETs because of their high dielectric constant at room temperature [160,167,168,169,170,171,172,173,174,175,176,177,178,179]. Laudari et al. investigated the effect of poling in ferroelectric polymers using 6,13-bis(triisopropylsilylethynyl)pentacene (TIPS-pentacene) organic FETs to achieve reliable operation of FET devices based on ferroelectric polymer dielectrics [160]. The poling of the PVDF-TrFE was controlled by applying an external electric field during film crystallization, which enabled the fabrication of uniform, vertical poling, and textured poling (Figure 4a). Compared to the vertically poled PVDF-TrFE-based devices, TIPS-pentacene FETs with textured poled PVDF-TrFE exhibited an enhanced on-current and slightly decreased off-current (Figure 4b,c). Quantitatively, the textured poled PVDF-TrFE FET device showed the μ of ~1 cm2·V−1·s−1, current on–off ratio (Ion/Ioff) of 105, whereas vertically poled PVDF-TrFE FET showed μ of ~0.08 cm2·V−1·s−1 and 104. The device performance was improved by an order of magnitude in textured poled PVDF-TrFE FETs while maintaining the low operating voltage of less than 4 V.
Blending and layer-by-layer structures have been proposed to improve the electrical and thermal properties of the ferroelectric polymer dielectrics [180,181,182,183,184,185,186,187]. Crosslinking is an efficient way to improve the electrical characteristics of the ferroelectric polymers and the resulting FET devices [161,188,189,190]. For example, Kwon et al. introduced fluorophenyl azide (FPA) to induce the crosslinking network in poly(vinylidene fluoride-co-hexafluoropropylene) (PVDF-HFP) (Figure 4d) [161]. A series of FPA was synthesized with a different number of fluorobenzene rings (2, 3, and 4), which can react with C−H bonds in conventional polymers under ultraviolet (UV) irradiation. Hysteresis in the transfer curves of 2,9-di-decyl-dinaphtho-[2,3-b:2′,3′-f]-thieno-[3,2-b]-thiophene (C10-DNTT) organic FETs disappeared through FPA crosslinking (UV-treated FPVDF-HFP) (Figure 4e). In the PVDF-based ferroelectric polymer dielectric, carbon (C)−fluorine (F) dipoles were rearranged under a high electric field, enabling ferroelectric behavior. This rearrangement of C−F dipoles remained even after the applied electric field was removed, causing the clockwise hysteresis in the current−voltage (IV) characteristics of FET devices. On the other hand, with the addition of the FPA crosslinker, the grain size of the VDF crystals and the number of ferroelectric domains were decreased, reducing the hysteresis behavior of the FET devices.
NC behavior, which achieves a lower subthreshold swing (SS) than the theoretical limit at room temperature (~60 mV·dec−1.), is another attractive property of ferroelectric polymer dielectric materials [162,163,191,192,193,194]. Wang et al. developed NC-FETs based on 2D MoS2 and ferroelectric P(VDF-TrFE) (Figure 4f) [162]. The thickness of P(VDF-TrFE) in 2D MoS2 NC-FETs was varied from 50 to 300 nm and all the devices exhibited SS values lower than 60 mV·dec−1. (Figure 4g). In particular, the SS values decreased gradually as the thickness of P(VDF-TrFE) increased because there was a certain layer at the interface between P(VDF-TrFE) and the top Al electrode that barely contributed to the NC behavior. This was confirmed by linear fitting, where a fixed capacitance of 5.5 nF was observed even when the thickness of P(VDF-TrFE) was projected to be zero. Nevertheless, the ultralow SS value of ~24.2 mV·dec−1. was obtained with a thick P(VDF-TrFE) dielectric layer (~300 nm). Liu et al. also developed NC-FETs using 2D MoS2 and P(VDF-TrFE), and 4 nm-thick HfO2 was inserted between 2D MoS2 and P(VDF-TrFE) to achieve the reliable NC behavior and robust operation of FETs [163]. Furthermore, silver nanowires (AgNWs) were used as gate electrodes to reduce the channel length and maximize the output current of the NC-FETs. An SS value lower than 60 mV·dec−1. and high transconductance (gm) were maintained with channel lengths greater than 60 nm (Figure 4h). With further decreases in channel length, SS increased, and gm decreased because of the short channel effects. Nevertheless, this study showed the great potential of 2D channel material and ferroelectric polymers in high-performance nano-scale electronics.
Figure 4. (a) Schematic diagram of vertically poled and textured poled P(VDF-TrFE). The transfer curves of TIPS-pentacene FETs with (b) vertically poled and (c) textured poled P(VDF-TrFE) as a gate dielectric layer [160]. Copyright © 2019, John Wiley and Sons. (d) The schematic diagram of PVDF-HFP and FDA-crosslinked PVDF-HFP (FPVDF-HFP). (e) The transfer curves of C10-DNTT organic FETs based on pristine PVDF-HFP and crosslinked FPVDF-HFP [161]. The arrows represent forward and backward sweep of VG. Copyright © 2020, American Chemical Society. (f) The schematic diagram of the device structure of the 2D MoS2 NC-FETs and (g) their transfer curves with different P(VDF-TrFE) thicknesses [162]. Copyright © 2017, Springer Nature. (h) The change in SS and gm according to the channel length of 2D MoS2 NC-FETs based on HfO2/P(VDF-TrFE) gate dielectric layer [163]. Copyright © 2018, John Wiley and Sons.
Figure 4. (a) Schematic diagram of vertically poled and textured poled P(VDF-TrFE). The transfer curves of TIPS-pentacene FETs with (b) vertically poled and (c) textured poled P(VDF-TrFE) as a gate dielectric layer [160]. Copyright © 2019, John Wiley and Sons. (d) The schematic diagram of PVDF-HFP and FDA-crosslinked PVDF-HFP (FPVDF-HFP). (e) The transfer curves of C10-DNTT organic FETs based on pristine PVDF-HFP and crosslinked FPVDF-HFP [161]. The arrows represent forward and backward sweep of VG. Copyright © 2020, American Chemical Society. (f) The schematic diagram of the device structure of the 2D MoS2 NC-FETs and (g) their transfer curves with different P(VDF-TrFE) thicknesses [162]. Copyright © 2017, Springer Nature. (h) The change in SS and gm according to the channel length of 2D MoS2 NC-FETs based on HfO2/P(VDF-TrFE) gate dielectric layer [163]. Copyright © 2018, John Wiley and Sons.
Micromachines 15 01115 g004

3.2.4. Polymer–Inorganic Hybrid Dielectric Materials for FET Devices

Organic–inorganic hybrid dielectric materials have been developed to accommodate the advantages of organic and inorganic materials, and polymers have been incorporated as organic components in hybrid dielectrics [100,195,196]. For example, polymer dielectric layers have been inserted between inorganic dielectric layers and semiconductors to passivate interface traps of the inorganic layers [107,164,197,198,199,200,201,202]. Feng et al. used poly(methyl methacrylate) (PMMA)/Al2O3 bilayer dielectrics for 2D indium selenide (InSe) FET devices (Figure 5a) [107]. The PMMA layer passivated the hydroxyl (−OH) groups of Al2O3 and suppressed charge carrier scattering, which was supported by analyzing the performance of FET. The μ value of the 2D InSe FET with the Al2O3 dielectric layer was ~64 cm2·V−1·s−1, which was enhanced drastically to ~1055 cm2·V−1·s−1 with the PMMA layer. In addition, the SS value was improved from 600 to 300 mV·dec−1. by inserting PMMA. This enhancement in the FET performance was consistent when SiO2 replaced Al2O3 as the inorganic dielectric layer, and the increase in the μ value was commonly obtained with different thicknesses of 2D InSe (10–50 nm). These results suggested that the Coulomb impurities or surface polar phonon scattering can be suppressed by the polymer dielectric layer. Similarly, Choi et al. developed polymer–inorganic hybrid dielectrics consisting of plasma-grown Al2O3 and high-k p(CEA-co-DEGDVE) [164]. The polymeric layer thickness was adjusted precisely on the nanometer scale by tuning the deposition time in the iCVD process. μ was increased, and the interface trap density (NSS) was decreased in C8-BTBT organic FETs by introducing the polymer dielectric layer (Figure 5b). The thickness of the polymer dielectric was optimized to 10 nm to achieve high-performance FET devices and reliable dielectric properties, and the total thickness of the layered hybrid dielectric was 15 nm (5 nm-thick Al2O3 and 10 nm-thick high-k polymer). Based on this high dielectric constant and ultralow thickness, low-voltage (<3 V), flexible C8-BTBT organic FETs were demonstrated, which exhibited 100% yield and a high areal uniformity in the FET characteristics.
FET devices with high operational stability were achieved by exploiting the reduced trap densities of layered hybrid dielectrics [165,200,203,204,205,206,207]. Jia et al. fabricated TIPS-pentacene/poly[bis(4-phenyl) (2,4,6-trimethylphenyl) amine] (PTAA) organic FETs and 35 nm-thick CYTOP and 33 nm-thick HfO2-Al2O3 nanolaminate were used as a gate dielectric layer [165]. Extremely high operational stability was achieved in the organic FETs where the change in ID was less than 4% of the initial ID even after 40 h of the continuous bias stress, which was well-matched with modeling results using double-stretched exponential (DSE) function (Figure 5c). This long-term bias stress stability could be explained by two competing processes, where ID decreased by charge trapping at the interface between the semiconductor and polymer dielectric layer, being compensated by remnant dipole polarization in the inorganic layer [164,208]. In addition, the FETs were fabricated in top-gate geometry, allowing for an improved environmental stability.
Incorporating polymer dielectric materials into hybrid dielectrics can improve mechanical flexibility by mitigating the brittleness of inorganic dielectric materials [4,106,209,210,211]. For example, Chen et al. developed polymeric self-assembled nanodielectric (PSAND) by the solution-processed layer-by-layer assembly of a P-PAE polymer chromophore and zirconia (ZrOx) [106]. The thickness of the hybrid insulating layer was adjusted precisely according to the assembly frequency (PSAND-x), and the fabricated hybrid insulating layer exhibited high capacitance (558, 472, 406, and 339 nF·cm−2 for PSAND-1, 2, 3, and 4, respectively). In particular, the Ci value varied linearly with the PSAND thickness, indicating a well-defined multilayer structure. Diketopyrrolopyrrole-dithienyl-thieno[3,2-b] thiophene (DPP-DTT) organic FETs were fabricated on polyimide (PI) substrates using the PSAND-4 as a dielectric layer (Figure 5d), with an operating voltage less than 3 V enabled by high Ci of PSAND-4. The fabricated flexible organic FETs only exhibited a marginal change in their μ and VT with the bending radius from ∞ to 1 mm, showing good mechanical flexibility of the self-assembled hybrid dielectrics.
A nanocomposite is another type of polymer–inorganic hybrid dielectric material, where inorganic nanoparticles are dispersed within a polymer matrix [166,212,213,214,215,216]. Cai et al. developed fully printable, semiconducting single-walled carbon nanotube (SWCNT) FETs using the nanocomposite hybrid dielectrics consisting of barium titanate (BaTiO3) nanoparticles and PDMS (Figure 5f,g) [166]. Unsorted carbon nanotubes (CNTs) were used as gate and source/drain electrodes, of which the resistance decreased significantly as the number of printing cycles increased. The dielectric constant of the BaTiO3/PDMS composite increased with the BaTiO3 volume content increased, and it reached ~9 at a BaTiO3 volume content of ~26%. A Ci value of ~2.8 nF·cm−2 was obtained, even with a thickness of ~2 μm owing to the high dielectric constant. The composite dielectric layer exhibited robust insulating properties with J < 10−8 A·cm−2 under a voltage as high as ±200 V. Moreover, the Ci values of the BaTiO3/PDMS composite showed only a marginal change under the repeated stretching with the applied strain of 50% for 2000 cycles. This excellent mechanical property made it possible to fabricate stretchable SWCNT FETs on a PDMS substrate, which retained its FET operation under an applied strain of 50% and under the 1400 cycles of repeated stretching with 50% strain. The μ value tended to decrease, and Ion/Ioff increased according to the applied strain and strain cycle, possibly due to the structural changes in the SWCNT network and the stronger gating effect, respectively (Figure 5h). Furthermore, stretchable logic circuits, including NOR and NAND gates, were demonstrated by integrating FET devices.
In addition to the hybrid dielectrics described above, homogeneous polymer–inorganic hybrid dielectric materials were developed using solution-phase fabrication [108,217,218,219] and vacuum deposition [109,220,221] to improve the electrical and mechanical properties of the dielectric layers.
Figure 5. (a) Schematic diagram of the 2D InSe FET with a PMMA/Al2O3 gate dielectric layer [107]. Copyright © 2014, John Wiley and Sons. (b) Change in the μ and NSS of C8-BTBT FETs according to the thickness of the polymeric layer in p(CEA-co-DEGDVE)/Al2O3 hybrid dielectric [164]. Black and blue lines represent mobility and trap density, respectively. Copyright © 2022, John Wiley and Sons. (c) ID change with respect to the bias stress time of TIPS-pentacene/PTAA FETs with the CYTOP/HfO2-Al2O3 dielectric layer [165]. Copyright © 2018, American Association for the Advancement of Science. (d) The schematic diagram of flexible DPP-DTT FETs based on PSAND-4 dielectrics and (e) their mobility and VT change according to the bending radius [106]. Copyright © 2020, John Wiley and Sons. (f) The schematic diagram of the fabrication process of SWCNT FETs with the BaTiO3/PDMS composite dielectric layer and (g) the optical microscopy image of the fabricated device. (h) The change in μ and Ion/Ioff with respect to the applied strain and strain cycles of SWCNT FETs [166]. Red and blue dots represent mobility and Ion/Ioff in the upper graph, respectively, and green and yellow dots represent mobility and Ion/Ioff in the lower graph, respectively. Copyright © 2016, American Chemical Society.
Figure 5. (a) Schematic diagram of the 2D InSe FET with a PMMA/Al2O3 gate dielectric layer [107]. Copyright © 2014, John Wiley and Sons. (b) Change in the μ and NSS of C8-BTBT FETs according to the thickness of the polymeric layer in p(CEA-co-DEGDVE)/Al2O3 hybrid dielectric [164]. Black and blue lines represent mobility and trap density, respectively. Copyright © 2022, John Wiley and Sons. (c) ID change with respect to the bias stress time of TIPS-pentacene/PTAA FETs with the CYTOP/HfO2-Al2O3 dielectric layer [165]. Copyright © 2018, American Association for the Advancement of Science. (d) The schematic diagram of flexible DPP-DTT FETs based on PSAND-4 dielectrics and (e) their mobility and VT change according to the bending radius [106]. Copyright © 2020, John Wiley and Sons. (f) The schematic diagram of the fabrication process of SWCNT FETs with the BaTiO3/PDMS composite dielectric layer and (g) the optical microscopy image of the fabricated device. (h) The change in μ and Ion/Ioff with respect to the applied strain and strain cycles of SWCNT FETs [166]. Red and blue dots represent mobility and Ion/Ioff in the upper graph, respectively, and green and yellow dots represent mobility and Ion/Ioff in the lower graph, respectively. Copyright © 2016, American Chemical Society.
Micromachines 15 01115 g005

3.3. Triboelectric Nanogenerators (TENGs)

3.3.1. Overview of Polymer Dielectric Materials for TENGs

Various types of dielectric materials have been introduced to accommodate the important requirements of TENGs, including high output performance, low operating voltage, and device stability. Directly injecting electric charge into the surface of a dielectric layer significantly improves the surface charge density and enhances the output performance of TENGs, adding a gradient charge-confinement layer containing mesoporous carbon spheres, further boosting these parameters [222,223]. Similarly, using the self-assembled monolayer (SAM) method for the atomic-level chemical functionalization of polymer surfaces significantly alters the dielectric properties and enhances the triboelectric performance of TENGs [224,225]. In addition, incorporating metal oxide nanoparticles into the triboelectric layer increases the dielectric constant and surface charge density, further improving the output performance of TENGs [226,227]. Dissolving P(VDF-TrFE) in high dipole moment solvents enhances the polymer chain length, crystallinity, and dipole alignment, leading to superior triboelectric properties and improved TENG output performance [228,229].
Moreover, the application of electrical polarization to ferroelectric composite layers by controlling the electric field, temperature, and process time improves the ferroelectric properties and allows the precise modulation of charge transfer and output voltage in TENGs [230,231]. Furthermore, multilayered PVDF-TrFE/BaTiO3 films with aligned BaTiO3 nanoparticles significantly enhance the dielectric constant, local field, and polarization, resulting in a higher output current density and improved TENG performance [27]. The subsequent section will review the latest research breakthroughs in enhancing polymer-based dielectric materials specifically for TENGs. Table 4 provides a summary of the parameters for the TENG devices that utilize polymer dielectric materials, which are the focus of this paper.

3.3.2. Polymer-Based Dielectric Materials and Strategies for Enhancing TENG Devices

Directly injecting electric charge into the surface of a dielectric layer is a straightforward yet highly effective method for greatly increasing surface charge density. Wang et al. demonstrated that the output performance of TENGs could be enhanced by using ionized air to inject charges into a fluorinated ethylene propylene (FEP) film with an air ionization gun (Figure 6a) [232]. The electret film created through ion implantation maintained a surface charge density (Δσ) of approximately 200 μC·m−2 for 160 days, with a minimal loss rate of 16.6%. The TENG with an ionized FEP film exhibited a five-fold increase in the short-circuit current density compared to the untreated FEP film (Figure 6b,c).
In another study, Cha et al. introduced a charge injection process to a gradient charge-confinement layer composed of electrospun nanofibers to enhance the surface charge density of the dielectric layer [233,234,235]. Each successive nanofibrous layer, from the innermost to the outermost, contained increasingly larger amounts of mesoporous carbon spheres (mCS), as illustrated in Figure 6d. These mCSs facilitated the transport of charges from the outer surface, influenced by the charge injection, to the deeper regions of the nanofibrous layer while also minimizing charge loss through confinement. This gradient arrangement of mCSs resulted in a high space charge density within the charge-injected composite layer (Figure 6e). When the charge injection was carried out under a strong external field of 7 kV, the surface charge density in the gradient charge-confinement layer was approximately 7.5 times greater than in the absence of mCSs. As a result, the output voltage surged dramatically to 600 V after the high-voltage charge injection compared to just 15.2 V before the injection (Figure 6f). In summary, the combination of the charge injection with the strategic structural design of the dielectric layer presents an effective approach for creating high-performance TENGs.
The SAM method [225,236,237], which relies on the spontaneous organization of chemisorbed surfactant molecules on solid surfaces, serves as another effective method for chemical surface modification. Shin et al. introduced a straightforward SAM approach to achieve atomic-scale chemical functionalization aimed at modifying the dielectric properties of polymer surfaces (Figure 6g). In their study, SAMs were methodically formed on a polyethylene terephthalate (PET) substrate through the sequential application of various halogens and amines. The process began with the functionalization of the PET surface using hydroxyl groups (−OH) via oxygen plasma treatment, creating an intermediate layer that promotes chemical bonding with desired functional molecules. Subsequently, this hydroxyl-coated layer was treated with halogen-terminated (Br, F, and Cl) phenyl derivatives, acting as tribo-negative materials, and different amine compounds, serving as tribo-positive materials, to impart negative or positive triboelectric characteristics to the PET substrate accordingly. The resulting aminated PET substrates demonstrated strong tribo-positive behavior, while the halogenated PET substrates exhibited pronounced tribo-negative properties. Constructing TENGs by pairing samples with opposing triboelectric polarities, such as PEI(b)-PET and Cl-PET, yielded high voltage outputs of 300 V and 200 V for each layer, respectively. Overall, this atomic-level chemical functionalization technique provides a simple and efficient strategy to enhance the design versatility of polymer substrates in the development of high-performance TENGs.
Another effective strategy for achieving high-performance TENGs is to directly adjust the intrinsic dielectric constant of the active triboelectric layer, which maximizes dielectric polarization across the triboelectric layers. This can be accomplished by creating a triboelectric layer in the form of a nanocomposite that incorporates high-k metal oxide nanoparticles (such as BaTiO3, BiFeO3, and CCTO) [238,239,240,241] dispersed within a ferroelectric polymer matrix, thereby effectively increasing the dielectric constant. These high-k nanoparticles boost the surface charge density of the triboelectric layer, which in turn enhances the output performance of TENGs.
Kim et al. developed a composite triboelectric layer composed of butylated melamine formaldehyde (BMF) and high-dielectric CaCuTi4O12 (CCTO) nanoparticles designed to produce a high-output TENG with consistent performance (Figure 6j–l). The CCTO nanoparticles, with a high dielectric constant of 7500, induced substantial dielectric polarization in the composite triboelectric layer, leading to more efficient charge induction at the counter electrode and boosting the TENG’s output performance. The dielectric constant of the BMF-CCTO composite layer (at 1 wt.%) was measured to be 21.74, which was approximately three times higher than that of pristine BMF, BMF-Al2O3, and BMF-TiO2. The relationship between permittivity and electrical polarization can be expressed as
P = ε 0 ε r 1 E
where P , ε r , and E are the electrical polarization within a material, relative permittivity, and applied electric field, respectively [242]. According to Equation (6), a higher dielectric constant leads to stronger internal polarization when subjected to an electric field. Incorporating 1 wt.% CCTO into BMF generated a voltage of 268 V and a current density of 25.8 mA/m2 [60]. In addition, ferroelectric materials were used to alter dipole alignment, which further boosted the performance of TENGs. The described advancements in enhancing the performance of TENGs highlight the significance of combining innovative charge injection techniques with material engineering. The use of ionized air, gradient charge-confinement layers, and high-dielectric materials demonstrates the potential for significantly increasing surface charge density and output efficiency. These approaches not only push the boundaries of TENG technology but also open up new possibilities for designing high-performance energy-harvesting devices. It could also be broadly applicable to other dielectric-based devices, such as capacitors or sensors, where maximizing charge storage and dielectric polarization is essential.
Figure 6. (a) Schematic diagram of the basic ion injection process onto the FEP film. The short-circuit current density of the TENG (b) before and (c) after the injection process. Reprinted with permission from [243], Copyright 2014, Wiley-VCH. (d) The schematic diagram of charge transport and confinement mechanisms in different charge confinement layers. (e) The comparison of the output voltages based on the gradient arrangement of mCSs. (f) The comparison of the surface charge densities based on the gradient arrangement of mCSs. Reprinted with permission from [223], Copyright 2022, Wiley-VCH. (g) The schematic diagram of surface-functionalized negative and positive PETs with adopted molecules. (h,i) The surface charge density for various functional groups. Reprinted with permission from [244], Copyright 2017, American Chemical Society. (j) The PE curves of pure BMF and BMF–CCTO 1 wt.% composite material (inset: chemical structure of CCTO and the relationship between relative permittivity and polarization under an electric field). (k) The relative permittivity of pure BMF, BMF–Al2O3, BMF–TiO2, and BMF–CCTO composite materials (1 wt.%). (l) The VRMS and JRMS of pure BMF, BMF–Al2O3, BMF–TiO2, and BMF–CCTO (1 wt.%) composite materials in rotation-type freestanding mode TENGs. Reprinted with permission from [60], Copyright 2020, Wiley-VCH.
Figure 6. (a) Schematic diagram of the basic ion injection process onto the FEP film. The short-circuit current density of the TENG (b) before and (c) after the injection process. Reprinted with permission from [243], Copyright 2014, Wiley-VCH. (d) The schematic diagram of charge transport and confinement mechanisms in different charge confinement layers. (e) The comparison of the output voltages based on the gradient arrangement of mCSs. (f) The comparison of the surface charge densities based on the gradient arrangement of mCSs. Reprinted with permission from [223], Copyright 2022, Wiley-VCH. (g) The schematic diagram of surface-functionalized negative and positive PETs with adopted molecules. (h,i) The surface charge density for various functional groups. Reprinted with permission from [244], Copyright 2017, American Chemical Society. (j) The PE curves of pure BMF and BMF–CCTO 1 wt.% composite material (inset: chemical structure of CCTO and the relationship between relative permittivity and polarization under an electric field). (k) The relative permittivity of pure BMF, BMF–Al2O3, BMF–TiO2, and BMF–CCTO composite materials (1 wt.%). (l) The VRMS and JRMS of pure BMF, BMF–Al2O3, BMF–TiO2, and BMF–CCTO (1 wt.%) composite materials in rotation-type freestanding mode TENGs. Reprinted with permission from [60], Copyright 2020, Wiley-VCH.
Micromachines 15 01115 g006

3.3.3. Ferroelectric Polymers for TENG Devices

Figure 7a shows P(VDF-TrFE) dissolved in a solvent with a high dipole moment, leading to longer polymer chain lengths, improved chain orientation, and increased crystallinity. The enhanced crystallinity of P(VDF-TrFE) in these solvents is mainly attributed to the extended end-to-end chain length of the polymer [245,246]. Kim et al. determined the Mw values using gel permeation chromatography (GPC) because the relative chain length of P(VDF-TrFE) can be expressed in terms of the weight average molecular weight (Mw). The GPC of P(VDF-TrFE) dissolved in tetrahydrofuran (THF), methyl ethyl ketone (MEK), dimethylformamide (DMF), and dimethyl sulfoxide (DMSO) revealed 2.37 × 10⁵, 2.44 × 10⁵, 2.75 × 10⁵, and 2.76 × 10⁵ g·mol−1, respectively (Figure 7b). These results confirm that P(VDF-TrFE) dissolved in higher dipole moment solvents (e.g., DMSO) has a higher Mw and a longer chain length. The alignment of dipoles plays a crucial role in the performance of ferroelectric polymer-based TENGs, as greater dipole alignment in the polymer can lead to increased output. To evaluate improvements in energy harvesting, the triboelectric properties of P(VDF-TrFE) dissolved in four different solvents were analyzed. Figure 7c illustrates the contact potential differences (CPDs) of P(VDF-TrFE) in these solvents, measured using Kelvin probe force microscopy (KPFM). The CPDs between the P(VDF-TrFE) and the Pt tip were −0.87, −1.10, −1.63, and −2.37 V for THF, MEK, DMF, and DMSO, respectively. Based on these CPD measurements, the corresponding work functions of P(VDF-TrFE) dissolved in THF, MEK, DMF, and DMSO were calculated to be 5.91, 6.14, 6.67, and 7.41 eV, respectively. Consequently, P(VDF-TrFE) dissolved in a solvent with a higher dipole moment exhibits a more negative CPD and a larger work function, suggesting a greater capacity to attract electrons during triboelectrification and enhanced triboelectric properties [247].
Beyond controlling chain length, ferroelectric composite layers can enhance the ferroelectric properties of friction layers by triggering dipole rearrangement through the application of external electric fields [230,248]. Achieving the effective polarization of each composite requires optimizing factors such as the electric field, temperature, and processing time, which are influenced by the material’s coercive field, insulation resistance, and Curie temperature [249]. Additionally, the polarization direction must be carefully selected to strengthen the polarity of the friction surface (Figure 7d). The output voltages of positively poled P(VDF-TrFE)-based TENGs were opposite in sign to those of negatively poled P(VDF-TrFE)-based TENGs, suggesting that the polarization direction influences the sign of charge transfer during the triboelectric phenomenon.
Moreover, although the bare P(VDF-TrFE) device showed an output voltage of approximately 15 V, positively and negatively poled P(VDF-TrFE)-based TENGs under a vertical compressive force of 2 kg exhibited output voltages of up to 400 V and 100 V, respectively. These values are significantly higher than those of the bare P(VDF-TrFE)-based TENGs. This can be explained by the incomplete switching of the bare P(VDF-TrFE) in one direction. The varying output voltages also confirmed that the amount of charge transfer can be controlled by adjusting the polarization states. Figure 7d,e show that the sign and amount of charge transfer can be controlled by the polarization states.
Park et al. evaluated the triboelectric performance of three different film types to assess the impact of BaTiO3 (BTO) nanoparticle (NP) interlayers on triboelectric properties: a single PVDF-TrFE film, a single PVDF-TrFE/BTO nanocomposite, and multilayered PVDF-TrFE/BTO films (Figure 7f). The distribution of the electric field indicated that the multilayer structure, with BTO NPs arranged on the coplanar layer, offers more efficiently connected interfacial charges at closer distances than composites with randomly dispersed BTO NPs. This arrangement leads to a significantly enhanced local electric field and a greater ferroelectric polarization of the polymer [250].
Simulations using COMSOL Multiphysics revealed that the interfacial polarization in the multilayer structure (8.4 mC·m−2) was greater than in the single composite (8.06 mC·m−2). This increased polarization contributes to a higher dielectric constant in the multilayered film. Figure 7g illustrates how the dielectric constant varies with frequency across different films. While the PVDF-TrFE/BTO films exhibited a higher dielectric constant (15.9 at 10 kHz) compared to the PVDF-TrFE films (13.9 at 10 kHz) due to the presence of high-k BTO nanoparticles, the multilayered PVDF-TrFE/BTO film achieved the highest dielectric constant (17.06 at 10 kHz), attributed to the alignment and closer spacing of BTO nanoparticles in the coplanar layer. The polymer interlayers between the inorganic nanoparticle layers in these multilayered films also enhanced breakdown strength by reducing leakage current in the insulating polymer layers [251], resulting in superior dielectric properties compared to single composites. Figure 7h displays the average output current density for PVDF-TrFE and PVDF-TrFE/BTO films in both the single composite and multilayer configurations. The multilayered PVDF-TrFE/BTO films demonstrated an output current density of 1.77 μA·cm−2, which is 1.5 times higher than that of the composite film without a multilayer structure (1.2 μA·cm−2). The enhanced output performance of the multilayered PVDF-TrFE/BTO films, in comparison to single composite films, is due to the alternating PVDF-TrFE/BTO layer structure, which increases stress-induced polarization. Furthermore, the stronger local field generated by effective interfacial polarization within the multilayered PVDF-TrFE/BTO film also played a significant role in boosting the output performance.
Figure 7. Schematic diagram of the ferroelectric P(VDF-TrFE) to modulate output performance and related methods. (a) The schematic diagram of P(VDF-TrFE) dissolved in a high dipole moment solvent and the resulting P(VDF-TrFE) film. (b) The GPC measurement of the molecular weight (Mw) of P(VDF-TrFE) dissolved in four solvents (THF, MEK, DMF, and DMSO); a higher Mw indicates a longer chain length of P(VDF-TrFE). (c) The CPDs of P(VDF-TrFE) dissolved in the four solvents; a bias was applied to the central area of P(VDF-TrFE) for poling. Reproduced from [247], Copyright 2017, WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim. (d,e) Controllable charge transfer through ferroelectric polarization. Reproduced from [252], Copyright 2016, WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim. (f) The schematic diagram of three different sample types with the same thickness: a PVDF-TrFE film, a single PVDF-TrFE/BTO composite, and a multi-layered PVDF-TrFE/BTO film. (g) The comparison of the dielectric constants of the three different types of films. (h) The comparison of the output current density under a vertical pressure of 98 kPa at 2 Hz. Reproduced from [250], Copyright 2020, American Chemical Society.
Figure 7. Schematic diagram of the ferroelectric P(VDF-TrFE) to modulate output performance and related methods. (a) The schematic diagram of P(VDF-TrFE) dissolved in a high dipole moment solvent and the resulting P(VDF-TrFE) film. (b) The GPC measurement of the molecular weight (Mw) of P(VDF-TrFE) dissolved in four solvents (THF, MEK, DMF, and DMSO); a higher Mw indicates a longer chain length of P(VDF-TrFE). (c) The CPDs of P(VDF-TrFE) dissolved in the four solvents; a bias was applied to the central area of P(VDF-TrFE) for poling. Reproduced from [247], Copyright 2017, WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim. (d,e) Controllable charge transfer through ferroelectric polarization. Reproduced from [252], Copyright 2016, WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim. (f) The schematic diagram of three different sample types with the same thickness: a PVDF-TrFE film, a single PVDF-TrFE/BTO composite, and a multi-layered PVDF-TrFE/BTO film. (g) The comparison of the dielectric constants of the three different types of films. (h) The comparison of the output current density under a vertical pressure of 98 kPa at 2 Hz. Reproduced from [250], Copyright 2020, American Chemical Society.
Micromachines 15 01115 g007
Overall, electrically polarizing ferroelectric composite materials allows for the adjustment of the dielectric properties and the triboelectric friction surface, thereby enhancing TENG performance. Moreover, the polarized dipoles within the composite layer influence charge carrier movement, leading to an additional increase in the triboelectric output [252,253,254]. The findings highlight the importance of solvent dipole moments and controlled polarization in optimizing the dielectric properties and output performance of ferroelectric polymers in TENGs. By carefully selecting solvents and employing strategic polarization techniques, the molecular structure and dipole alignment of materials like P(VDF-TrFE) can be precisely controlled, leading to substantial enhancements in triboelectric performance. These insights extend beyond TENGs, offering a pathway for improving other dielectric-based devices where charge carrier movement and dielectric constant optimization are important, such as in capacitors and memory devices. The demonstrated benefits of multilayered structures and aligned nanoparticles suggest that similar strategies could be employed to enhance the efficiency and performance of a wide range of dielectric materials and devices.

3.3.4. Polymer–Inorganic Hybrid Dielectric Materials for TENG Devices

Beyond surface and bulk modifications, engineering intermediate layers presents a promising strategy for enhancing triboelectric performance. These interlayer modifications are introduced between the triboelectric materials and electrodes, allowing for control over the interfacial properties based on the physical attributes of the interlayer materials. Two primary challenges include surface charge decay and the hindrance of electrostatic induction due to charge recombination at the electrodes [255]. Surface charges in contact with air can gradually dissipate because their electric field attracts ions that are constantly present in the atmosphere [256]. Furthermore, surface charges can decrease due to a drift process caused by the electric field and a diffusion process driven by the electron concentration gradient. This reduction in surface charges leads to a decline in the triboelectric performance. Electrostatic induction can also be hindered at the interface between the triboelectric materials and electrodes, where the presence of numerous free electrons in the electrodes may neutralize electrostatic induction through recombination [257]. The suppression of electrostatic induction has a detrimental effect on triboelectric behavior, resulting in lower performance. To address these challenges and enhance triboelectric efficiency, various interfacial engineering techniques have been thoroughly researched. This section examines interfacial engineering strategies, highlighting two primary approaches: electron-trapping layers (ETLs) and electron-blocking layers (EBLs).
Kim et al. introduced an innovative design for a TENG incorporating a polymeric interlayer. In this study, electrospun PVDF-TrFE served as the negative tribo-material, while PDMS was used as the charge-trapping layer [258,259] (Figure 8a). The design featured an interfused charge-trapping layer intended to capture triboelectric charges within the PDMS interlayer (Figure 8b). This interfused structure improved the contact between PVDF-TrFE and PDMS, thereby enhancing charge transfer from PVDF-TrFE to PDMS. As a result, the charge-trapping efficiency of the PDMS layer was significantly increased, leading to a triboelectric performance boost of approximately two times compared to a PVDF-TrFE-based TENG and six times higher than a TENG with a simply bonded PVDF-TrFE/PDMS structure (Figure 8c).
Table 4. Summary of polymer dielectric-based TENGs.
Table 4. Summary of polymer dielectric-based TENGs.
Dielectric
Material
Working ModeSurface Charge
Density
Output VoltageOutput
Current
Output
Power
Applied Force, Pressure[Ref.]
FEPSingle electrode~200
μC·m−2
~1000 VPP~78 mA/m2
(IPP)
315 W/m220 N[243]
mCSs 1Single electrode~75
μC·m−2
~600 VPP12.8 μA
(IPP)
5.83 mW80 N[223]
PEI(b)-PET 2Double electrode52
μC m−2
~520 VPP110 mA/m2
(IPP)
55 W/m20.15 MPa[244]
BMF-CCTO 3Free-standingN/A268 VRMS25.8 mA/m2
(IRMS)
25.8 W/m2N/A[60]
PVDF-TrFE
(DMSO)
Double electrodeN/A~340 VPP~220 μA
(IPP)
N/A1 kgf[247]
PVDF-TrFE
(poled)
Double electrode20.86
nC·m−2
~400 VampN/AN/A1 kgf[252]
PVDF-TrFE/BTODouble electrodeN/A~44 Vamp1.77 μA/cm2
(Iamp)
29.4 μW/cm298 kPa[250]
PVDF-TrFE/PDMSDouble electrodeN/A~40 VPP~350 μA
(IPP)
125.5 mW/cm2N/A[260]
PI/rGO 4Double electrodeN/A190 Vamp~70 μA
(Iamp)
6.3 W/m2N/A[261]
PDMS/TiO2Double electrode30
μC·m−2
272 Vamp~9.1 μA
(Iamp)
N/A5 N[257]
1 mesoporous carbon spheres (mCSs). 2 branched polyethylenimine (PEI(b)-PET) functionalized on the O2 plasma-treated PET. 3 butylated melamine formaldehyde (BMF)–CaCu3Ti4O12 (CCTO). 4 reduced graphene oxide (rGO) on PI.
Researchers have investigated advanced charge-trapping layers made from conductive materials and polymer composites. Wu et al. demonstrated improved triboelectric performance using a reduced graphene oxide (rGO)-based electron-trapping layer (ETL) (Figure 8d). Polyimide (PI) was employed as the negative tribo-material, with a PI/rGO composite layer integrated within the PI layer to function as an electron storage layer. Additionally, rGO adjusted the interfacial energy band alignment (Figure 8e), enhancing its ability to trap electrons. Electron trapping resulted in more than three times improvement in the triboelectric performance compared to a pristine PI-based TENG device (Figure 8f).
Electron-blocking layers (EBLs) have been researched as a means to reduce electrostatic induction hindrance at the interface between triboelectric materials and electrodes. Park et al. proposed an EBL by incorporating a sputtered titanium oxide (TiOx) layer (Figure 8g). PDMS served as the friction layer and an Al film as the electrode in this setup. The electron-blocking mechanism relies on the capacity of oxygen vacancies in the TiOx layer to capture free electrons (Figure 8h). These oxygen vacancies, which act as electrically positive defects, can trap free electrons from the Al electrode when an external electric field is applied. The efficiency of the TiOx layer in blocking electrons is closely tied to the quantity of oxygen vacancies, highlighting the importance of increasing the ratio of these vacancies in the interlayer. Moreover, the interlayer affects both the electron-blocking capabilities and the device’s capacitance. As the material thickness increases, both the capacitance and electrostatic induction may decrease, making it crucial to achieve an optimal thickness for the interlayer. A 25-fold enhancement in the triboelectric performance was achieved using the electron-blocking effect (Figure 8i).
Figure 8. The designs of charge-trapping layers and their impact on triboelectric performance. (a) The illustration of the TENG device featuring an interfused charge-trapping layer. (b) The FE-SEM results of an electrospun PVDF layer and a PVDF/PDMS interfused friction layer. (c) The dependency of the thickness of the interlayer and configuration on the output voltage. Reprinted with permission from [260], Copyright 2022, Elsevier. (d,e) The schematic diagram of the PI/rGO multilayer TENG. (f) The measured open-circuit voltages (VOC) of the TENG based on the PI/rGO multilayer. Reproduced from [261], Copyright 2017, Elsevier. (g) The device design of an interface-engineered TENG based on the intermediate layer (TiOx) and (h) the description regarding the electron block event by the TiOx interlayer. (i) Measured voltages as a function of the thickness of the intermediate layer. Reprinted with permission from [257], Copyright 2018, Elsevier.
Figure 8. The designs of charge-trapping layers and their impact on triboelectric performance. (a) The illustration of the TENG device featuring an interfused charge-trapping layer. (b) The FE-SEM results of an electrospun PVDF layer and a PVDF/PDMS interfused friction layer. (c) The dependency of the thickness of the interlayer and configuration on the output voltage. Reprinted with permission from [260], Copyright 2022, Elsevier. (d,e) The schematic diagram of the PI/rGO multilayer TENG. (f) The measured open-circuit voltages (VOC) of the TENG based on the PI/rGO multilayer. Reproduced from [261], Copyright 2017, Elsevier. (g) The device design of an interface-engineered TENG based on the intermediate layer (TiOx) and (h) the description regarding the electron block event by the TiOx interlayer. (i) Measured voltages as a function of the thickness of the intermediate layer. Reprinted with permission from [257], Copyright 2018, Elsevier.
Micromachines 15 01115 g008

4. Challenges

4.1. Challenges in the Area of Dielectric Polymers for Future Memory Technologies

Polymer dielectric materials have been used extensively in memory devices, specifically in developing conductive filament memristors and charge-trap memory transistors. These applications leverage the unique electrical properties, processability, and flexibility of polymers to enhance device performance. On the other hand, several challenges and opportunities exist for polymer dielectric materials in memory applications:
(1)
Electrical Property Improvement: Polymers have shown high dielectric constants and reliable insulating properties, but their electrical characteristics are often not as advanced as those of inorganic materials. Enhancing the dielectric constant and minimizing the leakage current paths, even at extremely thin thicknesses, requires molecular-level material and structural design. For example, incorporating polar functional groups that respond effectively to electric fields can improve the dielectric constant [262,263]. In addition, forming a denser polymer matrix can help reduce the leakage currents [97].
(2)
Process Optimization: Polymer dielectric films are primarily fabricated through solution-based processing methods, which offer simplicity and potential for large-scale production [96,264]. On the other hand, these methods can introduce residual solvents or additives that adversely affect electrical properties [80]. Exploring alternative processing techniques, such as chemical vapor deposition (CVD), can help achieve high-purity polymer films and homogeneous mixing, improving the overall performance of memory devices [51,55,85,86,95].
Despite these challenges, the flexibility, tunability, and ease of the processing of polymer dielectric materials make them promising candidates for next-generation memory devices. Ongoing research aim to address these hurdles and unlock the full potential of polymer dielectrics in advanced memory applications.

4.2. Challenges in the Area of Dielectric Polymers for Future FETs

Polymer dielectric materials have been used extensively in FET devices based on emerging semiconductors, as highlighted in this paper. Polymer materials have been designed to have high dielectric constant and reliable insulating performance to fabricate high-performance and low-power FETs. Some polymers can exhibit unique characteristics such as dopant properties, barrier properties, or ferroelectricity. Polymers have also been used as organic moieties in hybrid dielectrics, enhancing the FET performance, operational stability, or mechanical deformability. Nevertheless, there are still challenges and opportunities for polymer dielectric materials in FETs, as follows:
(1)
The electrical properties should be further improved. Although a high dielectric constant and reliable insulating properties have been secured in some polymers, polymer dielectric materials are still considered inferior to inorganic counterparts in their electrical characteristics. Overcoming these limitations and developing polymer dielectric layers with high capacitance will need molecular-level material and structural design [5,36]. For example, increasing the dielectric constant by allowing polar functional groups to respond to the applied electric field may be necessary [114,117,121,126], and ferroelectric polymer can be a representative example of this design [171,182,188]. In addition, it is essential to densify the polymer matrix to minimize the leakage current path even at extremely thin thicknesses [115,124,127,128].
(2)
Appropriate material design and process optimization are necessary. Polymer dielectric films are mostly fabricated through solution-based processing methods, which are advantageous in terms of process simplicity and potential printability [47,118,166]. On the other hand, solution processes can affect the electrical properties of polymer dielectric layers due to residual solvents or additives. Research has been performed on polymer dielectric materials based on chemical vapor deposition processes such as parylene [31,32,265]. Nevertheless, these processes require vacuum equipment and cost. Furthermore, reliable and uniform processability on a large scale should also be ensured to expand the practical application potential for polymer dielectric materials [9,50,164].
Despite these challenges, polymer dielectric materials are expected to be highly valuable because of their unique characteristics, such as tunability in chemistry and excellent mechanical properties, particularly in form factor-free, future electronics.

4.3. Challenges in the Area of Dielectric Polymers for Future TENGs

Polymer dielectric materials are used widely in TENGs to convert mechanical energy into electrical energy. These materials have been engineered to enhance dielectric properties and charge-trapping capabilities, which are essential for high-performance TENGs. Polymers, such as PVDF and its copolymers, have been popular because of their high dielectric constant and flexibility, allowing for efficient energy harvesting and mechanical adaptability. On the other hand, there are still challenges and opportunities for polymer dielectric materials in TENGs, as follows:
(1)
Efficiency and output performance: Although some polymer dielectrics show promising triboelectric properties, achieving higher efficiency and output performance remains challenging. Enhancing the dielectric constant and surface charge density through material innovation and surface engineering is critical to improving the efficiency of TENGs. For example, integrating high-dielectric nanoparticles, such as BaTiO3 or TiO2, into the polymer matrix can significantly boost the output performance. In addition, producing nanocomposite structures that incorporate conductive fillers like Au nanoparticles can further enhance charge transport and storage capabilities, increasing the overall energy conversion efficiency of TENGs [266].
(2)
Durability and wear resistance: TENGs are subject to repeated mechanical contact and friction, which can degrade polymer materials over time. Developing polymers with high mechanical durability and wear resistance is essential for ensuring long-term reliability and consistent performance [267]. Innovations can enhance durability, such as crosslinking polymer chains to produce a more robust network or adding wear-resistant additives. For example, incorporating silicone-based elastomers or other flexible yet tough materials can reduce surface wear and maintain the integrity of the triboelectric layers. Furthermore, applying protective coatings that resist abrasion and environmental degradation can help preserve the functional properties of the polymer dielectric materials in TENGs.
(3)
Environmental stability: TENGs often operate under various environmental conditions, such as humidity, temperature fluctuations, and UV exposure. Ensuring that polymer dielectrics maintain their performance under these conditions requires the development of environmentally robust materials. Polymers with inherent environmental stability, such as fluorinated polymers or those that can be chemically modified to resist harsh conditions, will be crucial for the reliable operation of TENGs in diverse settings. For example, surface modification with hydrophobic coatings can prevent moisture absorption [268], while UV stabilizers protect the material from photodegradation. In addition, selecting polymers that have a broad operational temperature range can ensure consistent performance in both high- and low-temperature environments.
Despite these challenges, polymer dielectric materials are expected to be highly valuable for TENG applications because of their unique properties and potential for innovative designs. The ability to tailor polymer properties at the molecular level offers significant opportunities for optimizing TENG performance. Ongoing research and development are needed to overcome these hurdles and unlock the full potential of polymer dielectrics in next-generation energy-harvesting devices.

5. Conclusions and Outlook

Polymer dielectric materials have potential applications in memory devices, FETs, and TENGs. The integration and synergy among these devices can lead to innovative solutions and advancements in flexible and wearable electronics, self-powered systems, and sustainable technologies.

5.1. Interconnected Functionalities

(1)
Memory devices and FETs: Polymer dielectric materials can be used to develop flexible, low-power memory devices that integrate seamlessly with FETs. The high dielectric constants and reliable insulating properties of polymers enable the development of memory–transistor hybrid systems that combine data storage with logic functions, such as logic-in-memory systems. This integration can result in compact, high-performance devices capable of performing complex computations while retaining data, all within a flexible form factor.
(2)
TENGs and FETs: TENGs can be self-powered sources for FET-based electronic circuits and sensors. By converting mechanical energy from movements or environmental vibrations into electrical energy, TENGs can provide a sustainable power supply for FETs, eliminating the need for external batteries. This integration is beneficial for wearable sensors and remote monitoring devices where replacing or recharging batteries is impractical.

5.2. Specific Integration Strategies in Artificial Intelligence of Things (AIoT) Devices

Integrated systems that combine TENGs, FETs, and memory devices can play a crucial role in AIoT applications where continuous and reliable power supply is essential. These advanced systems can harness energy from environmental sources using TENGs, ensuring a sustainable power supply without external batteries. The harvested energy can power FETs for data processing and memory devices for data storage, enabling smart sensors and devices to function autonomously.
For example, smart infrastructure sensors can use TENGs to convert mechanical energy from vibrations or movements into electrical energy. This energy can power FETs that process real-time data and memory devices that store historical data. These AIoT devices can then use artificial intelligence algorithms to analyze the data, predict maintenance needs, and optimize performance. This integration allows for real-time decision making and adaptive responses to changing conditions, enhancing the efficiency and reliability of smart infrastructure.
In agricultural settings, AIoT devices can monitor soil conditions, weather patterns, and crop health. TENGs can harvest energy from wind, rain, or mechanical movements, powering the sensors and processors needed to collect and analyze data. Memory devices can store large datasets for longitudinal analysis, while AI algorithms can provide insights and recommendations for improving crop yields and resource management.
Wearable health monitors in the AIoT ecosystem can benefit from this integration by harvesting energy from the user’s movements. TENGs can generate power from everyday activities, enabling the continuous monitoring of vital signs without frequent recharging. The FETs process the health data in real time, and the memory devices store critical health information. AI algorithms can analyze data to detect anomalies, provide health recommendations, and alert users or healthcare providers to potential issues.
AIoT systems can achieve a high level of autonomy and intelligence by combining TENGs, FETs, and memory devices, operating efficiently in various environments and applications. This integration enhances the capabilities of IoT devices, making them more sustainable, intelligent, and responsive to real-world conditions.

Author Contributions

W.C. and J.C. performed the literature survey and research analysis. Y.H. reviewed the manuscript. H.Y. and H.-J.Y. supervised the work. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Research Foundation of Korea (NRF) grant funded by the Korean government (MSIT) (RS-2023-00253474).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Li, Q.; Chen, L.; Gadinski, M.R.; Zhang, S.; Zhang, G.; Li, H.U.; Iagodkine, E.; Haque, A.; Chen, L.-Q.; Jackson, T.N. Flexible high-temperature dielectric materials from polymer nanocomposites. Nature 2015, 523, 576–579. [Google Scholar] [CrossRef]
  2. Seong, H.; Pak, K.; Joo, M.; Choi, J.; Im, S.G. Vapor-phase deposited ultrathin polymer gate dielectrics for high-performance organic thin film transistors. Adv. Electron. Mater. 2016, 2, 1500209. [Google Scholar] [CrossRef]
  3. Wang, X.; Wang, X.; Pi, M.; Ran, R. High-strength, highly conductive and woven organic hydrogel fibers for flexible electronics. J. Chem. Eng. 2022, 428, 131172. [Google Scholar] [CrossRef]
  4. Zhang, Z.; Du, C.; Jiao, H.; Zhang, M. Polyvinyl alcohol/SiO2 hybrid dielectric for transparent flexible/stretchable all-carbon-nanotube thin-film-transistor integration. Adv. Electron. Mater. 2020, 6, 1901133. [Google Scholar] [CrossRef]
  5. Choi, J.; Kang, J.; Lee, C.; Jeong, K.; Im, S.G. Heavily crosslinked, high-k ultrathin polymer dielectrics for flexible, low-power organic thin-film transistors (OTFTs). Adv. Electron. Mater. 2020, 6, 2000314. [Google Scholar] [CrossRef]
  6. Hung, C.C.; Wu, H.C.; Chiu, Y.C.; Tung, S.H.; Chen, W.C. Crosslinkable high dielectric constant polymer dielectrics for low voltage organic field-effect transistor memory devices. J. Polym. Sci. Part A Polym. Chem. 2016, 54, 3224–3236. [Google Scholar] [CrossRef]
  7. Wang, S.; Yang, C.; Li, X.; Jia, H.; Liu, S.; Liu, X.; Minari, T.; Sun, Q. Polymer-based dielectrics with high permittivity and low dielectric loss for flexible electronics. J. Mater. Chem. C 2022, 10, 6196–6221. [Google Scholar] [CrossRef]
  8. Zheng, X.; Pu, Z.; Hu, L.; Tian, Y.; Xia, J.; Cheng, J.; Zhong, J. Synthesis and properties of high performance polysulfone resin with low dielectric constant and dielectric loss. J. Mater. Sci. Mater. Electron. 2019, 30, 18168–18176. [Google Scholar] [CrossRef]
  9. Kim, W.; Kwon, J.; Takeda, Y.; Sekine, T.; Tokito, S.; Jung, S. Flexible and printed organic nonvolatile memory transistor with bilayer polymer dielectrics. Adv. Mater. Technol. 2021, 6, 2100141. [Google Scholar] [CrossRef]
  10. Wang, K.; Liu, J.; El-Khouly, M.E.; Cui, X.; Che, Q.; Zhang, B.; Chen, Y. Water-soluble polythiophene-conjugated polyelectrolyte-based memristors for transient electronics. ACS Appl. Mater. Interfaces 2022, 14, 36987–36997. [Google Scholar] [CrossRef]
  11. Woo, M.H.; Jang, B.C.; Choi, J.; Lee, K.J.; Shin, G.H.; Seong, H.; Im, S.G.; Choi, S.Y. Low-power nonvolatile charge storage memory based on MoS2 and an ultrathin polymer tunneling dielectric. Adv. Funct. Mater. 2017, 27, 1703545. [Google Scholar] [CrossRef]
  12. Han, Y.; Lee, S.; Kim, M.; Shin, W.; Lee, H.k.; Koo, R.H.; Lee, S.T.; Kim, C.H.; Yoo, H. Charge Transport Advancement in Anti-Ambipolar Transistors: Spatially Separating Layer Sandwiched between N-Type Metal Oxides and P-Type Small Molecules. Adv. Funct. Mater. 2024, 34, 2316217. [Google Scholar] [CrossRef]
  13. Lee, S.-H.; Xu, Y.; Khim, D.; Park, W.-T.; Kim, D.-Y.; Noh, Y.-Y. Effect of polymer gate dielectrics on charge transport in carbon nanotube network transistors: Low-k insulator for favorable active interface. ACS Appl. Mater. Interfaces 2016, 8, 32421–32431. [Google Scholar] [CrossRef] [PubMed]
  14. Seo, J.; Kim, C.-H.; Yoo, H. Dual-Gate Operation and Configurable Logic From Solution Pattern-Based Zinc Tin Oxide Thin-Film Transistors. IEEE Trans. Electron. Devices 2024, 71, 3020–3025. [Google Scholar] [CrossRef]
  15. Walser, M.; Kalb, W.; Mathis, T.; Brenner, T.; Batlogg, B. Stable complementary inverters with organic field-effect transistors on Cytop fluoropolymer gate dielectric. Appl. Phys. Lett. 2009, 94, 053303. [Google Scholar] [CrossRef]
  16. Fan, F.-R.; Tian, Z.-Q.; Wang, Z.L. Flexible triboelectric generator. Nano Energy 2012, 1, 328–334. [Google Scholar] [CrossRef]
  17. Hinchet, R.; Seung, W.; Kim, S.W. Recent progress on flexible triboelectric nanogenerators for selfpowered electronics. ChemSusChem 2015, 8, 2327–2344. [Google Scholar] [CrossRef] [PubMed]
  18. An, J.; Wang, Z.M.; Jiang, T.; Liang, X.; Wang, Z.L. Whirling-folded triboelectric nanogenerator with high average power for water wave energy harvesting. Adv. Funct. Mater. 2019, 29, 1904867. [Google Scholar] [CrossRef]
  19. Yang, H.; Wang, Z.; Guo, X.; Su, H.; Sun, K.; Yang, D.; Xiao, W.; Wang, Q.; He, D. Controlled growth of fine multifilaments in polymer-based memristive devices via the conduction control. ACS Appl. Mater. Interfaces 2020, 12, 34370–34377. [Google Scholar] [CrossRef]
  20. Chou, Y.-H.; Chang, H.-C.; Liu, C.-L.; Chen, W.-C. Polymeric charge storage electrets for non-volatile organic field effect transistor memory devices. Polym. Chem. 2015, 6, 341–352. [Google Scholar] [CrossRef]
  21. Hsu, L.-C.; Isono, T.; Lin, Y.-C.; Kobayashi, S.; Chiang, Y.-C.; Jiang, D.-H.; Hung, C.-C.; Ercan, E.; Yang, W.-C.; Hsieh, H.-C. Stretchable OFET memories: Tuning the morphology and the charge-trapping ability of conjugated block copolymers through soft segment branching. ACS Appl. Mater. Interfaces 2021, 13, 2932–2943. [Google Scholar] [CrossRef] [PubMed]
  22. Zhang, L.-X.; Gao, X.; Lv, J.-J.; Zhong, Y.-N.; Xu, C.; Xu, J.-L.; Wang, S.-D. Filter-free selective light monitoring by organic field-effect transistor memories with a tunable blend charge-trapping layer. ACS Appl. Mater. Interfaces 2019, 11, 40366–40371. [Google Scholar] [CrossRef] [PubMed]
  23. Kwon, H.j.; Ye, H.; Shim, K.; Girma, H.G.; Tang, X.; Lim, B.; Kim, Y.; Lee, J.; Park, C.E.; Jung, S.H. Newly synthesized nonvacuum processed high-k polymeric dielectrics with carboxyl functionality for highly stable operating printed transistor applications. Adv. Funct. Mater. 2021, 31, 2007304. [Google Scholar] [CrossRef]
  24. Liu, Z.; Yin, Z.; Chen, S.-C.; Dai, S.; Huang, J.; Zheng, Q. Binary polymer composite dielectrics for flexible low-voltage organic field-effect transistors. Org. Electron. 2018, 53, 205–212. [Google Scholar] [CrossRef]
  25. Rahmanudin, A.; Tate, D.J.; Marcial-Hernandez, R.; Bull, N.; Garlapati, S.K.; Zamhuri, A.; Khan, R.U.; Faraji, S.; Gollu, S.R.; Persaud, K.C. Robust high-capacitance polymer gate dielectrics for stable low-voltage organic field-effect transistor sensors. Adv. Electron. Mater. 2020, 6, 1901127. [Google Scholar] [CrossRef]
  26. Davies, D. Charge generation on dielectric surfaces. J. Phys. D Appl. Phys. 1969, 2, 1533. [Google Scholar] [CrossRef]
  27. Seung, W.; Yoon, H.J.; Kim, T.Y.; Ryu, H.; Kim, J.; Lee, J.H.; Lee, J.H.; Kim, S.; Park, Y.K.; Park, Y.J. Boosting power-generating performance of triboelectric nanogenerators via artificial control of ferroelectric polarization and dielectric properties. Adv. Energy Mater. 2017, 7, 1600988. [Google Scholar] [CrossRef]
  28. Wan, Q.; Chen, Q.; Freithaler, M.A.; Velagala, S.R.; Liu, Y.; To, A.C.; Mahajan, A.; Mukkamala, R.; Xiong, F. Toward Real-Time Blood Pressure Monitoring via High-Fidelity Iontronic Tonometric Sensors with High Sensitivity and Large Dynamic Ranges. Adv. Healthc. Mater. 2023, 12, 2202461. [Google Scholar] [CrossRef]
  29. Song, J.; Liu, H.; Zhao, Z.; Lin, P.; Yan, F. Flexible organic transistors for biosensing: Devices and applications. Adv. Mater. 2024, 36, 2300034. [Google Scholar] [CrossRef]
  30. Chen, X.; Han, X.; Shen, Q.D. PVDF-based ferroelectric polymers in modern flexible electronics. Adv. Electron. Mater. 2017, 3, 1600460. [Google Scholar] [CrossRef]
  31. Kwon, J.; Takeda, Y.; Shiwaku, R.; Tokito, S.; Cho, K.; Jung, S. Three-dimensional monolithic integration in flexible printed organic transistors. Nat. Commun. 2019, 10, 54. [Google Scholar] [CrossRef]
  32. Stucchi, E.; Dell’Erba, G.; Colpani, P.; Kim, Y.H.; Caironi, M. Low-voltage, printed, all-polymer integrated circuits employing a low-leakage and high-yield polymer dielectric. Adv. Electron. Mater. 2018, 4, 1800340. [Google Scholar] [CrossRef]
  33. Wu, X.; Wang, S.; Huang, W.; Dong, Y.; Wang, Z.; Huang, W. Wearable in-sensor reservoir computing using optoelectronic polymers with through-space charge-transport characteristics for multi-task learning. Nat. Commun. 2023, 14, 468. [Google Scholar] [CrossRef] [PubMed]
  34. Dai, Y.; Hu, H.; Wang, M.; Xu, J.; Wang, S. Stretchable transistors and functional circuits for human-integrated electronics. Nat. Electron. 2021, 4, 17–29. [Google Scholar] [CrossRef]
  35. Wang, W.; Jiang, Y.; Zhong, D.; Zhang, Z.; Choudhury, S.; Lai, J.-C.; Gong, H.; Niu, S.; Yan, X.; Zheng, Y. Neuromorphic sensorimotor loop embodied by monolithically integrated, low-voltage, soft e-skin. Science 2023, 380, 735–742. [Google Scholar] [CrossRef]
  36. Wang, B.; Huang, W.; Chi, L.; Al-Hashimi, M.; Marks, T.J.; Facchetti, A. High-k gate dielectrics for emerging flexible and stretchable electronics. Chem. Rev. 2018, 118, 5690–5754. [Google Scholar] [CrossRef]
  37. Kang, J.; Yoon, J.; Lee, B.; Jung, H.; Kim, J.; Nam, W.; Jeong, K.; Choi, J.; Son, D.; Im, S.G. Ultrathin, solvent-resistant dielectric for monolithic fabrication of low-power, intrinsically stretchable active-matrix electronic skin. Device 2024, 2, 100426. [Google Scholar] [CrossRef]
  38. Xiong, Y.; Luo, L.; Yang, J.; Han, J.; Liu, Y.; Jiao, H.; Wu, S.; Cheng, L.; Feng, Z.; Sun, J.; et al. Scalable spinning, winding, and knitting graphene textile TENG for energy harvesting and human motion recognition. Nano Energy 2023, 107, 108137. [Google Scholar] [CrossRef]
  39. Cheng, P.; Dai, S.; Liu, Y.; Li, Y.; Hayashi, H.; Papani, R.; Su, Q.; Li, N.; Dai, Y.; Liu, W.; et al. An intrinsically stretchable power-source system for bioelectronics. Device 2024, 2, 100216. [Google Scholar] [CrossRef]
  40. Hu, S.; Chang, S.; Xiao, G.; Lu, J.; Gao, J.; Zhang, Y.; Tao, Y. A Stretchable Multimode Triboelectric Nanogenerator for Energy Harvesting and Self-Powered Sensing. Adv. Mater. Technol. 2021, 7, 2100870. [Google Scholar] [CrossRef]
  41. Anju, V.; Narayanankutty, S.K. High dielectric constant polymer nanocomposite for embedded capacitor applications. Mater. Sci. Eng. B 2019, 249, 114418. [Google Scholar] [CrossRef]
  42. Thakur, Y.; Zhang, T.; Iacob, C.; Yang, T.; Bernholc, J.; Chen, L.; Runt, J.; Zhang, Q. Enhancement of the dielectric response in polymer nanocomposites with low dielectric constant fillers. Nanoscale 2017, 9, 10992–10997. [Google Scholar] [CrossRef]
  43. Zhang, Q.; Chen, X.; Zhang, B.; Zhang, T.; Lu, W.; Chen, Z.; Liu, Z.; Kim, S.H.; Donovan, B.; Warzoha, R.J. High-temperature polymers with record-high breakdown strength enabled by rationally designed chain-packing behavior in blends. Matter 2021, 4, 2448–2459. [Google Scholar] [CrossRef]
  44. Kang, S.; Kim, S.; Yoo, H. Light-Assisted/Light-Driven Memory Behaviors with Small Molecule-Fluoropolymer-Small Molecule-Stacked Floating-Gate Heterostructures. ACS Photonics 2023, 10, 1310–1319. [Google Scholar] [CrossRef]
  45. Ren, H.; Cui, N.; Tang, Q.; Tong, Y.; Zhao, X.; Liu, Y. High-performance, ultrathin, ultraflexible organic thin-film transistor array via solution process. Small 2018, 14, 1801020. [Google Scholar] [CrossRef] [PubMed]
  46. Xu, W.; Hu, Z.; Liu, H.; Lan, L.; Peng, J.; Wang, J.; Cao, Y. Flexible all-organic, all-solution processed thin film transistor array with ultrashort channel. Sci. Rep. 2016, 6, 29055. [Google Scholar] [CrossRef] [PubMed]
  47. Chung, S.; Jang, M.; Ji, S.B.; Im, H.; Seong, N.; Ha, J.; Kwon, S.K.; Kim, Y.H.; Yang, H.; Hong, Y. Flexible High-Performance All-Inkjet-Printed Inverters: Organo-Compatible and Stable Interface Engineering. Adv. Mater. 2013, 25, 4773–4777. [Google Scholar] [CrossRef]
  48. Conti, S.; Lai, S.; Cosseddu, P.; Bonfiglio, A. An Inkjet-Printed, Ultralow Voltage, Flexible Organic Field Effect Transistor. Adv. Mater. Technol. 2017, 2, 1600212. [Google Scholar] [CrossRef]
  49. Kim, I.; Ju, B.; Zhou, Y.; Li, B.M.; Jur, J.S. Microstructures in all-inkjet-printed textile capacitors with bilayer interfaces of polymer dielectrics and metal–organic decomposition silver electrodes. ACS Appl. Mater. Interfaces 2021, 13, 24081–24094. [Google Scholar] [CrossRef]
  50. Yang, S.C.; Choi, J.; Jang, B.C.; Hong, W.; Shim, G.W.; Yang, S.Y.; Im, S.G.; Choi, S.Y. Large-scale, low-power nonvolatile memory based on few-layer MoS2 and ultrathin polymer dielectrics. Adv. Electron. Mater. 2019, 5, 1800688. [Google Scholar] [CrossRef]
  51. Kim, N.; Oh, J.; Kim, S.; Cha, J.H.; Choi, J.; Im, S.G.; Choi, S.Y.; Jang, B.C. Realistic Chest X-Ray Image Synthesis via Generative Network with Stochastic Memristor Array for Machine Learning-Based Medical Diagnosis. Adv. Funct. Mater. 2024, 34, 2305136. [Google Scholar] [CrossRef]
  52. Han, J.S.; Le, Q.V.; Choi, J.; Hong, K.; Moon, C.W.; Kim, T.L.; Kim, H.; Kim, S.Y.; Jang, H.W. Air-stable cesium lead iodide perovskite for ultra-low operating voltage resistive switching. Adv. Funct. Mater. 2018, 28, 1705783. [Google Scholar] [CrossRef]
  53. Zhang, Z.C.; Li, Y.; Li, J.; Chen, X.D.; Yao, B.W.; Yu, M.X.; Lu, T.B.; Zhang, J. An ultrafast nonvolatile memory with low operation voltage for high-speed and low-power applications. Adv. Funct. Mater. 2021, 31, 2102571. [Google Scholar] [CrossRef]
  54. Hung, C.C.; Chiu, Y.C.; Wu, H.C.; Lu, C.; Bouilhac, C.; Otsuka, I.; Halila, S.; Borsali, R.; Tung, S.H.; Chen, W.C. Conception of stretchable resistive memory devices based on nanostructure-controlled carbohydrate-block-polyisoprene block copolymers. Adv. Funct. Mater. 2017, 27, 1606161. [Google Scholar] [CrossRef]
  55. Lee, S.; Seong, H.; Im, S.G.; Moon, H.; Yoo, S. Organic flash memory on various flexible substrates for foldable and disposable electronics. Nat. Commun. 2017, 8, 725. [Google Scholar] [CrossRef] [PubMed]
  56. Choi, J.; Yoo, H. Combination of Polymer Gate Dielectric and Two-Dimensional Semiconductor for Emerging Field-Effect Transistors. Polymers 2023, 15, 1395. [Google Scholar] [CrossRef] [PubMed]
  57. Guo, S.; Wang, Z.; Chen, X.; Li, L.; Li, J.; Ji, D.; Li, L.; Hu, W. Low-voltage polymer-dielectric-based organic field-effect transistors and applications. Nano Sel. 2022, 3, 20–38. [Google Scholar] [CrossRef]
  58. Kim, S.; Seo, J.; Choi, J.; Yoo, H. Vertically integrated electronics: New opportunities from emerging materials and devices. Nanomicro Lett. 2022, 14, 201. [Google Scholar] [CrossRef]
  59. Yu, H.; Chen, Y.; Wei, H.; Gong, J.; Xu, W. High-k polymeric gate insulators for organic field-effect transistors. Nanotechnology 2019, 30, 202002. [Google Scholar] [CrossRef]
  60. Kim, J.; Ryu, H.; Lee, J.H.; Khan, U.; Kwak, S.S.; Yoon, H.J.; Kim, S.W. High permittivity CaCu3Ti4O12 particle-induced internal polarization amplification for high performance triboelectric nanogenerators. Adv. Energy Mater. 2020, 10, 1903524. [Google Scholar] [CrossRef]
  61. Zhou, Y.S.; Wang, S.; Yang, Y.; Zhu, G.; Niu, S.; Lin, Z.-H.; Liu, Y.; Wang, Z.L. Manipulating nanoscale contact electrification by an applied electric field. Nano Lett. 2014, 14, 1567–1572. [Google Scholar] [CrossRef] [PubMed]
  62. Chen, Q.; Lin, M.; Wang, Z.; Zhao, X.; Cai, Y.; Liu, Q.; Fang, Y.; Yang, Y.; He, M.; Huang, R. Low power parylene-based memristors with a graphene barrier layer for flexible electronics applications. Adv. Electron. Mater. 2019, 5, 1800852. [Google Scholar] [CrossRef]
  63. Jang, B.C.; Kim, S.; Yang, S.Y.; Park, J.; Cha, J.-H.; Oh, J.; Choi, J.; Im, S.G.; Dravid, V.P.; Choi, S.-Y. Polymer analog memristive synapse with atomic-scale conductive filament for flexible neuromorphic computing system. Nano Lett. 2019, 19, 839–849. [Google Scholar] [CrossRef]
  64. Baeg, K.-J.; Noh, Y.-Y.; Kim, D.-Y. Charge transfer and trapping properties in polymer gate dielectrics for non-volatile organic field-effect transistor memory applications. Solid State Electron. 2009, 53, 1165–1168. [Google Scholar] [CrossRef]
  65. Feng, Y.; Gao, X.; Zhong, Y.-N.; Wu, J.-L.; Xu, J.-L.; Wang, S.-D. Solution-processed polymer thin-film memristors with an electrochromic feature and frequency-dependent synaptic plasticity. Adv. Intell. Syst. 2019, 1, 1900022. [Google Scholar] [CrossRef]
  66. Forrest, S.R. The path to ubiquitous and low-cost organic electronic appliances on plastic. Nature 2004, 428, 911–918. [Google Scholar] [CrossRef]
  67. Park, H.L.; Kim, M.H.; Lee, S.H. Introduction of interfacial load polymeric layer to organic flexible memristor for regulating conductive filament growth. Adv. Electron. Mater. 2020, 6, 2000582. [Google Scholar] [CrossRef]
  68. Scott, J.C.; Bozano, L.D. Nonvolatile memory elements based on organic materials. Adv. Mater. 2007, 19, 1452–1463. [Google Scholar] [CrossRef]
  69. Gao, Y.; Zhang, Y.; Wang, X.; Sim, K.; Liu, J.; Chen, J.; Feng, X.; Xu, H.; Yu, C. Moisture-triggered physically transient electronics. Sci. Adv. 2017, 3, e1701222. [Google Scholar] [CrossRef]
  70. Fu, K.K.; Wang, Z.; Dai, J.; Carter, M.; Hu, L. Transient electronics: Materials and devices. Chem. Mater. 2016, 28, 3527–3539. [Google Scholar] [CrossRef]
  71. Oh, S.; Kim, H.; Kim, S.E.; Kim, M.-H.; Park, H.-L.; Lee, S.-H. Biodegradable and flexible polymer-based memristor possessing optimized synaptic plasticity for eco-friendly wearable neural networks with high energy efficiency. Adv. Intell. Syst. 2023, 5, 2200272. [Google Scholar] [CrossRef]
  72. Lee, S.H.; Park, H.L.; Keum, C.M.; Lee, I.H.; Kim, M.H.; Lee, S.D. Organic flexible memristor with reduced operating voltage and high stability by interfacial control of conductive filament growth. Phys. Status Solidi-Rapid Res. Lett. 2019, 13, 1900044. [Google Scholar] [CrossRef]
  73. Dorigato, A.; Pegoretti, A. Biodegradable single-polymer composites from polyvinyl alcohol. Colloid Polym. Sci. 2012, 290, 359–370. [Google Scholar] [CrossRef]
  74. Ahmed, H.; Hashim, A. Geometry optimization, optical and electronic characteristics of novel PVA/PEO/SiC structure for electronics applications. Silicon 2021, 13, 2639–2644. [Google Scholar] [CrossRef]
  75. Fortunato, E.; Correia, N.; Barquinha, P.; Pereira, L.; Gonçalves, G.; Martins, R. High-performance flexible hybrid field-effect transistors based on cellulose fiber paper. IEEE Electron Device Lett. 2008, 29, 988–990. [Google Scholar] [CrossRef]
  76. Yang, D.; Yang, H.; Guo, X.; Zhang, H.; Jiao, C.; Xiao, W.; Guo, P.; Wang, Q.; He, D. Robust polyethylenimine electrolyte for high performance and thermally stable atomic switch memristors. Adv. Funct. Mater. 2020, 30, 2004514. [Google Scholar] [CrossRef]
  77. Zhang, W.; Wang, C.; Liu, G.; Zhu, X.; Chen, X.; Pan, L.; Tan, H.; Xue, W.; Ji, Z.; Wang, J. Thermally-stable resistive switching with a large ON/OFF ratio achieved in poly (triphenylamine). Chem. Commun. 2014, 50, 11856–11858. [Google Scholar] [CrossRef]
  78. Krishnan, K.; Aono, M.; Tsuruoka, T. Thermally stable resistive switching of a polyvinyl alcohol-based atomic switch. J. Mater. Chem. C 2018, 6, 6460–6464. [Google Scholar] [CrossRef]
  79. Li, J.; Qian, Y.; Li, W.; Yu, S.; Ke, Y.; Qian, H.; Lin, Y.H.; Hou, C.H.; Shyue, J.J.; Zhou, J. Polymeric Memristor Based Artificial Synapses with Ultra-Wide Operating Temperature. Adv. Mater. 2023, 35, 2209728. [Google Scholar] [CrossRef]
  80. Moon, H.; Seong, H.; Shin, W.C.; Park, W.-T.; Kim, M.; Lee, S.; Bong, J.H.; Noh, Y.-Y.; Cho, B.J.; Yoo, S. Synthesis of ultrathin polymer insulating layers by initiated chemical vapour deposition for low-power soft electronics. Nat. Mater. 2015, 14, 628–635. [Google Scholar] [CrossRef]
  81. Gohil, J.; Bhattacharya, A.; Ray, P. Studies on the crosslinking of poly (vinyl alcohol). J. Polym. Res. 2006, 13, 161–169. [Google Scholar] [CrossRef]
  82. Xu, X.; Zhou, X.; Wang, T.; Shi, X.; Liu, Y.; Zuo, Y.; Xu, L.; Wang, M.; Hu, X.; Yang, X. Robust DNA-Bridged Memristor for Textile Chips. Angew. Chem. Int. Ed. 2020, 59, 12762–12768. [Google Scholar] [CrossRef] [PubMed]
  83. Zhang, X.; Wu, C.; Lv, Y.; Zhang, Y.; Liu, W. High-performance flexible polymer memristor based on stable filamentary switching. Nano Lett. 2022, 22, 7246–7253. [Google Scholar] [CrossRef]
  84. Che, Q.; Wang, X.; El-Khouly, M.E.; Li, G.; Liu, J.; Zhang, B.; Chen, Y. Improving the Long-Term Stability of BPQD-Based Memory Device via Modification with Polyvinylpyrrolidone-Grafted Polydopamine. Adv. Electron. Mater. 2022, 8, 2101057. [Google Scholar] [CrossRef]
  85. Jang, B.C.; Seong, H.; Kim, S.K.; Kim, J.Y.; Koo, B.J.; Choi, J.; Yang, S.Y.; Im, S.G.; Choi, S.-Y. Flexible nonvolatile polymer memory array on plastic substrate via initiated chemical vapor deposition. ACS Appl. Mater. Interfaces 2016, 8, 12951–12958. [Google Scholar] [CrossRef]
  86. Bae, H.; Jang, B.C.; Park, H.; Jung, S.-H.; Lee, H.M.; Park, J.-Y.; Jeon, S.-B.; Son, G.; Tcho, I.-W.; Yu, K. Functional circuitry on commercial fabric via textile-compatible nanoscale film coating process for fibertronics. Nano Lett. 2017, 17, 6443–6452. [Google Scholar] [CrossRef]
  87. Lee, S.; Kim, S.; Yoo, H. Contribution of polymers to electronic memory devices and applications. Polymers 2021, 13, 3774. [Google Scholar] [CrossRef]
  88. Chen, M.-N.; Chang, S.-W.; Prakoso, S.P.; Li, Y.-T.; Chen, K.-L.; Chiu, Y.-C. Unveiling the photoinduced recovery mystery in conjugated polymer-based transistor memory. ACS Appl. Mater. Interfaces 2021, 13, 44656–44662. [Google Scholar] [CrossRef] [PubMed]
  89. Shih, C.-C.; Chiang, Y.-C.; Hsieh, H.-C.; Lin, Y.-C.; Chen, W.-C. Multilevel photonic transistor memory devices using conjugated/insulated polymer blend electrets. ACS Appl. Mater. Interfaces 2019, 11, 42429–42437. [Google Scholar] [CrossRef] [PubMed]
  90. Liu, Z.; Yu, T.; Wan, Z.; Wang, Y.; Li, Z.; Yin, J.; Gao, X.; Xia, Y.; Liu, Z. High-Performance Pentacene-Based Field-Effect Transistor Memory Using the Electrets of Polymer Blends. Adv. Electron. Mater. 2022, 8, 2101342. [Google Scholar] [CrossRef]
  91. Lin, C.F.; Lin, Y.C.; Yang, W.C.; Hsu, L.C.; Ercan, E.; Hung, C.C.; Yu, Y.Y.; Chen, W.C. Multiband Photoresponding Field-Effect Transistor Memory Using Conjugated Block Copolymers with Pendent Isoindigo Coils as a Polymer Electret. Adv. Electron. Mater. 2021, 7, 2100655. [Google Scholar] [CrossRef]
  92. Chi, H.-Y.; Hsu, H.-W.; Tung, S.-H.; Liu, C.-L. Nonvolatile organic field-effect transistors memory devices using supramolecular block copolymer/functional small molecule nanocomposite electret. ACS Appl. Mater. Interfaces 2015, 7, 5663–5673. [Google Scholar] [CrossRef]
  93. Shu, S.; Xu, T.; Su, J. A flexible floating-gate based organic field-effect transistor non-volatile memory based on F8BT/PMMA integrated floating-gate/tunneling layer. Phys. Scr. 2023, 98, 045402. [Google Scholar] [CrossRef]
  94. He, W.; Xu, W.; He, H.; Jing, X.; Liu, C.; Feng, J.; Luo, C.; Fan, Z.; Wu, S.; Gao, J. From Unipolar, WORM-Type to Ambipolar, Bistable Organic Electret Memory Device by Controlling Minority Lateral Transport. Adv. Electron. Mater. 2020, 6, 1901320. [Google Scholar] [CrossRef]
  95. Pak, K.; Choi, J.; Lee, C.; Im, S.G. Low-power, flexible nonvolatile organic transistor memory based on an ultrathin bilayer dielectric stack. Adv. Electron. Mater. 2019, 5, 1800799. [Google Scholar] [CrossRef]
  96. Li, Z.; An, L.; Khuje, S.; Tan, J.; Hu, Y. Solution-shearing of dielectric polymer with high thermal conductivity and electric insulation. Sci. Adv. 2021, 7, eabi7410. [Google Scholar] [CrossRef]
  97. Yoo, S.; Kim, D.-G.; Ha, T.; Chan Won, J.; Jang, K.-S.; Kim, Y.H. Solution-processable, thin, and high-κ dielectric polyurea gate insulator with strong hydrogen bonding for low-voltage organic thin-film transistors. ACS Appl. Mater. Interfaces 2018, 10, 32462–32470. [Google Scholar] [CrossRef] [PubMed]
  98. Kim, J.Y.; Choi, M.-J.; Jang, H.W. Ferroelectric field effect transistors: Progress and perspective. APL Mater. 2021, 9, 021102. [Google Scholar] [CrossRef]
  99. Chen, R.S.; Lu, Y. Negative capacitance field effect transistors based on van der Waals 2D materials. Small 2023, 2304445. [Google Scholar] [CrossRef]
  100. Nketia-Yawson, B.; Noh, Y.Y. Recent progress on high-capacitance polymer gate dielectrics for flexible low-voltage transistors. Adv. Funct. Mater. 2018, 28, 1802201. [Google Scholar] [CrossRef]
  101. Kang, B.; Song, E.; Lee, S.B.; Chae, B.-G.; Ahn, H.; Cho, K. Stretchable polymer gate dielectric with segmented elastomeric network for organic soft electronics. Chem. Mater. 2018, 30, 6353–6360. [Google Scholar] [CrossRef]
  102. Koo, J.H.; Kang, J.; Lee, S.; Song, J.-K.; Choi, J.; Yoon, J.; Park, H.J.; Sunwoo, S.-H.; Kim, D.C.; Nam, W. A vacuum-deposited polymer dielectric for wafer-scale stretchable electronics. Nat. Electron. 2023, 6, 137–145. [Google Scholar] [CrossRef]
  103. Yoo, H.; Heo, K.; Ansari, M.H.R.; Cho, S. Recent advances in electrical doping of 2D semiconductor materials: Methods, analyses, and applications. Nanomaterials 2021, 11, 832. [Google Scholar] [CrossRef]
  104. Park, H.; Choi, J.; Shim, J.; Lee, S.M.; On, S.; Yun, H.J.; Kim, S.; Im, S.G.; Yoo, H. Functional polymeric passivation-led improvement of bias stress with long-term durability of edge-rich nanoporous MoS2 thin-film transistors. npj 2D Mater. Appl. 2022, 6, 21. [Google Scholar] [CrossRef]
  105. Jeong, J.W.; Hwang, H.S.; Choi, D.; Ma, B.C.; Jung, J.; Chang, M. Hybrid polymer/metal oxide thin films for high performance, flexible transistors. Micromachines 2020, 11, 264. [Google Scholar] [CrossRef]
  106. Chen, Y.; Zhuang, X.; Goldfine, E.A.; Dravid, V.P.; Bedzyk, M.J.; Huang, W.; Facchetti, A.; Marks, T.J. Printable organic-inorganic nanoscale multilayer gate dielectrics for thin-film transistors enabled by a polymeric organic interlayer. Adv. Funct. Mater. 2020, 30, 2005069. [Google Scholar] [CrossRef]
  107. Feng, W.; Zheng, W.; Cao, W.; Hu, P. Back gated multilayer InSe transistors with enhanced carrier mobilities via the suppression of carrier scattering from a dielectric interface. Adv. Mater. 2014, 26, 6587–6593. [Google Scholar] [CrossRef]
  108. Kwon, H.j.; Ye, H.; Baek, Y.; Hong, J.; Wang, R.; Choi, Y.; Lee, I.; Park, C.E.; Nam, S.; Kim, J. Printable ultra-flexible fluorinated organic–inorganic nanohybrid sol–gel derived gate dielectrics for highly stable organic thin-film transistors and other practical applications. Adv. Funct. Mater. 2021, 31, 2009539. [Google Scholar] [CrossRef]
  109. Kim, M.J.; Jeong, J.; Lee, T.I.; Kim, J.; Tak, Y.; Park, H.; Im, S.G.; Cho, B.J. Large-Area, Conformal, and Uniform Synthesis of Hybrid Polymeric Film via Initiated Chemical Vapor Deposition. Macromol. Mater. Eng. 2021, 306, 2000608. [Google Scholar] [CrossRef]
  110. Li, Y.; Wang, H.; Shi, Z.; Mei, J.; Wang, X.; Yan, D.; Cui, Z. Novel high-k polymers as dielectric layers for organic thin-film transistors. Polym. Chem. 2015, 6, 6651–6658. [Google Scholar] [CrossRef]
  111. Li, Y.; Wang, H.; Zhang, C.; Zhang, Y.; Cui, Z.; Yan, D.; Shi, Z. Organic thin-film transistors with novel high-k polymers as dielectric layers. Polym. Chem. 2015, 6, 3685–3693. [Google Scholar] [CrossRef]
  112. Ye, H.; Ryu, K.Y.; Kwon, H.-j.; Lee, H.; Wang, R.; Hong, J.; Choi, H.H.; Nam, S.Y.; Lee, J.; Kong, H. Amorphous Fluorinated Acrylate Polymer Dielectrics for Flexible Transistors and Logic Gates with High Operational Stability. ACS Appl. Mater. Interfaces 2023, 15, 32610–32620. [Google Scholar] [CrossRef]
  113. Zou, J.; Li, S.; Wang, H.; Wang, W.; Shi, Z.; Jiang, Y.; Cui, Z.; Yan, D. High-k polymer materials containing cyclic carbonate as gate dielectrics for application in low-voltage operating organic thin-film transistors. J. Mater. Chem. C 2019, 7, 15357–15363. [Google Scholar] [CrossRef]
  114. Choe, Y.-S.; Yi, M.H.; Kim, J.-H.; Ryu, G.-S.; Noh, Y.-Y.; Kim, Y.H.; Jang, K.-S. Crosslinked polymer-mixture gate insulator for high-performance organic thin-film transistors. Org. Electron. 2016, 36, 171–176. [Google Scholar] [CrossRef]
  115. Choi, J.; Joo, M.; Seong, H.; Pak, K.; Park, H.; Park, C.W.; Im, S.G. Flexible, low-power thin-film transistors made of vapor-phase synthesized high-k, ultrathin polymer gate dielectrics. ACS Appl. Mater. Interfaces 2017, 9, 20808–20817. [Google Scholar] [CrossRef]
  116. Choi, J.; Seong, H.; Pak, K.; Im, S.G. Vapor-phase deposition of the fluorinated copolymer gate insulator for the p-type organic thin-film transistor. J. Inf. Disp. 2016, 17, 43–49. [Google Scholar] [CrossRef]
  117. Choi, J.; Yoon, J.; Kim, M.J.; Pak, K.; Lee, C.; Lee, H.; Jeong, K.; Ihm, K.; Yoo, S.; Cho, B.J. Spontaneous generation of a molecular thin hydrophobic skin layer on a sub-20 nm, high-k polymer dielectric for extremely stable organic thin-film transistor operation. ACS Appl. Mater. Interfaces 2019, 11, 29113–29123. [Google Scholar] [CrossRef]
  118. Feng, L.; Jiang, C.; Ma, H.; Guo, X.; Nathan, A. All ink-jet printed low-voltage organic field-effect transistors on flexible substrate. Org. Electron. 2016, 38, 186–192. [Google Scholar] [CrossRef]
  119. Hallani, R.K.; Moser, M.; Bristow, H.; Jenart, M.V.; Faber, H.; Neophytou, M.; Yarali, E.; Paterson, A.F.; Anthopoulos, T.D.; McCulloch, I. Low-temperature cross-linking benzocyclobutene based polymer dielectric for organic thin film transistors on plastic substrates. J. Org. Chem. 2019, 85, 277–283. [Google Scholar] [CrossRef]
  120. Kim, J.-Y.; Kim, J.W.; Lee, E.K.; Park, J.-I.; Lee, B.-L.; Kwon, Y.-N.; Byun, S.; Jung, M.-S.; Kim, J.-J. Cross-linked poly (hydroxy imide) gate-insulating materials for low-temperature processing of organic thin-film transistors. J. Mater. Chem. C 2018, 6, 13359–13366. [Google Scholar] [CrossRef]
  121. Kwon, J.S.; Park, H.W.; Kim, D.H.; Kwark, Y.-J. Solvent-free processable and photo-patternable hybrid gate dielectric for flexible top-gate organic field-effect transistors. ACS Appl. Mater. Interfaces 2017, 9, 5366–5374. [Google Scholar] [CrossRef] [PubMed]
  122. Lee, E.K.; Kim, J.Y.; Chung, J.W.; Lee, B.-L.; Kang, Y. Photo-crosslinkable polymer gate dielectrics for hysteresis-free organic field-effect transistors with high solvent resistance. RSC Adv. 2014, 4, 293–300. [Google Scholar] [CrossRef]
  123. Li, S.; Feng, L.; Zhao, J.; Guo, X.; Zhang, Q. Low temperature cross-linked, high performance polymer gate dielectrics for solution-processed organic field-effect transistors. Polym. Chem. 2015, 6, 5884–5890. [Google Scholar] [CrossRef]
  124. Sun, W.; Zhao, J.; Chen, S.; Guo, X.; Zhang, Q. Thermally cross-linked polyvinyl alcohol as gate dielectrics for solution processing organic field-effect transistors. Synth. Met. 2019, 250, 73–78. [Google Scholar] [CrossRef]
  125. Yang, H.; Kim, M.-J.; Kang, B.-G.; Lee, B.H. Organic Field-Effect Transistors with Bottlebrush Polymer Gate Dielectrics Thermally Cross-Linked in Less Than 1 min. Chem. Mater. 2021, 33, 6356–6364. [Google Scholar] [CrossRef]
  126. Ye, H.; Kwon, H.j.; Shin, S.C.; Lee, H.y.; Park, Y.H.; Tang, X.; Wang, R.; Lee, K.; Hong, J.; Li, Z. The hidden potential of polysilsesquioxane for high-k: Analysis of the origin of its dielectric nature and practical low-voltage-operating applications beyond the unit device. Adv. Funct. Mater. 2022, 32, 2104030. [Google Scholar] [CrossRef]
  127. Jung, S.; Albariqi, M.; Gruntz, G.; Al-Hathal, T.; Peinado, A.; Garcia-Caurel, E.; Nicolas, Y.; Toupance, T.; Bonnassieux, Y.; Horowitz, G. A TIPS-TPDO-tetraCN-based n-type organic field-effect transistor with a cross-linked PMMA polymer gate dielectric. ACS Appl. Mater. Interfaces 2016, 8, 14701–14708. [Google Scholar] [CrossRef]
  128. Noh, Y.-Y.; Sirringhaus, H. Ultra-thin polymer gate dielectrics for top-gate polymer field-effect transistors. Org. Electron. 2009, 10, 174–180. [Google Scholar] [CrossRef]
  129. Oh, J.Y.; Rondeau-Gagné, S.; Chiu, Y.-C.; Chortos, A.; Lissel, F.; Wang, G.-J.N.; Schroeder, B.C.; Kurosawa, T.; Lopez, J.; Katsumata, T. Intrinsically stretchable and healable semiconducting polymer for organic transistors. Nature 2016, 539, 411–415. [Google Scholar] [CrossRef]
  130. Wang, S.; Xu, J.; Wang, W.; Wang, G.-J.N.; Rastak, R.; Molina-Lopez, F.; Chung, J.W.; Niu, S.; Feig, V.R.; Lopez, J. Skin electronics from scalable fabrication of an intrinsically stretchable transistor array. Nature 2018, 555, 83–88. [Google Scholar] [CrossRef]
  131. Wang, W.; Wang, S.; Rastak, R.; Ochiai, Y.; Niu, S.; Jiang, Y.; Arunachala, P.K.; Zheng, Y.; Xu, J.; Matsuhisa, N. Strain-insensitive intrinsically stretchable transistors and circuits. Nat. Electron. 2021, 4, 143–150. [Google Scholar] [CrossRef]
  132. Xu, J.; Wang, S.; Wang, G.-J.N.; Zhu, C.; Luo, S.; Jin, L.; Gu, X.; Chen, S.; Feig, V.R.; To, J.W. Highly stretchable polymer semiconductor films through the nanoconfinement effect. Science 2017, 355, 59–64. [Google Scholar] [CrossRef] [PubMed]
  133. Xu, J.; Wu, H.-C.; Zhu, C.; Ehrlich, A.; Shaw, L.; Nikolka, M.; Wang, S.; Molina-Lopez, F.; Gu, X.; Luo, S. Multi-scale ordering in highly stretchable polymer semiconducting films. Nat. Mater. 2019, 18, 594–601. [Google Scholar] [CrossRef]
  134. Ho, J.-C.; Lin, Y.-C.; Chen, C.-K.; Hsu, L.-C.; Chen, W.-C. Hydrogel-based sustainable and stretchable field-effect transistors. Org. Electron. 2022, 100, 106358. [Google Scholar] [CrossRef]
  135. Hur, J.S.; Kim, J.O.; Kim, H.A.; Jeong, J.K. Stretchable polymer gate dielectric by ultraviolet-assisted hafnium oxide doping at low temperature for high-performance indium gallium tin oxide transistors. ACS Appl. Mater. Interfaces 2019, 11, 21675–21685. [Google Scholar] [CrossRef]
  136. Kim, J.O.; Hur, J.S.; Kim, D.; Lee, B.; Jung, J.M.; Kim, H.A.; Chung, U.J.; Nam, S.H.; Hong, Y.; Park, K.S. Network structure modification-enabled hybrid polymer dielectric film with zirconia for the stretchable transistor applications. Adv. Funct. Mater. 2020, 30, 1906647. [Google Scholar] [CrossRef]
  137. Lu, C.; Lee, W.-Y.; Shih, C.-C.; Wen, M.-Y.; Chen, W.-C. Stretchable polymer dielectrics for low-voltage-driven field-effect transistors. ACS Appl. Mater. Interfaces 2017, 9, 25522–25532. [Google Scholar] [CrossRef]
  138. Rao, Y.-L.; Chortos, A.; Pfattner, R.; Lissel, F.; Chiu, Y.-C.; Feig, V.; Xu, J.; Kurosawa, T.; Gu, X.; Wang, C. Stretchable self-healing polymeric dielectrics cross-linked through metal–ligand coordination. J. Am. Chem. Soc. 2016, 138, 6020–6027. [Google Scholar] [CrossRef]
  139. Zheng, Y.-Q.; Liu, Y.; Zhong, D.; Nikzad, S.; Liu, S.; Yu, Z.; Liu, D.; Wu, H.-C.; Zhu, C.; Li, J. Monolithic optical microlithography of high-density elastic circuits. Science 2021, 373, 88–94. [Google Scholar] [CrossRef]
  140. Choi, S.; Jo, J.-W.; Kim, J.; Song, S.; Kim, J.; Park, S.K.; Kim, Y.-H. Static and dynamic water motion-induced instability in oxide thin-film transistors and its suppression by using low-k fluoropolymer passivation. ACS Appl. Mater. Interfaces 2017, 9, 26161–26168. [Google Scholar] [CrossRef]
  141. Jia, J.; Jang, S.K.; Lai, S.; Xu, J.; Choi, Y.J.; Park, J.-H.; Lee, S. Plasma-treated thickness-controlled two-dimensional black phosphorus and its electronic transport properties. ACS Nano 2015, 9, 8729–8736. [Google Scholar] [CrossRef] [PubMed]
  142. Lee, H.; Hong, S.; Yoo, H. Interfacial doping effects in fluoropolymer-tungsten diselenide composites providing high-performance P-type transistors. Polymers 2021, 13, 1087. [Google Scholar] [CrossRef]
  143. Nair, S.; Kathiresan, M.; Mukundan, T.; Natarajan, V. Passivation of organic field effect transistor with photopatterned Parylene to improve environmental stability. Microelectron. Eng. 2016, 163, 36–42. [Google Scholar] [CrossRef]
  144. Oh, J.G.; Pak, K.; Kim, C.S.; Bong, J.H.; Hwang, W.S.; Im, S.G.; Cho, B.J. A high-performance top-gated graphene field-effect transistor with excellent flexibility enabled by an iCVD copolymer gate dielectric. Small 2018, 14, 1703035. [Google Scholar] [CrossRef]
  145. Pak, K.; Seong, H.; Choi, J.; Hwang, W.S.; Im, S.G. Synthesis of ultrathin, homogeneous copolymer dielectrics to control the threshold voltage of organic thin-film transistors. Adv. Funct. Mater. 2016, 26, 6574–6582. [Google Scholar] [CrossRef]
  146. Seong, H.; Choi, J.; Jang, B.C.; Kim, M.; Yoo, S.; Choi, S.Y.; Im, S.G. A low-voltage organic complementary inverter with high operation stability and flexibility using an ultrathin iCVD polymer dielectric and a hybrid encapsulation layer. Adv. Electron. Mater. 2016, 2, 1500385. [Google Scholar] [CrossRef]
  147. Shin, E.-S.; Park, W.-T.; Kwon, Y.-W.; Xu, Y.; Noh, Y.-Y. Spontaneous doping at the polymer–polymer interface for high-performance organic transistors. ACS Appl. Mater. Interfaces 2019, 11, 12709–12716. [Google Scholar] [CrossRef] [PubMed]
  148. Tak, Y.J.; Keene, S.T.; Kang, B.H.; Kim, W.-G.; Kim, S.J.; Salleo, A.; Kim, H.J. Multifunctional, room-temperature processable, heterogeneous organic passivation layer for oxide semiconductor thin-film transistors. ACS Appl. Mater. Interfaces 2019, 12, 2615–2624. [Google Scholar] [CrossRef]
  149. Chen, Y.; Huang, W.; Sangwan, V.K.; Wang, B.; Zeng, L.; Wang, G.; Huang, Y.; Lu, Z.; Bedzyk, M.J.; Hersam, M.C. Polymer doping enables a two-dimensional electron gas for high-performance homojunction oxide thin-film transistors. Adv. Mater. 2019, 31, 1805082. [Google Scholar] [CrossRef]
  150. Hong, S.; Yoo, G.; Kim, D.H.; Song, W.G.; Le, O.K.; Hong, Y.K.; Takahashi, K.; Omkaram, I.; Son, D.N.; Kim, S. The doping mechanism and electrical performance of polyethylenimine-doped MoS2 transistor. Phys. Status Solidi C 2017, 14, 1600262. [Google Scholar] [CrossRef]
  151. Huang, W.; Guo, P.; Zeng, L.; Li, R.; Wang, B.; Wang, G.; Zhang, X.; Chang, R.P.; Yu, J.; Bedzyk, M.J. Metal composition and polyethylenimine doping capacity effects on semiconducting metal oxide–polymer blend charge transport. J. Am. Chem. Soc. 2018, 140, 5457–5473. [Google Scholar] [CrossRef]
  152. Huang, W.; Zeng, L.; Yu, X.; Guo, P.; Wang, B.; Ma, Q.; Chang, R.P.; Yu, J.; Bedzyk, M.J.; Marks, T.J. Metal oxide transistors via polyethylenimine doping of the channel layer: Interplay of doping, microstructure, and charge transport. Adv. Funct. Mater. 2016, 26, 6179–6187. [Google Scholar] [CrossRef]
  153. Seo, Y.-M.; Jang, W.; Gu, T.; Whang, D. Highly efficient n-type doping of graphene by vacuum annealed amine-rich macromolecules. Materials 2020, 13, 2166. [Google Scholar] [CrossRef] [PubMed]
  154. Jiang, Y.; Liu, Z.; Yin, Z.; Zheng, Q. Sandwich structured dielectrics for air-stable and flexible low-voltage organic transistors in ultrasensitive pressure sensing. Mater. Chem. Front. 2020, 4, 1459–1470. [Google Scholar] [CrossRef]
  155. Liu, Z.; Yin, Z.; Jiang, Y.; Zheng, Q. Dielectric interface passivation of polyelectrolyte-gated organic field-effect transistors for ultrasensitive low-voltage pressure sensors in wearable applications. Mater. Today Electron. 2022, 1, 100001. [Google Scholar] [CrossRef]
  156. Liu, Z.; Yin, Z.; Li, J.; Wang, Z. Transparent polymer nanoheterostructure films for flexible low-power organic transistors with high mobility, decent photostability, and ultralong-term air stability. Mater. Today Phys. 2023, 37, 101206. [Google Scholar] [CrossRef]
  157. Liu, Z.; Yin, Z.; Wang, J.; Zheng, Q. Polyelectrolyte dielectrics for flexible low-voltage organic thin-film transistors in highly sensitive pressure sensing. Adv. Funct. Mater. 2019, 29, 1806092. [Google Scholar] [CrossRef]
  158. Liu, Z.; Zhao, Y.; Yin, Z. Low-power soft transistors triggering revolutionary electronics. Innovation 2024, 5, 100616. [Google Scholar] [CrossRef]
  159. Yin, Z.; Yin, M.J.; Liu, Z.; Zhang, Y.; Zhang, A.P.; Zheng, Q. Solution-processed bilayer dielectrics for flexible low-voltage organic field-effect transistors in pressure-sensing applications. Adv. Sci. 2018, 5, 1701041. [Google Scholar] [CrossRef] [PubMed]
  160. Laudari, A.; Pickett, A.; Shahedipour-Sandvik, F.; Hogan, K.; Anthony, J.E.; He, X.; Guha, S. Textured Poling of the Ferroelectric Dielectric Layer for Improved Organic Field-Effect Transistors. Adv. Mater. Interfaces 2019, 6, 1801787. [Google Scholar] [CrossRef]
  161. Kwon, H.-j.; Tang, X.; Shin, S.; Hong, J.; Jeong, W.; Jo, Y.; An, T.K.; Lee, J.; Kim, S.H. Facile photo-cross-linking system for polymeric gate dielectric materials toward solution-processed organic field-effect transistors: Role of a cross-linker in various polymer types. ACS Appl. Mater. Interfaces 2020, 12, 30600–30615. [Google Scholar] [CrossRef] [PubMed]
  162. Wang, X.; Chen, Y.; Wu, G.; Li, D.; Tu, L.; Sun, S.; Shen, H.; Lin, T.; Xiao, Y.; Tang, M. Two-dimensional negative capacitance transistor with polyvinylidene fluoride-based ferroelectric polymer gating. npj 2D Mater. Appl. 2017, 1, 38. [Google Scholar] [CrossRef]
  163. Liu, X.; Liang, R.; Gao, G.; Pan, C.; Jiang, C.; Xu, Q.; Luo, J.; Zou, X.; Yang, Z.; Liao, L. MoS2 Negative-Capacitance Field-Effect Transistors with Subthreshold Swing below the Physics Limit. Adv. Mater. 2018, 30, 1800932. [Google Scholar] [CrossRef] [PubMed]
  164. Choi, J.; Lee, C.; Kang, J.; Lee, C.; Lee, S.M.; Oh, J.; Choi, S.Y.; Im, S.G. A sub-20 nm organic/inorganic hybrid dielectric for ultralow-power organic thin-film transistor (OTFT) with enhanced operational stability. Small 2022, 18, 2203165. [Google Scholar] [CrossRef]
  165. 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]
  166. Cai, L.; Zhang, S.; Miao, J.; Yu, Z.; Wang, C. Fully printed stretchable thin-film transistors and integrated logic circuits. ACS Nano 2016, 10, 11459–11468. [Google Scholar] [CrossRef]
  167. Chen, Y.; Wang, X.; Wang, P.; Huang, H.; Wu, G.; Tian, B.; Hong, Z.; Wang, Y.; Sun, S.; Shen, H. Optoelectronic properties of few-layer MoS2 FET gated by ferroelectric relaxor polymer. ACS Appl. Mater. Interfaces 2016, 8, 32083–32088. [Google Scholar] [CrossRef] [PubMed]
  168. Kobayashi, T.; Hori, N.; Nakajima, T.; Kawae, T. Electrical characteristics of MoS2 field-effect transistor with ferroelectric vinylidene fluoride-trifluoroethylene copolymer gate structure. Appl. Phys. Lett. 2016, 108, 132903. [Google Scholar] [CrossRef]
  169. Li, D.; Wang, X.; Chen, Y.; Zhu, S.; Gong, F.; Wu, G.; Meng, C.; Liu, L.; Wang, L.; Lin, T. The ambipolar evolution of a high-performance WSe2 transistor assisted by a ferroelectric polymer. Nanotechnology 2018, 29, 105202. [Google Scholar] [CrossRef]
  170. Tian, B.; Liu, L.; Yan, M.; Wang, J.; Zhao, Q.; Zhong, N.; Xiang, P.; Sun, L.; Peng, H.; Shen, H. A robust artificial synapse based on organic ferroelectric polymer. Adv. Electron. Mater. 2019, 5, 1800600. [Google Scholar] [CrossRef]
  171. Wang, X.; Wang, P.; Wang, J.; Hu, W.; Zhou, X.; Guo, N.; Huang, H.; Sun, S.; Shen, H.; Lin, T. Ultrasensitive and broadband MoS2 photodetector driven by ferroelectrics. Adv. Mater. 2015, 27, 6575–6581. [Google Scholar] [CrossRef] [PubMed]
  172. Wang, Y.; Kizu, T.; Song, L.; Zhang, Y.; Jiang, S.; Qian, J.; Wang, Q.; Shi, Y.; Zheng, Y.; Nabatame, T. High-performance non-volatile field-effect transistor memories using an amorphous oxide semiconductor and ferroelectric polymer. J. Mater. Chem. C 2016, 4, 7917–7923. [Google Scholar] [CrossRef]
  173. Wen, J.; Yan, C.; Sun, Z. The application of a high-κ polymer dielectric in graphene transistors. Adv. Electron. Mater. 2020, 6, 2000031. [Google Scholar] [CrossRef]
  174. Chu, F.J.; Chen, Y.C.; Shih, L.C.; Mao, S.C.; Chen, J.S. Reconfigurable Physical Reservoir Enabled by Polarization of Ferroelectric Polymer P (VDF–TrFE) and Interface Charge-Trapping/Detrapping in Dual-Gate IGZO Transistor. Adv. Funct. Mater. 2024, 34, 2310951. [Google Scholar] [CrossRef]
  175. Jiao, H.; Wang, X.; Wu, S.; Chen, Y.; Chu, J.; Wang, J. Ferroelectric field effect transistors for electronics and optoelectronics. Appl. Phys. Rev. 2023, 10, 011310. [Google Scholar] [CrossRef]
  176. Li, L.; Liu, X.; Guo, J.; Ji, H.; Zhang, F.; Lou, Z.; Qin, L.; Hu, Y.; Hou, Y.; Teng, F. Low-operating-voltage two-dimensional tin perovskite field-effect transistors with multilayer gate dielectrics based on a fluorinated copolymer. J. Phys. Chem. Lett. 2023, 14, 2223–2233. [Google Scholar] [CrossRef]
  177. Luo, Z.D.; Yang, M.M.; Liu, Y.; Alexe, M. Emerging opportunities for 2D semiconductor/ferroelectric transistor-structure devices. Adv. Mater. 2021, 33, 2005620. [Google Scholar] [CrossRef]
  178. Niu, X.; Tian, B.; Zhu, Q.; Dkhil, B.; Duan, C. Ferroelectric polymers for neuromorphic computing. Appl. Phys. Rev. 2022, 9, 021309. [Google Scholar] [CrossRef]
  179. Yang, D.; Moon, Y.; Han, N.; Lee, M.; Beak, J.; Lee, S.-H.; Kim, D.-Y. Solution-processable low-voltage carbon nanotube field-effect transistors with high-k relaxor ferroelectric polymer gate insulator. Nanotechnology 2024, 35, 295202. [Google Scholar] [CrossRef]
  180. Faraji, S.; Hashimoto, T.; Turner, M.L.; Majewski, L.A. Solution-processed nanocomposite dielectrics for low voltage operated OFETs. Org. Electron. 2015, 17, 178–183. [Google Scholar] [CrossRef]
  181. Li, J.; Yan, F. Solution-processable low-voltage and flexible floating-gate memories based on an n-type polymer semiconductor and high-k polymer gate dielectrics. ACS Appl. Mater. Interfaces 2014, 6, 12815–12820. [Google Scholar] [CrossRef] [PubMed]
  182. Perinot, A.; Scuratti, F.; Scaccabarozzi, A.D.; Tran, K.; Salazar-Rios, J.M.; Loi, M.A.; Salvatore, G.; Fabiano, S.; Caironi, M. Solution-Processed Polymer Dielectric Interlayer for Low-Voltage, Unipolar n-Type Organic Field-Effect Transistors. ACS Appl. Mater. Interfaces 2023, 15, 56095–56105. [Google Scholar] [CrossRef] [PubMed]
  183. Ramos, B.; Lopes, M.; Buso, D.; Ternisien, M. Performance enhancement in N-channel organic field-effect transistors using ferroelectric material as a gate dielectric. IEEE Trans. Nanotechnol. 2017, 16, 773–777. [Google Scholar] [CrossRef]
  184. Tang, W.; Zhao, J.; Huang, Y.; Ding, L.; Li, Q.; Li, J.; You, P.; Yan, F.; Guo, X. Bias stress stability improvement in solution-processed low-voltage organic field-effect transistors using relaxor ferroelectric polymer gate dielectric. IEEE Electron. Device Lett. 2017, 38, 748–751. [Google Scholar] [CrossRef]
  185. Thuau, D.; Kallitsis, K.; Ha, S.; Bargain, F.; Soulestin, T.; Pecastaings, G.; Tencé-Girault, S.; Domingues Dos Santos, F.; Hadziioannou, G. High and temperature-independent dielectric constant dielectrics from PVDF-based terpolymer and copolymer blends. Adv. Electron. Mater. 2020, 6, 1901250. [Google Scholar] [CrossRef]
  186. Wu, Y.; Wang, Z.; Yang, L.; Qiao, Y.; Chang, D.; Yan, Y.; Wu, Z.; Hu, Z.; Zhang, J.; Lu, X. Constrain Effect of Charge Traps in Organic Field-Effect Transistors with Ferroelectric Polymer as a Dielectric Interfacial Layer. ACS Appl. Mater. Interfaces 2022, 14, 3095–3102. [Google Scholar] [CrossRef]
  187. Xu, M.; Xiang, L.; Xu, T.; Wang, W.; Xie, W.; Zhou, D. Low-voltage operating flexible ferroelectric organic field-effect transistor nonvolatile memory with a vertical phase separation P (VDF-TrFE-CTFE)/PS dielectric. Appl. Phys. Lett. 2017, 111, 183302. [Google Scholar] [CrossRef]
  188. Adhikari, J.M.; Gadinski, M.R.; Li, Q.; Sun, K.G.; Reyes-Martinez, M.A.; Iagodkine, E.; Briseno, A.L.; Jackson, T.N.; Wang, Q.; Gomez, E.D. Controlling Chain Conformations of High-k Fluoropolymer Dielectrics to Enhance Charge Mobilities in Rubrene Single-Crystal Field-Effect Transistors. Adv. Mater. 2016, 28, 10095–10102. [Google Scholar] [CrossRef]
  189. Shin, E.Y.; Cho, H.J.; Jung, S.; Yang, C.; Noh, Y.Y. A high-k fluorinated P (VDF-TrFE)-g-PMMA gate dielectric for high-performance flexible field-effect transistors. Adv. Funct. Mater. 2018, 28, 1704780. [Google Scholar] [CrossRef]
  190. Shin, Y.J.; Kang, S.J.; Jung, H.J.; Park, Y.J.; Bae, I.; Choi, D.H.; Park, C. Chemically cross-linked thin poly (vinylidene fluoride-co-trifluoroethylene) films for nonvolatile ferroelectric polymer memory. ACS Appl. Mater. Interfaces 2011, 3, 582–589. [Google Scholar] [CrossRef]
  191. Cho, H.; Jin, H.J.; Lee, S.; Jeon, S.; Cho, Y.; Park, S.; Jang, M.; Widiapradja, L.J.; Ryu, D.Y.; Park, J.H. 5 nm Ultrathin Crystalline Ferroelectric P (VDF-TrFE)-Brush Tuned for Hysteresis-Free Sub 60 mV dec−1 Negative-Capacitance Transistors. Adv. Mater. 2023, 35, 2300478. [Google Scholar] [CrossRef]
  192. Jo, J.; Shin, C. Negative capacitance field effect transistor with hysteresis-free sub-60-mV/decade switching. IEEE Electron. Device Lett. 2016, 37, 245–248. [Google Scholar] [CrossRef]
  193. McGuire, F.A.; Cheng, Z.; Price, K.; Franklin, A.D. Sub-60 mV/decade switching in 2D negative capacitance field-effect transistors with integrated ferroelectric polymer. Appl. Phys. Lett. 2016, 109, 093101. [Google Scholar] [CrossRef]
  194. Zhang, H.; Chen, Y.; Ding, S.; Wang, J.; Bao, W.; Zhang, D.W.; Zhou, P. 2D negative capacitance field-effect transistor with organic ferroelectrics. Nanotechnology 2018, 29, 244004. [Google Scholar] [CrossRef] [PubMed]
  195. Ha, Y.-G.; Everaerts, K.; Hersam, M.C.; Marks, T.J. Hybrid gate dielectric materials for unconventional electronic circuitry. Acc. Chem. Res. 2014, 47, 1019–1028. [Google Scholar] [CrossRef]
  196. Ortiz, R.P.; Facchetti, A.; Marks, T.J. High-k organic, inorganic, and hybrid dielectrics for low-voltage organic field-effect transistors. Chem. Rev. 2010, 110, 205–239. [Google Scholar] [CrossRef]
  197. Chen, H.; Zhang, W.; Li, M.; He, G.; Guo, X. Interface engineering in organic field-effect transistors: Principles, applications, and perspectives. Chem. Rev. 2020, 120, 2879–2949. [Google Scholar] [CrossRef] [PubMed]
  198. Lee, H.S.; Shin, J.M.; Jeon, P.J.; Lee, J.; Kim, J.S.; Hwang, H.C.; Park, E.; Yoon, W.; Ju, S.Y.; Im, S. Few-Layer MoS2–Organic Thin-Film Hybrid Complementary Inverter Pixel Fabricated on a Glass Substrate. Small 2015, 11, 2132–2138. [Google Scholar] [CrossRef]
  199. Li, Q.; Li, S.; Yang, D.; Su, W.; Wang, Y.; Zhou, W.; Liu, H.; Xie, S. Designing hybrid gate dielectric for fully printing high-performance carbon nanotube thin film transistors. Nanotechnology 2017, 28, 435203. [Google Scholar] [CrossRef]
  200. Seong, H.; Choi, J.; Kim, B.J.; Park, H.; Im, S.G. Vapor-phase synthesis of sub-15 nm hybrid gate dielectrics for organic thin film transistors. J. Mater. Chem. C 2017, 5, 4463–4470. [Google Scholar] [CrossRef]
  201. Sun, Q.J.; Zhuang, J.; Yan, Y.; Zhou, Y.; Han, S.T.; Zhou, L.; Roy, V.A. Investigation on the mobility and stability in organic thin film transistors consisting of bilayer gate dielectrics. Phys. Status Solidi A 2016, 213, 79–84. [Google Scholar] [CrossRef]
  202. Yoon, M.; Ko, K.R.; Min, S.-W.; Im, S. Polymer/oxide bilayer dielectric for hysteresis-minimized 1 V operating 2D TMD transistors. RSC Adv. 2018, 8, 2837–2843. [Google Scholar] [CrossRef] [PubMed]
  203. Bulgarevich, K.; Sakamoto, K.; Yasuda, T.; Minari, T.; Takeuchi, M. Operational Stability Enhancement of Polymeric Organic Field-Effect Transistors by Amorphous Perfluoropolymers Chemically Anchored to Gate Dielectric Surfaces. Adv. Electron. Mater. 2020, 6, 2000161. [Google Scholar] [CrossRef]
  204. Held, M.; Schießl, S.P.; Miehler, D.; Gannott, F.; Zaumseil, J. Polymer/metal oxide hybrid dielectrics for low voltage field-effect transistors with solution-processed, high-mobility semiconductors. Appl. Phys. Lett. 2015, 107, 083301. [Google Scholar] [CrossRef]
  205. Kim, G.; Fuentes-Hernandez, C.; Jia, X.; Kippelen, B. Organic thin-film transistors with a bottom bilayer gate dielectric having a low operating voltage and high operational stability. ACS Appl. Electron. Mater. 2020, 2, 2813–2818. [Google Scholar] [CrossRef]
  206. Wang, C.-Y.; Fuentes-Hernandez, C.; Chou, W.-F.; Kippelen, B. Top-gate organic field-effect transistors fabricated on paper with high operational stability. Org. Electron. 2017, 41, 340–344. [Google Scholar] [CrossRef]
  207. Wang, C.-Y.; Fuentes-Hernandez, C.; Liu, J.-C.; Dindar, A.; Choi, S.; Youngblood, J.P.; Moon, R.J.; Kippelen, B. Stable low-voltage operation top-gate organic field-effect transistors on cellulose nanocrystal substrates. ACS Appl. Mater. Interfaces 2015, 7, 4804–4808. [Google Scholar] [CrossRef] [PubMed]
  208. Hwang, D.K.; Fuentes-Hernandez, C.; Kim, J.; Potscavage, W.J.; Kim, S.J.; Kippelen, B. Top-gate organic field-effect transistors with high environmental and operational stability. Adv. Mater. 2011, 10, 1293–1298. [Google Scholar] [CrossRef]
  209. Song, W.G.; Kwon, H.J.; Park, J.; Yeo, J.; Kim, M.; Park, S.; Yun, S.; Kyung, K.U.; Grigoropoulos, C.P.; Kim, S. High-performance flexible multilayer MoS2 transistors on solution-based polyimide substrates. Adv. Funct. Mater. 2016, 26, 2426–2434. [Google Scholar] [CrossRef]
  210. Kim, J.H.; Hwang, B.-U.; Kim, D.-I.; Kim, J.S.; Seol, Y.G.; Kim, T.W.; Lee, N.-E. Nanocomposites of polyimide and mixed oxide nanoparticles for high performance nanohybrid gate dielectrics in flexible thin film transistors. Electron. Mater. Lett. 2017, 13, 214–221. [Google Scholar] [CrossRef]
  211. Kim, J.; Kim, J.; Seol, Y.; Kim, T.; Lee, N.-E. High-k Nanocomposite Gate Dielectrics Highly Loaded with Inorganic Nanoparticles by Self-Assembly for Flexible Thin Film Transistors. J. Nanosci. Nanotechnol. 2016, 16, 11335–11342. [Google Scholar] [CrossRef]
  212. Canimkurbey, B.; Çakirlar, Ç.; Piravadili Mucur, S.; Yasin, M.; Berber, S. Influence of Al2O3 nanoparticles incorporation on the dielectric properties of solution processed PVA films for organic field effect transistor applications. J. Mater. Sci. Mater. Electron. 2019, 30, 18384–18390. [Google Scholar] [CrossRef]
  213. Chen, X.; Zhang, H.; Zhang, Y.; Guan, X.; Zhang, Z.; Chen, D. Low-power flexible organic field-effect transistors with solution-processable polymer-ceramic nanoparticle composite dielectrics. Nanomaterials 2020, 10, 518. [Google Scholar] [CrossRef]
  214. Faraji, S.; Danesh, E.; Tate, D.J.; Turner, M.L.; Majewski, L.A. Cyanoethyl cellulose-based nanocomposite dielectric for low-voltage, solution-processed organic field-effect transistors (OFETs). J. Phys. D Appl. Phys. 2016, 49, 185102. [Google Scholar] [CrossRef]
  215. Lim, S.; Lee, K.H.; Kim, H.; Kim, S.H. Optimization of nanocomposite gate insulators for organic thin film transistors. Org. Electron. 2015, 17, 144–150. [Google Scholar] [CrossRef]
  216. Yu, Y.-Y.; Jiang, A.-H.; Lee, W.-Y. Organic/inorganic nano-hybrids with high dielectric constant for organic thin film transistor applications. Nanoscale Res. Lett. 2016, 11, 488. [Google Scholar] [CrossRef]
  217. Alvarado-Beltrán, C.G.; Almaral-Sánchez, J.L.; Mejia, I.; Quevedo-López, M.A.; Ramirez-Bon, R. Sol–gel PMMA–ZrO2 hybrid layers as gate dielectric for low-temperature ZnO-based thin-film transistors. ACS Omega 2017, 2, 6968–6974. [Google Scholar] [CrossRef]
  218. Kim, G.I.; Jung, J.; Min, W.K.; Kim, M.S.; Jung, S.; Choi, D.H.; Chung, J.; Kim, H.J. Mechanically durable organic/high-k inorganic hybrid gate dielectrics enabled by plasma-polymerization of PTFE for flexible electronics. ACS Appl. Mater. Interfaces 2022, 14, 28085–28096. [Google Scholar] [CrossRef]
  219. Kim, K.; Song, H.W.; Shin, K.; Kim, S.H.; Park, C.E. Photo-cross-linkable organic–inorganic hybrid gate dielectric for high performance organic thin film transistors. J. Phys. Chem. C 2016, 120, 5790–5796. [Google Scholar] [CrossRef]
  220. Kim, M.J.; Lee, C.; Jeong, J.; Kim, S.; Lee, T.I.; Shin, E.J.; Hwang, W.S.; Im, S.G.; Cho, B.J. Hf-and Ti-Based Organic/Inorganic Hybrid Dielectrics Synthesized via Chemical Vapor Phase for Advanced Gate Stack in Flexible Electronic Devices. Adv. Electron. Mater. 2021, 7, 2001197. [Google Scholar] [CrossRef]
  221. Kim, M.J.; Pak, K.; Choi, J.; Lee, T.I.; Hwang, W.S.; Im, S.G.; Cho, B.J. Ultrathin ZrO x-organic hybrid dielectric (EOT 3.2 nm) via initiated chemical vapor deposition for high-performance flexible electronics. ACS Appl. Mater. Interfaces 2019, 11, 44513–44520. [Google Scholar] [CrossRef]
  222. Jang, J.; Cho, W.; Kim, B.; Shin, J.H.; Kim, S.; Go, M.; Choi, S.S.; Kim, S.-W.; Kim, J.K.; Jeong, U. Design to secure temporal invariance of triboelectric sensing signals. Nano Energy 2023, 117, 108926. [Google Scholar] [CrossRef]
  223. Cha, S.; Cho, Y.; Kim, J.G.; Choi, H.; Ahn, D.; Sun, J.; Kang, D.s.; Pak, C.; Park, J.J. Controllable Triboelectric Series Using Gradient Positive and Negative Charge-Confinement Layer with Different Particle Sizes of Mesoporous Carbon Materials. Small Methods 2022, 6, 2101545. [Google Scholar] [CrossRef] [PubMed]
  224. Song, G.; Kim, Y.; Yu, S.; Kim, M.-O.; Park, S.-H.; Cho, S.M.; Velusamy, D.B.; Cho, S.H.; Kim, K.L.; Kim, J. Molecularly engineered surface triboelectric nanogenerator by self-assembled monolayers (METS). Chem. Mater. 2015, 27, 4749–4755. [Google Scholar] [CrossRef]
  225. Wang, C.-C.; Chang, C.-Y. Enhanced output performance and stability of triboelectric nanogenerators by employing silane-based self-assembled monolayers. J. Mater. Chem. C 2020, 8, 4542–4548. [Google Scholar] [CrossRef]
  226. He, W.; Qian, Y.; Lee, B.S.; Zhang, F.; Rasheed, A.; Jung, J.-E.; Kang, D.J. Ultrahigh output piezoelectric and triboelectric hybrid nanogenerators based on ZnO nanoflakes/polydimethylsiloxane composite films. ACS Appl. Mater. Interfaces 2018, 10, 44415–44420. [Google Scholar] [CrossRef]
  227. Aazem, I.; Mathew, D.T.; Radhakrishnan, S.; Vijoy, K.; John, H.; Mulvihill, D.M.; Pillai, S.C. Electrode materials for stretchable triboelectric nanogenerator in wearable electronics. RSC Adv. 2022, 12, 10545–10572. [Google Scholar] [CrossRef]
  228. Kuang, X.; Gao, Q.; Zhu, H. Effect of calcination temperature of TiO2 on the crystallinity and the permittivity of PVDF-TrFE/TiO2 composites. J. Appl. Polym. Sci. 2013, 129, 296–300. [Google Scholar] [CrossRef]
  229. Choi, Y.S.; Kim, S.K.; Smith, M.; Williams, F.; Vickers, M.E.; Elliott, J.A.; Kar-Narayan, S. Unprecedented dipole alignment in α-phase nylon-11 nanowires for high-performance energy-harvesting applications. Sci. Adv. 2020, 6, eaay5065. [Google Scholar] [CrossRef]
  230. Guo, H.; Ma, C.; Liu, X.; Tan, X. Electrical poling below coercive field for large piezoelectricity. Appl. Phys. Lett. 2013, 102, 092902. [Google Scholar] [CrossRef]
  231. Li, L.; Xie, L.; Pan, X. Real-time studies of ferroelectric domain switching: A review. Rep. Prog. Phys. 2019, 82, 126502. [Google Scholar] [CrossRef] [PubMed]
  232. Bouza, M.; Li, Y.; Wu, C.; Guo, H.; Wang, Z.L.; Fernández, F.M. Large-Area Triboelectric Nanogenerator Mass Spectrometry: Expanded Coverage, Double-Bond Pinpointing, and Supercharging. J. Am. Soc. Mass Spectrom. 2020, 31, 727–734. [Google Scholar] [CrossRef] [PubMed]
  233. Wang, H.; Xu, L.; Bai, Y.; Wang, Z.L. Pumping up the charge density of a triboelectric nanogenerator by charge-shuttling. Nat. Commun. 2020, 11, 4203. [Google Scholar] [CrossRef] [PubMed]
  234. Cheng, L.; Xu, Q.; Zheng, Y.; Jia, X.; Qin, Y. A self-improving triboelectric nanogenerator with improved charge density and increased charge accumulation speed. Nat. Commun. 2018, 9, 3773. [Google Scholar] [CrossRef] [PubMed]
  235. Sun, J.; Lee, J.; Han, S.; Li, Y.; Park, J.-J.; Bae, J. Output Performance Enhanced Triboelectric Nanogenerators Induced by Magnetic Ink Trapping Property Act as Wearable Sensors. ACS Omega 2024, 9, 3565–3573. [Google Scholar] [CrossRef]
  236. Menge, H.G.; Huynh, N.D.; Hwang, H.J.; Han, S.; Choi, D.; Park, Y.T. Designable skin-like triboelectric nanogenerators using layer-by-layer self-assembled polymeric nanocomposites. ACS Energy Lett. 2021, 6, 2451–2459. [Google Scholar] [CrossRef]
  237. Zhou, J.; Gao, M.; Choi, J. Enhancement of output power and durability of DLC-based sliding TENGs modified with self-assembled monolayers. ACS Appl. Electron. Mater. 2023, 5, 2853–2861. [Google Scholar] [CrossRef]
  238. Hajra, S.; Padhan, A.M.; Sahu, M.; Alagarsamy, P.; Lee, K.; Kim, H.J. Lead-free flexible Bismuth Titanate-PDMS composites: A multifunctional colossal dielectric material for hybrid piezo-triboelectric nanogenerator to sustainably power portable electronics. Nano Energy 2021, 89, 106316. [Google Scholar] [CrossRef]
  239. Han, S.A.; Seung, W.; Kim, J.H.; Kim, S.-W. Ultrathin noncontact-mode triboelectric nanogenerator triggered by giant dielectric material adaption. ACS Energy Lett. 2021, 6, 1189–1197. [Google Scholar] [CrossRef]
  240. Liu, J.; Yu, D.; Zheng, Z.; Huangfu, G.; Guo, Y. Lead-free BiFeO3 film on glass fiber fabric: Wearable hybrid piezoelectric-triboelectric nanogenerator. Ceram. Int. 2021, 47, 3573–3579. [Google Scholar] [CrossRef]
  241. Zhang, P.; Zhang, W.; Deng, L.; Zhang, H. A triboelectric nanogenerator based on temperature-stable high dielectric BaTiO3-based ceramic powder for energy harvesting. Nano Energy 2021, 87, 106176. [Google Scholar] [CrossRef]
  242. Küchler, A. High Voltage Engineering: Fundamentals-Technology-Applications; Springer: Berlin/Heidelberg, Germany, 2017. [Google Scholar]
  243. Wang, S.; Xie, Y.; Niu, S.; Lin, L.; Liu, C.; Zhou, Y.S.; Wang, Z.L. Maximum Surface Charge Density for Triboelectric Nanogenerators Achieved by Ionized-Air Injection: Methodology and Theoretical Understanding. Adv. Mater. 2014, 26, 6720–6728. [Google Scholar] [CrossRef]
  244. Shin, S.-H.; Bae, Y.E.; Moon, H.K.; Kim, J.; Choi, S.-H.; Kim, Y.; Yoon, H.J.; Lee, M.H.; Nah, J. Formation of Triboelectric Series via Atomic-Level Surface Functionalization for Triboelectric Energy Harvesting. ACS Nano 2017, 11, 6131–6138. [Google Scholar] [CrossRef]
  245. Patnam, H.; Dudem, B.; Graham, S.A.; Yu, J.S. High-performance and robust triboelectric nanogenerators based on optimal microstructured poly (vinyl alcohol) and poly (vinylidene fluoride) polymers for self-powered electronic applications. Energy 2021, 223, 120031. [Google Scholar] [CrossRef]
  246. Chen, F.; Wu, Y.; Ding, Z.; Xia, X.; Li, S.; Zheng, H.; Diao, C.; Yue, G.; Zi, Y. A novel triboelectric nanogenerator based on electrospun polyvinylidene fluoride nanofibers for effective acoustic energy harvesting and self-powered multifunctional sensing. Nano Energy 2019, 56, 241–251. [Google Scholar] [CrossRef]
  247. Kim, J.; Lee, J.H.; Ryu, H.; Lee, J.H.; Khan, U.; Kim, H.; Kwak, S.S.; Kim, S.W. High-performance piezoelectric, pyroelectric, and triboelectric nanogenerators based on P (VDF-TrFE) with controlled crystallinity and dipole alignment. Adv. Funct. Mater. 2017, 27, 1700702. [Google Scholar] [CrossRef]
  248. Granzow, T.; Kounga, A.; Aulbach, E.; Rödel, J. Electromechanical poling of piezoelectrics. Appl. Phys. Lett. 2006, 88, 252907. [Google Scholar] [CrossRef]
  249. Kamel, T.M.; de With, G. Poling of hard ferroelectric PZT ceramics. J. Eur. Ceram. Soc. 2008, 28, 1827–1838. [Google Scholar] [CrossRef]
  250. Park, Y.; Shin, Y.-E.; Park, J.; Lee, Y.; Kim, M.P.; Kim, Y.-R.; Na, S.; Ghosh, S.K.; Ko, H. Ferroelectric multilayer nanocomposites with polarization and stress concentration structures for enhanced triboelectric performances. ACS Nano 2020, 14, 7101–7110. [Google Scholar] [CrossRef]
  251. Pei, J.-Y.; Zhong, S.-L.; Zhao, Y.; Yin, L.-J.; Feng, Q.-K.; Huang, L.; Liu, D.-F.; Zhang, Y.-X.; Dang, Z.-M. All-organic dielectric polymer films exhibiting superior electric breakdown strength and discharged energy density by adjusting the electrode–dielectric interface with an organic nano-interlayer. Energy Environ. Sci. 2021, 14, 5513–5522. [Google Scholar] [CrossRef]
  252. Lee, K.Y.; Kim, S.K.; Lee, J.H.; Seol, D.; Gupta, M.K.; Kim, Y.; Kim, S.W. Controllable charge transfer by ferroelectric polarization mediated triboelectricity. Adv. Funct. Mater. 2016, 26, 3067–3073. [Google Scholar] [CrossRef]
  253. Bai, P.; Zhu, G.; Zhou, Y.S.; Wang, S.; Ma, J.; Zhang, G.; Wang, Z.L. Dipole-moment-induced effect on contact electrification for triboelectric nanogenerators. Nano Res. 2014, 7, 990–997. [Google Scholar] [CrossRef]
  254. Shin, K.S.; Kim, T.Y.; Yoon, G.C.; Gupta, M.K.; Kim, S.K.; Seung, W.; Kim, H.; Kim, S.; Kim, S.; Kim, S.W. Ferroelectric Coupling Effect on the Energy-Band Structure of Hybrid Heterojunctions with Self-Organized P (VDF-TrFE) Nanomatrices. Adv. Mater. 2014, 26, 5619–5625. [Google Scholar] [CrossRef] [PubMed]
  255. Cui, N.; Gu, L.; Lei, Y.; Liu, J.; Qin, Y.; Ma, X.; Hao, Y.; Wang, Z.L. Dynamic behavior of the triboelectric charges and structural optimization of the friction layer for a triboelectric nanogenerator. ACS Nano 2016, 10, 6131–6138. [Google Scholar] [CrossRef]
  256. Kornfeld, M. Frictional electrification. J. Phys. D Appl. Phys. 1976, 9, 1183. [Google Scholar] [CrossRef]
  257. Park, H.-W.; Huynh, N.D.; Kim, W.; Lee, C.; Nam, Y.; Lee, S.; Chung, K.-B.; Choi, D. Electron blocking layer-based interfacial design for highly-enhanced triboelectric nanogenerators. Nano Energy 2018, 50, 9–15. [Google Scholar] [CrossRef]
  258. Lee, J.-H.; Hinchet, R.; Kim, T.Y.; Ryu, H.; Seung, W.; Yoon, H.-J.; Kim, S.-W. Control of Skin Potential by Triboelectrification with Ferroelectric Polymers. Adv. Mater. 2015, 27, 5553–5558. [Google Scholar] [CrossRef]
  259. Khan, U.; Kim, T.H.; Ryu, H.; Seung, W.; Kim, S.W. Graphene tribotronics for electronic skin and touch screen applications. Adv. Mater. 2017, 29, 1603544. [Google Scholar] [CrossRef]
  260. Kim, Y.W.; Lee, H.B.; Yoon, J.; Park, S.-H. 3D customized triboelectric nanogenerator with high performance achieved via charge-trapping effect and strain-mismatching friction. Nano Energy 2022, 95, 107051. [Google Scholar] [CrossRef]
  261. Wu, C.; Kim, T.W.; Choi, H.Y. Reduced graphene-oxide acting as electron-trapping sites in the friction layer for giant triboelectric enhancement. Nano Energy 2017, 32, 542–550. [Google Scholar] [CrossRef]
  262. Qiao, Y.; Yin, X.; Zhu, T.; Li, H.; Tang, C. Dielectric polymers with novel chemistry, compositions and architectures. Prog. Polym. Sci. 2018, 80, 153–162. [Google Scholar] [CrossRef]
  263. Wei, J.; Zhang, Z.; Tseng, J.-K.; Treufeld, I.; Liu, X.; Litt, M.H.; Zhu, L. Achieving high dielectric constant and low loss property in a dipolar glass polymer containing strongly dipolar and small-sized sulfone groups. ACS Appl. Mater. Interfaces 2015, 7, 5248–5257. [Google Scholar] [CrossRef] [PubMed]
  264. Koutsiaki, C.; Kaimakamis, T.; Zachariadis, A.; Papamichail, A.; Kamaraki, C.; Fachouri, S.; Gravalidis, C.; Laskarakis, A.; Logothetidis, S. Efficient combination of Roll-to-Roll compatible techniques towards the large area deposition of a polymer dielectric film and the solution-processing of an organic semiconductor for the field-effect transistors fabrication on plastic substrate. Org. Electron. 2019, 73, 231–239. [Google Scholar] [CrossRef]
  265. Yu, S.J.; Pak, K.; Kwak, M.J.; Joo, M.; Kim, B.J.; Oh, M.S.; Baek, J.; Park, H.; Choi, G.; Kim, D.H. Initiated chemical vapor deposition: A versatile tool for various device applications. Adv. Eng. Mater. 2018, 20, 1700622. [Google Scholar] [CrossRef]
  266. Chun, J.; Kim, J.W.; Jung, W.-s.; Kang, C.-Y.; Kim, S.-W.; Wang, Z.L.; Baik, J.M. Mesoporous pores impregnated with Au nanoparticles as effective dielectrics for enhancing triboelectric nanogenerator performance in harsh environments. Energy Environ. Sci. 2015, 8, 3006–3012. [Google Scholar] [CrossRef]
  267. Kwak, S.S.; Kim, S.M.; Ryu, H.; Kim, J.; Khan, U.; Yoon, H.-J.; Jeong, Y.H.; Kim, S.-W. Butylated melamine formaldehyde as a durable and highly positive friction layer for stable, high output triboelectric nanogenerators. Energy Environ. Sci. 2019, 12, 3156–3163. [Google Scholar] [CrossRef]
  268. Lee, Y.; Cha, S.H.; Kim, Y.-W.; Choi, D.; Sun, J.-Y. Transparent and attachable ionic communicators based on self-cleanable triboelectric nanogenerators. Nat. Commun. 2018, 9, 1804. [Google Scholar] [CrossRef]
Table 1. Comparison of polymer dielectric-based memristor devices.
Table 1. Comparison of polymer dielectric-based memristor devices.
Active Materials
(Method)
Resistance
on/off Ratio
Switching
Cycles
Retention Time
(s)
Working Temperature
(°C)
Air
Stability
(Days)
Biodegradable[Ref.]
PVA
(Spin-coating)
N/A5 × 103104From 20 to 80N/A[71]
PPTNMI+Br− 1
(Spin-coating)
N/A102104Room
temperature
N/A[10]
MDMO-PPV
(Spin-coating)
N/A104104From −196 to 300N/A[79]
PEI
(Spin-coating)
~105103105From 25 to 150N/A[76]
PTPA
(Spin-coating)
~108<108 × 103From −243 to 117N/A[77]
PEI-AgClO4
(Spin-coating)
~1035 × 1027 × 102Room
temperature
30[83]
BPQDs@PDA-PVP 3
(Spin-coating)
N/A2 × 1022 × 104Room
temperature
90[84]
pV3D3
(iCVD)
~109105105From 27 to 85N/A[51]
pV3D3
(iCVD)
~108105105From 27 to 8727[85]
pEDGMA 2
(iCVD)
2.5 × 1025 × 102107Room
temperature
N/A[86]
1 reactant of poly[thiophene-alt-4,4-bis(6-bromohexyl)-4H-cyclopenta(1,2-b:5,4-b′)dithiophene] (PTT-Br) with N-methylimidazole (NMI). 2 poly(ethylene glycol dimethacrylate) (pEDGMA). 3 blend of black phosphorus quantum dots (BPQDs) with polyvinylpyrrolidone-grafted polydopamine (PDA-PVP).
Table 2. Comparison of polymer dielectric-based charge trap devices.
Table 2. Comparison of polymer dielectric-based charge trap devices.
Charge Trap StructureDevice TypeCurrent on/off RatioMemory Window (Operating Voltage)Switching CyclesRetention Time (s)[Ref.]
pV3D3/AuNP/pC1D1Floating gate~106~5.7 V (±13 V)1.5 × 103104[50]
pV3D3/Al/pEDGMAFloating gate1065.5 V (±10 V)1033.2 × 108[55]
PMMA@F8BT/P(VDF-TrFE-CFE) 2Floating gate6.5 × 1039.3 V (±40 V)102>104[93]
PFOPolymer electrets~10776 V (Light/−100 V)N/A~4 × 103[88]
PαMS 1Polymer electrets10523 V (±20 V)5 × 103104[94]
pV3D3Polymer electrets>1055.3 V (±14 V)50>105[95]
PF-b-PDLPolymer electrets105102 V (±140 V)N/A104[21]
PF-b-Piso 3Polymer electrets10633 V (Light/−40 V)10104[91]
PVN
[email protected]% PCBM
PVN@30% PCBM
Polymer electrets>104
>104
>104
~12 V (±20 V)
~14 V (±20 V)
~15 V (±20 V)
N/A
N/A
N/A
103
103
103
[22]
N2200@ PVN 4Polymer electrets10530 V (+30 V/−18 V)103104[90]
1 poly(α-methylstyrene) (PαMS). 2 poly(9,9-dioctylflfluorene-co-benzothiadiazole) (F8BT) and poly(vinylidene flfluoride-triflfluoroethy-lene-chloroflfluoroethylene) (P(VDF-TrFE-CFE)). 3 poly [2,7-(9,9-dihexylfluorene)]-b- poly(pendent isoindigo) (PF-b-Piso). 4 blend of poly [96] (N2200) with poly(2-vinyl naphthalene) (PVN).
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.; Choi, J.; Han, Y.; Yoo, H.; Yoon, H.-J. Polymer Dielectric-Based Emerging Devices: Advancements in Memory, Field-Effect Transistor, and Nanogenerator Technologies. Micromachines 2024, 15, 1115. https://doi.org/10.3390/mi15091115

AMA Style

Choi W, Choi J, Han Y, Yoo H, Yoon H-J. Polymer Dielectric-Based Emerging Devices: Advancements in Memory, Field-Effect Transistor, and Nanogenerator Technologies. Micromachines. 2024; 15(9):1115. https://doi.org/10.3390/mi15091115

Chicago/Turabian Style

Choi, Wangmyung, Junhwan Choi, Yongbin Han, Hocheon Yoo, and Hong-Joon Yoon. 2024. "Polymer Dielectric-Based Emerging Devices: Advancements in Memory, Field-Effect Transistor, and Nanogenerator Technologies" Micromachines 15, no. 9: 1115. https://doi.org/10.3390/mi15091115

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