Next Article in Journal
Multiple Moving Vehicles Tracking Algorithm with Attention Mechanism and Motion Model
Next Article in Special Issue
Evaluation of the Electronic Properties of Atomic Layer Deposition-Grown Ge-Doped Zinc Oxide Thin Films at Elevated Temperatures
Previous Article in Journal
Emotional and Mental Nuances and Technological Approaches: Optimising Fact-Check Dissemination through Cognitive Reinforcement Technique
Previous Article in Special Issue
Semi-Analytical Approach versus Finite Element Method for Analysis of Propagation Properties in Rectangular Waveguides: Silica-Titania Technological Platform
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Active-Matrix Array Based on Thin-Film Transistors Using Emerging Materials for Application: From Lab to Industry

by
Seongjae Kim
and
Hocheon Yoo
*
Department of Electronic Engineering, Gachon University, Seongnam 13120, Republic of Korea
*
Author to whom correspondence should be addressed.
Electronics 2024, 13(1), 241; https://doi.org/10.3390/electronics13010241
Submission received: 4 December 2023 / Revised: 24 December 2023 / Accepted: 2 January 2024 / Published: 4 January 2024

Abstract

:
The active-matrix technology incorporates a transistor to exert precise control over each pixel within a pixel array, eliminating the issue of crosstalk between neighboring pixels that is prevalent in the passive-matrix approach. Consequently, the active-matrix method facilitates the realization of high-resolution arrays, and this inherent advantage has propelled its widespread adoption, not only in display applications but also in diverse sensor arrays from lab to industry. In this comprehensive review, we delve into instances of active-matrix arrays utilizing thin-film transistors (TFTs) that leverage emerging materials such as organic semiconductors, metal oxide semiconductors, two-dimensional materials, and carbon nanotubes (CNTs). Our examination encompasses a broad classification of active-matrix research into two main categories: (i) displays and (ii) sensors. We not only assess the performance of TFTs based on emerging materials within the active-matrix framework, but also explore the evolving trends and directions in active-matrix-based displays and sensors.

1. Introduction

Thin-film transistors (TFTs), based on field-effect transistors (FETs), are manufactured through a bottom–up process that involves the systematic stacking of layers. This bottom–up approach bestows significant process advantages, offering a considerable level of flexibility in structural design and material selection, thereby providing TFTs with a highly versatile platform. This design flexibility allows TFTs to extend beyond display backplanes [1,2,3] into various device areas, including digital and analog circuits [4,5,6,7,8,9] and image sensors [10,11,12]. In particular, many studies have been carried out recently to apply emerging materials such as organics [13,14,15,16,17], metal-oxide semiconductors [18,19,20,21], two-dimensional (2D) materials [22,23,24], and carbon nanotubes (CNTs) [25,26,27,28], rather than Si, as TFT channels.
While demonstrations have often been limited to the unit device level, proving practical device feasibility requires expanding to array scale. These array types are broadly categorized as passive-matrix [29,30] and active-matrix [31,32]. The passive-matrix takes the form of a crossbar array and has the advantage of a simple structure and low manufacturing cost. However, resolution improvement is constrained by crosstalk issues between adjacent pixels in the passive-matrix [32]. On the other hand, active-matrix, with a TFT in each pixel enabling independent pixel operation, avoids crosstalk and can achieve higher resolution [33]. This has led to the adoption of active-matrix not only in displays but also in various sensors, including photodetectors [34,35,36] and pressure sensors [37,38,39], etc.
In this review, we highlight research outcomes on active-matrix-based displays and various sensors, focusing particularly on studies implementing active-matrix with TFTs based on emerging materials. Furthermore, we revisit the specific purposes served by each sensor and explore the technologies they use. This review explains the interest of the use of the active-matrix approach across diverse applications and investigates its potential for expanding practicality from lab to industry.

2. Thin-Film Transistors and Active-Matrix Array

TFTs are mainly fabricated in a bottom–up process, unlike Si-based metal-oxide-semiconductor field-effect transistors (MOSFETs), which are manufactured in a top-down process [40]. A bottom–up process-based TFT is not only free in terms of designing device structure but also has excellent substrate compatibility [41,42,43]. In addition, a TFT can be implemented based on various processes such as spin coating [44,45], inkjet printing [46,47], photolithography [48,49], and thermal evaporation [50,51]. Through considerable effort, large-area processes have also become possible. Securing large-area process technology has enabled development beyond the characteristic evaluation at the unit device level to TFT-based system levels, such as digital and analog circuits, including active-matrix arrays. In particular, active matrices can be applied to various applications and offer many advantages. The active-matrix array is a method originally used for addressing the pixels in the display. Unlike the passive-matrix, which is a crossbar array that crosses horizontally and vertically, the active-matrix array contains one transistor per unit pixel. In the active-matrix, pixels are selected by controlling the on/off of the TFT. The very low off current of the TFT prevents the crosstalk issues that are problematic in the passive-matrix. This leads the active-matrix array to be used in various ways, such as sensor arrays and displays.
As mentioned earlier, the active-matrix array was initially adopted as an addressing method for displays and was later used to drive sensor arrays beyond displays. The driving methods of a sensor array and a display are not significantly different. A display converts electrical signals within the system into light. On the other hand, in a sensor array, the opposite process is achieved by converting information such as light and temperature into electrical signals through sensor devices and processing them in the internal system. In other words, both the display and the sensor array are the same up to the point of selecting a specific column through a switching transistor, and the only difference is the direction of signal transmission through the data line.
The implementation of an active-matrix array allows for the expression and sensing of more information, such as displaying a specific image or sensing the distribution of temperature or pressure. These unique features of the active-matrix array contribute to the ongoing development of advanced display and sensor systems.

3. Active-Matrix-Based Displays

3.1. Liquid Crystal Display (LCD)

A liquid crystal display (LCD) is one of the types of displays and is currently used in most televisions and monitors. The working principle is to control the amount of light transmitted by controlling the arrangement of liquid crystals located between two polarizers through voltage. To express color, color filters corresponding to red, green, and blue are used. Additionally, there are LCDs that are driven by using a reflector instead of a backlight to reflect external light. Cong et al. demonstrated an active-matrix polymer-dispersed liquid crystal (PDLC) seven-segment display using CNT TFT [52]. They utilized PDLC, formed by embedding micro-sized liquid crystals into a polymer matrix, and controlled the transparent or opaque state based on the refractive index difference between the polymer matrix and the liquid crystal, depending on the applied voltage (Figure 1a,b). The proposed PDLC display has a structure in which two ITO electrodes are sandwiched above and below the liquid crystal (Figure 1c). Figure 1d illustrates a CNT TFT-based backplane for driving a seven-segment display. The fabricated PDLC display demonstrated the ability to clearly distinguish light transmittance in the visible light region based on the voltage applied to both ends of the PDLC. Building on this capability, they successfully displayed numerical information by selectively controlling the liquid crystal of each pixel through CNT TFT.

3.2. Light-Emitting Diode (LED)

Unlike LCD, light-emitting diode does not require a backlight and the pixels emit light on their own; so, the pixels can be turned off completely, making real darkness possible. In addition, since there is no need for layers such as backlights and polarizers, as required in LCD, it is possible to implement thinner displays. These light-emitting diodes (LEDs) are divided into organic LED (OLED) [53,54,55,56,57], micro- LED (μ-LED) [58,59], and quantum dots LED (QLED) [60,61], depending on the materials used, and each is being actively researched and developed by several researchers. In particular, many attempts to apply emerging materials such as organic, oxide semiconductor, and 2D materials as backplanes have been reported.
Peng et al. developed a driving circuit for an 8 × 8 active-matrix LED based on organic TFTs (OTFTs) on paper [62]. They utilized paper with high roughness as a substrate, gradually alleviating the roughness by forming screen-printed silver and parylene-C as the gate electrode and gate dielectric. The reduced surface roughness led to the successful implementation of an 8 × 8 organic transistor array with a 100% yield (Figure 2a,b). Finally, they integrated LEDs into the array and demonstrated the display of character information by controlling the active-matrix LEDs with the proposed organic transistor array (Figure 2c). This highlights the feasibility of implementing functional circuits, such as active-matrix arrays, using organic semiconductors even on rough substrates like paper.
Um et al. developed an active-matrix LED display by integrating GaN μ-LEDs with an amorphous indium-gallium-zinc-oxide (a-IGZO) TFT backplane [63]. First, they fabricated the a-IGZO TFT backplane and GaN μ-LED pixel array on glass and sapphire substrates, respectively. Subsequently, they completed the implementation of a 128 × 384 μ-LED array display by integrating the GaN μ-LED on the sapphire substrate into the a-IGZO TFT backplane through flip chip bonding (Figure 2d). The flip chip bonding process was conducted below 100 °C, ensuring no deterioration in the characteristics of the TFT backplane, and consequently, text and portrait images were successfully displayed on the display.
Choi et al. attempted to configure the backplane of the OLED display with MoS2 TFT [64]. They optimized the characteristics of MoS2 TFT before applying it as a driving TFT. The top and bottom of the MoS2 channel were encapsulated with Al2O3, resulting in high on-current and low hysteresis characteristics. The introduction of the Al2O3 layer reduced the charge trap density at the interface, contributing to improved electrical properties. Using MoS2 with enhanced characteristics, they implemented an OLED display on a flexible polymer substrate. The positive threshold voltage of the MoS2 TFT reduced power consumption as it did not require a separate voltage to turn off the pixel. They demonstrated real-time control of text images by attaching a flexible OLED display with the manufactured MoS2 backplane to a wrist (Figure 2e).
Baek et al. implemented a CNT TFT-based active-matrix QLED display [65]. In particular, to investigate the optimum design for a high-performance QLED array, they compared the characteristics of a total of four types of QLED arrays according to the type of driving TFT (p- or n-type) and the connection configuration between QLED and the driving transistor (conventional QLED or inverted QLED) (Figure 2f). They passivated Si3N4 on p-type CNT TFT to prepare n-type CNT TFT. The conversion of CNT TFT to n-type operation through Si3N4 passivation is because the positive fixed charge in Si3N4 causes n-type doping and protects it from atmospheric oxygen and water, which cause p-type doping effects [65,66]. Subsequently, the distinction between conventional QLED and inverted QLED was defined based on whether the upper Al and lower ITO electrodes served as cathode/anode or anode/cathode, respectively. At this point, depending on whether ITO, the bottom electrode of QLED, is in contact with the drain or source of the driving transistor, the cases were classified into four structures: (i) drain contact with the bottom anode (DCBA), (ii) source contact with the bottom cathode (SCBC), (iii) drain contact with the bottom cathode (DCBC), and (iv) source contact with the bottom anode (SCBA). As a result of comparing the electrical and optical characteristics of the four structures, when the QLED was connected to the source of the driving TFT (SCBC, SCBA), it showed poorer performance than when connected to the drain. This is because the voltage between the gate and source, which determines the drain current, is reduced due to the voltage drop occurring at the QLED. Therefore, when applying a p-type or n-type TFT as a driving TFT, it is advantageous to adopt the DCBA and DCBC structures. Based on this, they ultimately produced a 5 × 5 QLED array with p-type and n-type CNT TFTs as driving TFTs and demonstrated real-time display of character images.

3.3. Light-Emitting Electrochemical Cell (LEC)

Unlike OLED, which requires additional layers such as an electron injection layer and an electron transport layer in addition to the light-emitting layer, a light-emitting electrochemical cell (LEC) has a simple structure consisting of a light-emitting layer and two metal electrodes [67,68,69,70]. This simple structure is a great advantage, especially when implementing stretchable light-emitting devices. In order to provide elasticity to existing OLEDs, elasticity must be imparted to all layers, including the electron and hole injection layers as well as the light-emitting layer. However, in the case of LEC, only the light-emitting layer needs to be elastic; therefore, the process’ difficulty is low.
Liu et al. utilized the advantages of LEC to implement an organic-based stretchable active-matrix organic LEC (OLEC) array [71]. They first fabricated a stretchable active-matrix TFT array to drive a stretchable OLEC array. They mixed a crosslinker containing flexible polydimethysiloxane into conjugated polymers to prevent the semiconductor layer from delamination from the perfluorinated elastomer used as the gate dielectric when the TFT is stretched (Figure 3a). At this time, the introduced crosslinker strengthened the bond between the two layers by crosslinking the conjugated polymers and the elastomeric dielectric layer during the annealing process, thereby preventing delamination of the semiconductor layer when stretching. Based on this, they successfully fabricated a stretchable TFT array (Figure 3b,c). Then, OLEC was integrated on the TFT array (Figure 3d), and it was demonstrated that OLEC could be driven through OTFT both without twisting and with twisting (Figure 3e,f).

4. Active-Matrix-Based Sensors

Unlike displays that process internal electrical signals to output visual information, sensors play a crucial role by receiving information about specific external elements and converting it into corresponding electrical signals. Sensors are categorized based on the detection target, ranging from particles like photons [72,73,74,75] and gases [76,77,78] to external physical changes such as pressure [37,79,80], strain [81,82,83], and temperature [83,84]. The development of these various sensors can provide many benefits to humans, such as preventing accidents from occurring caused by harmful elements such as toxic gases and ultraviolet, and enabling an accurate evaluation of the strength of products produced at manufacturing sites. Furthermore, by mimicking human sense organs such as vision and touch, sensors can contribute to the ongoing development of artificial intelligence and robotics, paving the way for innovative technologies in the future.

4.1. Photosensors

Photosensors are devices that convert optical signals into electrical signals and can perform the function of vision among the human senses. In addition, photosensors can be manufactured by targeting visible light [12,34,85], ultraviolet [86,87,88], and infrared [35,89,90] depending on the purpose. In particular, it is interesting that the presence or absence of light irradiation that humans cannot see, such as ultraviolet and infrared, can be detected through the photosensor.
Takahashi et al. achieved an imager capable of detecting visible light by monolithically integrating a CNT-based backplane with organic photodiodes [91]. They utilized PI substrates to take advantage of the material advantages of CNT, regioregular poly(3-hexylthiophene) (P3HT), and [6,6]-phenyl C61-butyric acid methyl ester (PCBM) in both TFTs and photodiodes. This approach resulted in the successful demonstration of a visible light imager with flexibility. Moreover, they integrated a Gd2O2S/Tb scintillator film, capable of converting X-rays into green light, into the visible light imager (Figure 4a,b). The introduction of the Gd2O2S/Tb scintillator film facilitated X-ray detection without causing damage to the organic active layer of the photodiode (Figure 4c).
X-rays possess strong energy and can be harmful to the human body, necessitating their detection for safety. In contrast, near-infrared (NIR), like X-rays, is invisible to humans but has very weak energy and is not harmful to the human body. Detection of NIR is significant as it can be utilized in various applications, including temperature measurement, near-field communication, and medical fields. In medical applications, the deep penetration characteristics of NIR enable measurements such as heart rate and oxygen saturation. Li et al. employed solution-processed In2O3 to construct the backplane of a NIR sensor array [89]. Additionally, they integrated solution-processed In2O3, known for its high electron mobility, into an NIR-responsive hybrid phototransistor along with an organic bulk heterojunction (Figure 4d). PTB7-Th and BTPV-4F were used as donor and acceptor materials, respectively, for the organic bulk heterojunction. Among the two materials, BTPV-4F is responsible for the actual absorption of NIR, while PTB7-Th is intended to improve the separation of electron–hole pairs through a built-in electric field generated by the formation of an organic bulk heterojunction. Finally, they manufactured a 16 × 16 hybrid NIR sensor array based on unit pixels in the form of one transistor–one phototransistor and successfully detected the NIR irradiation area.
The detection of a specific wavelength band, such as X-ray or NIR, is determined by the intended use of the sensor [92]. To imitate human vision, the sensor must be designed to target the visible light region, and each pixel should have the ability to distinguish between red, green, and blue light. Kim et al. introduced CdSe quantum dots as the active layer of a phototransistor on top of an a-IGZO transistor as a method to implement a color-selective imaging function. They initially demonstrated photodetector arrays with phototransistors capable of detecting red, green, and blue light, respectively, arranged in the lateral direction. Subsequently, to enhance the resolution, they attempted to vertically stack the quantum dots responsible for red, green, and blue. By investigating the photoresponse characteristics according to the stacking order of the red, green, and blue quantum dots, it was shown that stacking them from the bottom is the optimal structure. Based on this, they finally implemented color-selective photodetection at the same region by integrating an a-IGZO TFT-based backplane and a phototransistor with vertically stacked quantum dots.
The primary function of a photosensor is to detect light, but when its light-sensing properties are utilized, it can be actively employed in medical applications, such as for measuring a person’s heart rate and oxygen saturation. Ruiz-Preciado et al. fabricated a highly flexible all-organic photosensor array by integrating organic photodiodes and OTFTs through an inkjet-printing process [93]. They demonstrated photoplethysmography (PPG) measurement by directing a red LED onto a finger and observing the pulsating blood flow through the current response in the photosensor array generated by the transmitted light.

4.2. Gas Sensor

Gas sensors are widely applied as devices that detect the presence of target gas in a specific space. In particular, measuring oxygen concentration or detecting harmful gases can inform people in the area of danger early and help them evacuate. Therefore, it is important for gas sensors to detect even minute levels of concentration quickly. Kim et al. demonstrated NO2 gas detection using a two-step grown MoS2 TFT-based gas sensor array (Figure 5a) [94]. Here, the two-step growing method proceeds with a sputtering step followed by a thermal sulfurization step. For both the switching TFT and the sensing TFT, two-step grown MoS2 was applied as a channel, but the switching TFT was passivated with a SiO2 layer to prevent reaction with gas (Figure 5b). In the case of the sensing TFT, the drain current decreased as the concentration of NO2 gas increased (Figure 5c). This means that NO2 gas increases the resistance of the MoS2 channel, and when the concentration of each NO2 gas was 128 ppm, the resistance increased by about 60% compared to the initial resistance (Figure 5d). The reason why MoS2 TFT was able to react to NO2 is because two-step grown MoS2 has a poly-crystalline structure and many grain boundaries. In fact, for this reason, defects are intentionally formed to improve reactivity to gas.
Unlike photosensors and pressure sensors, gas sensors have not yet been implemented at the array level in many cases. This may be because, unlike other sensors that can obtain meaningful information by detecting specific image information or pressure distribution, gas sensors only need to detect the presence and concentration level of a specific gas. Therefore, if expanded in the direction of simultaneously detecting two or more different gases rather than just one specific gas, a gas sensor array could also be attractive.

4.3. Pressure Sensors

Shin et al. implemented a pressure sensor array based on graphene FET with air-dielectric [95]. They used two plastic panels connected by elastic joints as a substrate and created source/drain patterns and gate patterns on each panel. Afterward, by folding the elastic joint to the boundary, the two panels were completely overlapped and the FET structure was completed. At this time, the thickness of the air-dielectric corresponding to the gap between the graphene and the top gate is determined by the thickness of the elastomeric partition spacers. Therefore, the thickness of the air-dielectric and the pressure are inversely proportional, and as the pressure increases, the thickness of the gate dielectric of the graphene FET decreases. This means that an increase in pressure induces the capacitance of the gate dielectric, ultimately leading to an increase in the current of the graphene FET. Additionally, they successfully demonstrated a transparent pressure sensor array by introducing Ag NWs as electrodes (Figure 6a–c). However, the proposed graphene FET-based pressure sensor has limitations due to high off current; so, for low-power operation, the addition of FETs or other alternatives was needed to select each cell.
Jang et al. replaced graphene with MoS2, a well-known 2D semiconductor, as the channel layer to address the high-off-current issue, a drawback in existing graphene FET-based pressure sensors [96]. Although MoS2 has a two-dimensional structure, it differs from graphene in having a band gap suitable for a semiconductor channel. Notably, MoS2 exhibits the ability to control electrical and optical properties depending on the number of layers. In the case of monolayer MoS2, it possesses a direct band gap and outstanding optical properties. By incorporating monolayer MoS2, which exhibits both excellent electrical and optical properties, into the pressure sensor, they successfully developed a pressure sensor array with low power consumption. Moreover, they leveraged the optical properties of monolayer MoS2 to enhance pressure sensitivity further by introducing mechanoluminescent materials that emit visible light when compression or friction force is applied (Figure 6d). Through this approach, when pressure is applied to the MoS2-based pressure sensor, not only does the dielectric thickness decrease, but the additional effect of excess carriers being created in MoS2 by the emission of mechanoluminescent materials is introduced. This results in an increased amount of change in current for the same pressure (Figure 6e).
Among the existing pressure sensor types, in addition to the previously mentioned FET-type, many piezoresistive pressure sensors have been reported. The piezoresistive type offers the advantage of not only having a simple structure but also exhibiting high pressure sensitivity and enabling continuous sensing. Zhao et al. developed a pressure sensor utilizing a 64 × 64 CNT TFT-based backplane and a piezoresistive film (PRF) blended with CNT and thermoplastic polyurethane (TPU) elastomer (Figure 6f,g) [97]. Before the full-scale demonstration, they conducted a comprehensive investigation into the characteristics of the pressure sensor. The sensor demonstrated faster response times, with rising and falling times of 5 ms and 3 ms, respectively, compared to existing PRF-based pressure sensors. Additionally, the proposed pressure sensor exhibited excellent sensitivity and spatial resolution. Notably, it was demonstrated that the sensor could detect even the footprint of an artificial honeybee weighing 6.7 g, as illustrated in Figure 6h.
The high sensitivity of the pressure sensor allows it to detect even minute changes in pressure; so, it can also be applied to biosignal monitoring applications with weak signal strength. Karner-Petritz et al. implemented a pressure sensor array, utilizing a poly(vinylidene fluoride: trifluoroethylene) (P(VDF:TrFE))-based ferroelectric transducer and a DNTT-based OTFT for each pixel [98]. They fabricated an organic-based pressure sensor array on a 1 um thick parylene substrate (Figure 6i). The ultra-flexible properties of the parylene substrate allow the pressure sensor array to be attached to uneven human skin. Exploiting this flexibility, they measured pulse waves using pressure sensors aligned with the direction of blood flow (Figure 6j). Moreover, they demonstrated the extraction of pulse wave velocity and blood pressure by analyzing the time difference at the peak point of the pulse wave from sensors spaced apart from each other.
Efforts persist in enhancing the performance of diverse pressure sensors, with endeavors directed towards incorporating additional features, including transparency. The pursuit of high-performance pressure sensors holds significance as it can pave the way for the creation of bio-health devices characterized by low latency and high precision. Such devices enable the monitoring of vital biological indicators like blood pressure and heart rate, contributing to advancements in healthcare technology.

4.4. Strain Sensors

Sun et al. fabricated an active-matrix strain sensor array by integrating a P(VDF-TrFE)-based piezoelectric nanogenerator (NG) and a graphene FET [99]. The bottom electrode of the piezoelectric NG and the gate electrode of the graphene FET are connected with patterned graphene (Figure 7a). Additionally, the conductivity of the graphene FET is determined by the potential of the graphene gate electrode through the electrolyte. As a result of evaluating the characteristics of the proposed strain sensor, they showed that when strain is applied to the device, the conductivity of the graphene FET is modulated as a piezoelectric potential is generated due to the alignment of the dipole in the PVDF that constitutes the piezoelectric NG. They further expanded the sensor to a 4 × 4 array based on piezopotential-gated graphene FETs and visualized the distribution of strain applied to the substrate.
In the context of strain sensors, being devices designed to measure externally applied stress, they operate in an environment where the components can be susceptible to damage through repetitive measurements. Consequently, there is a need for research aimed at enhancing the reliability of strain sensors to facilitate continuous sensing. Oh et al. introduced a semiconductor thin film with self-healing properties, offering insights into addressing reliability issues in strain sensors [100]. They employed a semiconductor thin film comprising a mixture of DPP-TVT-PDCA and PDMS-PDCA-Fe elastomer, determining through an examination of carrier mobility at various mixing ratios that the optimal ratio was 1:5. The self-healing properties of the proposed semiconductor thin film mixed with DPP-TVT-PDCA and PDMS-PDCA-Fe elastomer result from the presence of abundant dynamic metal–ligand coordination bonds [101]. Finally, they fabricated a 5 × 5 strain sensor array and performed additional passivation to prevent device malfunction due to human sweat, showing that the electrical characteristics remained stable even after 15 h after dropping artificial sweat on the sensor array. After verifying the stability aspects, including the self-healing properties, they successfully visualized the distribution of strain by mapping the current of the sensors that make up the array when the center of the 5 × 5 strain sensor array was pierced with a plastic tip (Figure 7c–e).
As mentioned earlier, in order to improve the reliability and durability of the strain sensor, it is necessary to not only ensure stability to the semiconductor layer but also consider the stress in the interconnection between the components that make up the device. Li et al. investigated the stress distribution throughout the strain sensor through simulation when a horizontally increasing strain was applied to a resistive strain sensor consisting of an elastic sensitive region and two electrodes at both ends (Figure 7f) [102]. A noticeable stress peak occurred at the interface between the elastic-sensitive region and the electrode, which means that in the strain sensor of the aforementioned structure, the junction between the electrode and the elastic-sensitive region is an unstable element with a high possibility of malfunction. Hence, they introduced an additional interconnected layer to mitigate stress at the contact area while preserving the strain sensing characteristics. Consequently, it was confirmed that the peak stress in the strain sensor was reduced by more than threefold (Figure 7g). Based on this, they introduced a CB-PDMS/Ecoflex interconnected layer into a strain sensor with Ag NW embedded in PDMS elastomer and Ag paste as the sensitive region and electrode, respectively, and secured stability at the contact area. Finally, by integrating a strain sensor into the flexible active-matrix OLED display, they could sense the distribution of strain according to the bending of the display and through this, implemented an interactive surface (Figure 7h).
Zhao et al. developed a multi-modal sensor capable of simultaneously detecting both strain and relative humidity [103]. They applied and integrated graphene and MoS2 to sense strain and relative humidity, respectively. Even though the two sensing functions were performed simultaneously, mutual interference was avoided because graphene was insensitive to humidity and, conversely, MoS2 had a negligible response to humidity. They attached the proposed multimodal sensor to a mask to measure human breathing frequency, conducting measurements both before and after exercise. During this process, they could extract the breathing frequency by analyzing the cycle in which the peak of resistance changes for the measured strain and humidity occurred. Additionally, it was possible to observe changes in humidity attributed to exhalation. Consequently, while enhancing the performance of sensors such as sensitivity and response time is crucial, expanding the capability to sense multiple elements within the same area represents a direction for further advancement in the sensor field.
Figure 7. Active-matrix array-based strain sensors. (a) Structural schematic diagram of an active-matrix strain sensor array based on a piezopotential-gated graphene transistor. (b) 2D mapping of the output current of each pixel comprising the strain sensor array when an external strain is applied (adapted from [99] with permission from John Wiley and Sons). (c) A photograph of a self-healable semiconducting polymer-based active-matrix strain sensor being poked with a plastic tip to evaluate its properties, and (d) the normalized on current of the strain sensor array at that time and (e) strain mapping obtained through simulation (adapted from [100] with permission from American Association for the Advancement of Science). When applying a horizontally increasing strain to a resistive strain sensor consisting of an elastic sensitive region and two electrodes at either end, stress distribution in the sensor region (f) with and (g) without an interconnection layer. (h) Mapping of strain distribution through resistance change according to bending operation after integrating the strain sensor with the display (adapted from [102] with permission from John Wiley and Sons). (i) A photo showing the insertion of a sensor capable of simultaneously measuring strain and humidity into a mask to monitor human breathing frequency, intensity, and RH. Changes in resistance over time of (j) strain-sensitive graphene and (k) humidity-sensitive MoS2 caused by human respiration before and after exercise (adapted from [103] with permission from John Wiley and Sons).
Figure 7. Active-matrix array-based strain sensors. (a) Structural schematic diagram of an active-matrix strain sensor array based on a piezopotential-gated graphene transistor. (b) 2D mapping of the output current of each pixel comprising the strain sensor array when an external strain is applied (adapted from [99] with permission from John Wiley and Sons). (c) A photograph of a self-healable semiconducting polymer-based active-matrix strain sensor being poked with a plastic tip to evaluate its properties, and (d) the normalized on current of the strain sensor array at that time and (e) strain mapping obtained through simulation (adapted from [100] with permission from American Association for the Advancement of Science). When applying a horizontally increasing strain to a resistive strain sensor consisting of an elastic sensitive region and two electrodes at either end, stress distribution in the sensor region (f) with and (g) without an interconnection layer. (h) Mapping of strain distribution through resistance change according to bending operation after integrating the strain sensor with the display (adapted from [102] with permission from John Wiley and Sons). (i) A photo showing the insertion of a sensor capable of simultaneously measuring strain and humidity into a mask to monitor human breathing frequency, intensity, and RH. Changes in resistance over time of (j) strain-sensitive graphene and (k) humidity-sensitive MoS2 caused by human respiration before and after exercise (adapted from [103] with permission from John Wiley and Sons).
Electronics 13 00241 g007

4.5. Temperature Sensors

Hong et al. constructed a temperature sensor array based on a CNT TFT backplane [104]. The temperature sensor array consists of a total of four layers, excluding Ecoflex, which is used as a flexible substrate. The CNT TFT and temperature sensor are positioned on the first and fourth layers, respectively, starting from the bottom (Figure 8a). The gate and source lines for array configuration are located on the second and third layers. Notably, the gate and source lines are constructed using Galinstan, ensuring that the electrical characteristics remain stable even when the Ecoflex board is deformed. Moreover, electrochemically synthesized polyaniline nanofiber film was employed as the temperature sensor channel and had a resistance sensitivity of 1.0% °C−1 in the range from 15 to 45 °C (Figure 8b). Subsequently, following the integration of the temperature sensor with the CNT TFT, a linear change in current corresponding to temperature was observed (Figure 8c). This enabled the calculation of the temperature based on the change in drain current when a finger contacted the 5 × 5 temperature sensor array. Remarkably, the calculated temperature value closely matched the measurement obtained with an infrared thermometer.
Ren et al. implemented a temperature sensor array consisting of a DNTT TFT and a pentacene/Ag NPs thermistor in a 16 × 16 array (Figure 8d) [105]. The constructed temperature sensor array demonstrated the ability to detect temperatures in the range of 20 to 100 °C, achieving a temperature resolution of 0.4 °C. In the case of the DNTT TFT, the change in electrical characteristics in the relevant temperature range was negligible. Instead, only the resistance of the pentacene/Ag NPs thermistor exhibited a sensitive response to temperature. To validate the practicality of the temperature sensor, they placed a 2 cm × 2 cm Peltier heater on the sensor, inducing a temperature change. They successfully performed 2D temperature mapping using the temperature sensor array (Figure 8e). Moreover, by utilizing the flexible properties of poly(ethylene naphthalate) (PEN) used as a substrate, they demonstrated the measurement of body temperature by attaching the sensor to a person’s forehead.
The two previous temperature sensor arrays were implemented by integrating a thermistor with a resistance change sensitive to temperature on a TFT-based backplane that is insensitive to temperature change. In addition, the focus was on implementing a temperature sensor and demonstrating the measurement of the temperature distribution of an object or body temperature. Furthermore, Kim et al. proposed a flexible temperature sensor array designed for low-power operation [106]. To achieve a temperature sensor capable of low-power operation, they used a voltage range corresponding to the subthreshold region of the transistor because the subthreshold region is located near 0 V, which naturally leads to low power consumption. In addition, the characteristic that transport of carriers in the subthreshold region is mainly achieved by temperature-sensitive diffusion could serve as an opportunity to sense temperature at the subthreshold region. The implemented subthreshold transistor-based temperature sensor operated stably even under strain of up to 100%, and its sensitivity to temperature was recorded at a very high 9.4% °C−1. Additionally, they successfully measured the surface temperature of a cold or hot spherical metal ball using the proposed temperature sensor (Figure 8f–h).

5. Summary

In this review, we reported on studies that implemented an active-matrix array using TFTs based on emerging materials such as organic semiconductors, metal oxide semiconductors, 2D materials, and CNTs (Table 1 and Table 2). The versatility of the TFT-based active-matrix, which can overcome resolution limitations due to crosstalk issues in the passive-matrix, became the reason for introducing the active-matrix not only in displays but also in various sensors. In addition, TFTs based on the bottom–up process offer a high degree of design flexibility, allowing for freedom in material and structure selection during production. The freedom of choice regarding design parameters allows TFTs to be manufactured not only on rigid substrates such as silicon and glass, but also on flexible substrates based on paper and polymers. As a result, many attempts have been reported to provide flexibility and stretchability to active-matrix-based displays and various sensors. In particular, the flexible and stretchable sensor array produced allows for close contact with surfaces that are not flat, such as human skin. This is used in healthcare to measure biological signals by minimizing the gap between the sensor and the surface to which it is attached. It enables precise measurements in the monitoring system.
There are many advantages to active-matrix using emerging-material-based TFTs, and the following matters need to be considered for further development to reach industrialization in the future. (i) Although it is important to improve the performance of TFTs based on emerging materials integrated into the active-matrix, it is necessary to secure reliability through continuous research and evaluation of operation stability that can maintain stable characteristics in various surrounding environments. (ii) In the case of active-matrix sensors, development is needed to detect two or more factors within the same area. This will be important in the development of gas sensors capable of detecting and identifying multiple gases, especially in the field of gas sensors that are less scalable to array scale. (iii) In addition, in order to advance to the industrialization stage, technology for manufacturing a high-resolution active-matrix array must be secured through improvements in the fine process level to improve the integration of TFTs based on emerging materials. As has been the case so far, the development of emerging-material-based TFT and active-matrix-based displays and sensors will continue to be implemented by many researchers, and it is clear that it will have a beneficial impact on our lives in the future.

Author Contributions

S.K. performed literature research, analysis, and wrote the paper. H.Y. initiated and supervised the work and wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This research was partly supported by the Technology Innovation Program (1415181712, Interface Technology of 3D Stacked Heterogeneous System for SCM-based Process-in-Memory) funded by the Ministry of Trade, Industry & Energy (MOTIE, Korea). This work was supported in part by the Gachon University Research Fund of 2023 (GCU-202307950001).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Zhu, H.; Shin, E.S.; Liu, A.; Ji, D.; Xu, Y.; Noh, Y.Y. Printable semiconductors for backplane TFTs of flexible OLED displays. Adv. Funct. Mater. 2020, 30, 1904588. [Google Scholar] [CrossRef]
  2. Wu, W.-J.; Chen, J.-W.; Wang, J.-S.; Zhou, L.; Tao, H.; Zou, J.-H.; Xu, M.; Wang, L.; Peng, J.-B.; Chan, M. High-resolution flexible AMOLED display integrating gate driver by metal–oxide TFTs. IEEE Electron Device Lett. 2018, 39, 1660–1663. [Google Scholar] [CrossRef]
  3. Meng, W.; Xu, F.; Yu, Z.; Tao, T.; Shao, L.; Liu, L.; Li, T.; Wen, K.; Wang, J.; He, L. Three-dimensional monolithic micro-LED display driven by atomically thin transistor matrix. Nat. Nanotechnol. 2021, 16, 1231–1236. [Google Scholar] [CrossRef] [PubMed]
  4. Lei, T.; Shao, L.-L.; Zheng, Y.-Q.; Pitner, G.; Fang, G.; Zhu, C.; Li, S.; Beausoleil, R.; Wong, H.-S.P.; Huang, T.-C. Low-voltage high-performance flexible digital and analog circuits based on ultrahigh-purity semiconducting carbon nanotubes. Nat. Commun. 2019, 10, 2161. [Google Scholar] [CrossRef] [PubMed]
  5. Janneck, R.; Nowack, T.S.; De Roose, F.; Ali, H.; Dehaene, W.; Heremans, P.; Genoe, J.; Rolin, C. Integration of highly crystalline C8-BTBT thin-films into simple logic gates and circuits. Org. Electron. 2019, 67, 64–71. [Google Scholar] [CrossRef]
  6. Andersson Ersman, P.; Lassnig, R.; Strandberg, J.; Tu, D.; Keshmiri, V.; Forchheimer, R.; Fabiano, S.; Gustafsson, G.; Berggren, M. All-printed large-scale integrated circuits based on organic electrochemical transistors. Nat. Commun. 2019, 10, 5053. [Google Scholar] [CrossRef] [PubMed]
  7. Shim, J.; woon Jang, S.; Lim, J.-H.; Kim, H.; Kang, D.-H.; Kim, K.-H.; Seo, S.; Heo, K.; Shin, C.; Yu, H.-Y. Polarity control in a single transition metal dichalcogenide (TMD) transistor for homogeneous complementary logic circuits. Nanoscale 2019, 11, 12871–12877. [Google Scholar] [CrossRef] [PubMed]
  8. Polyushkin, D.K.; Wachter, S.; Mennel, L.; Paur, M.; Paliy, M.; Iannaccone, G.; Fiori, G.; Neumaier, D.; Canto, B.; Mueller, T. Analogue two-dimensional semiconductor electronics. Nat. Electron. 2020, 3, 486–491. [Google Scholar] [CrossRef]
  9. Lu, W.; Lu, C.; Yang, G.; Liu, M.; Chen, K.; Liao, F.; Duan, X.; Lu, N.; Li, L. Monolithically Stacked Two Layers of a-IGZO-Based Transistors Upon a-IGZO-Based Analog/Logic Circuits. IEEE Trans. Electron Devices 2023, 70, 1697–1701. [Google Scholar] [CrossRef]
  10. Xu, Y.; Ruan, C.-P.; Zhou, L.; Zou, J.-H.; Xu, M.; Wu, W.-J.; Wang, L.; Peng, J.-B. A 256 × 256, 50-μm pixel pitch OPD image sensor based on an IZO TFT backplane. IEEE Sens. J. 2021, 21, 20824–20832. [Google Scholar] [CrossRef]
  11. Yokota, T.; Fukuda, K.; Someya, T. Recent progress of flexible image sensors for biomedical applications. Adv. Mater. 2021, 33, 2004416. [Google Scholar] [CrossRef] [PubMed]
  12. Chen, T.; Wang, C.; Yang, G.; Lou, Q.; Lin, Q.; Zhang, S.; Zhou, H. Monolithic Integration of Perovskite Photoabsorbers with IGZO Thin-Film Transistor Backplane for Phototransistor-Based Image Sensor. Adv. Mater. Technol. 2023, 8, 2200679. [Google Scholar] [CrossRef]
  13. Luo, Z.; Peng, B.; Zeng, J.; Yu, Z.; Zhao, Y.; Xie, J.; Lan, R.; Ma, Z.; Pan, L.; Cao, K. Sub-thermionic, ultra-high-gain organic transistors and circuits. Nat. Commun. 2021, 12, 1928. [Google Scholar] [CrossRef] [PubMed]
  14. Casula, G.; Lai, S.; Matino, L.; Santoro, F.; Bonfiglio, A.; Cosseddu, P. Printed, low-voltage, all-organic transistors and complementary circuits on paper substrate. Adv. Electron. Mater. 2020, 6, 1901027. [Google Scholar] [CrossRef]
  15. Jiang, C.; Choi, H.W.; Cheng, X.; Ma, H.; Hasko, D.; Nathan, A. Printed subthreshold organic transistors operating at high gain and ultralow power. Science 2019, 363, 719–723. [Google Scholar] [CrossRef] [PubMed]
  16. Shin, H.-S.; Yoo, H.; Kim, C.-H. Revealing Three-in-One Nature of Organic Negative Transconductance Transistors. IEEE Trans. Electron Devices 2022, 69, 5149–5154. [Google Scholar] [CrossRef]
  17. Kim, C.-H. Contact resistance in organic transistors: Use it or remove it. Appl. Phys. Rev. 2020, 7, 031306. [Google Scholar] [CrossRef]
  18. Carlos, E.; Leppäniemi, J.; Sneck, A.; Alastalo, A.; Deuermeier, J.; Branquinho, R.; Martins, R.; Fortunato, E. Printed, highly stable metal oxide thin-film transistors with ultra-thin high-κ oxide dielectric. Adv. Electron. Mater. 2020, 6, 1901071. [Google Scholar] [CrossRef]
  19. Zhao, Y.; Wang, Z.; Xu, G.; Cai, L.; Han, T.H.; Zhang, A.; Wu, Q.; Wang, R.; Huang, T.; Cheng, P. High Performance Indium-Gallium-Zinc Oxide Thin Film Transistor via Interface Engineering. Adv. Funct. Mater. 2020, 30, 2003285. [Google Scholar] [CrossRef]
  20. Bukke, R.N.; Mude, N.N.; Saha, J.K.; Jang, J. High performance of a-IZTO TFT by purification of the semiconductor oxide precursor. Adv. Mater. Interfaces 2019, 6, 1900277. [Google Scholar] [CrossRef]
  21. Li, G.; Wu, Y.; Li, Y.; Hong, Y.; Zhao, X.; Reyes, P.I.; Lu, Y. Early stage detection of Staphylococcus epidermidis biofilm formation using MgZnO dual-gate TFT biosensor. Biosens. Bioelectron. 2020, 151, 111993. [Google Scholar] [CrossRef]
  22. Das, S.; Sebastian, A.; Pop, E.; McClellan, C.J.; Franklin, A.D.; Grasser, T.; Knobloch, T.; Illarionov, Y.; Penumatcha, A.V.; Appenzeller, J. Transistors based on two-dimensional materials for future integrated circuits. Nat. Electron. 2021, 4, 786–799. [Google Scholar] [CrossRef]
  23. Daus, A.; Vaziri, S.; Chen, V.; Köroğlu, Ç.; Grady, R.W.; Bailey, C.S.; Lee, H.R.; Schauble, K.; Brenner, K.; Pop, E. High-performance flexible nanoscale transistors based on transition metal dichalcogenides. Nat. Electron. 2021, 4, 495–501. [Google Scholar] [CrossRef]
  24. Wang, Z.; Wang, P.; Wang, F.; Ye, J.; He, T.; Wu, F.; Peng, M.; Wu, P.; Chen, Y.; Zhong, F. A noble metal dichalcogenide for high-performance field-effect transistors and broadband photodetectors. Adv. Funct. Mater. 2020, 30, 1907945. [Google Scholar] [CrossRef]
  25. Liu, L.; Han, J.; Xu, L.; Zhou, J.; Zhao, C.; Ding, S.; Shi, H.; Xiao, M.; Ding, L.; Ma, Z. Aligned, high-density semiconducting carbon nanotube arrays for high-performance electronics. Science 2020, 368, 850–856. [Google Scholar] [CrossRef] [PubMed]
  26. Lin, Y.; Liang, S.; Xu, L.; Liu, L.; Hu, Q.; Fan, C.; Liu, Y.; Han, J.; Zhang, Z.; Peng, L.M. Enhancement-Mode Field-Effect Transistors and High-Speed Integrated Circuits Based on Aligned Carbon Nanotube Films. Adv. Funct. Mater. 2022, 32, 2104539. [Google Scholar] [CrossRef]
  27. Xu, L.; Qiu, C.; Peng, L.-m.; Zhang, Z. Suppression of leakage current in carbon nanotube field-effect transistors. Nano Res. 2021, 14, 976–981. [Google Scholar] [CrossRef]
  28. Pitner, G.; Zhang, Z.; Lin, Q.; Su, S.-K.; Gilardi, C.; Kuo, C.; Kashyap, H.; Weiss, T.; Yu, Z.; Chao, T.-A. Sub-0.5 nm interfacial dielectric enables superior electrostatics: 65 mV/dec top-gated carbon nanotube FETs at 15 nm gate length. In Proceedings of the 2020 IEEE International Electron Devices Meeting (IEDM), San Francisco, CA, USA, 12–18 December 2020; pp. 3.5.1–3.5.4. [Google Scholar]
  29. Bornemann, S.; Gülink, J.; Moro, V.; Gil, J.C.; Wolter, S.; Schöttler, G.; Bezshlyakh, D.; Prades, J.D.; Dieguez, A.; Waag, A. Processing and characterization of monolithic passive-matrix GaN-based microLED arrays with pixel sizes from 5 to 50 µm. IEEE Photonics J. 2021, 13, 1–9. [Google Scholar] [CrossRef]
  30. Nau, S.; Wolf, C.; Sax, S.; List-Kratochvil, E.J. Organic Non-Volatile Resistive Photo-Switches for Flexible Image Detector Arrays. Adv. Mater. 2015, 27, 1048–1052. [Google Scholar] [CrossRef]
  31. Choi, M.; Jang, B.; Lee, W.; Lee, S.; Kim, T.W.; Lee, H.J.; Kim, J.H.; Ahn, J.H. Stretchable active matrix inorganic light-emitting diode display enabled by overlay-aligned roll-transfer printing. Adv. Funct. Mater. 2017, 27, 1606005. [Google Scholar] [CrossRef]
  32. Park, Y.J.; Sharma, B.K.; Shinde, S.M.; Kim, M.-S.; Jang, B.; Kim, J.-H.; Ahn, J.-H. All MoS2-based large area, skin-attachable active-matrix tactile sensor. ACS Nano 2019, 13, 3023–3030. [Google Scholar] [CrossRef] [PubMed]
  33. Jang, J.; Oh, B.; Jo, S.; Park, S.; An, H.S.; Lee, S.; Cheong, W.H.; Yoo, S.; Park, J.U. Human-interactive, active-matrix displays for visualization of tactile pressures. Adv. Mater. Technol. 2019, 4, 1900082. [Google Scholar] [CrossRef]
  34. Liu, C.; Zhou, H.; Wu, Q.; Dai, F.; Lau, T.K.; Lu, X.; Yang, T.; Wang, Z.; Liu, X.; Liu, C. Guided Formation of Large Crystals of Organic and Perovskite Semiconductors by an Ultrasonicated Dispenser and Their Application as the Active Matrix of Photodetectors. ACS Appl. Mater. Interfaces 2018, 10, 39921–39932. [Google Scholar] [CrossRef] [PubMed]
  35. Wang, Y.; Chen, C.; Zou, T.; Yan, L.; Liu, C.; Du, X.; Zhang, S.; Zhou, H. Spin-On-Patterning of Sn–Pb Perovskite Photodiodes on IGZO Transistor Arrays for Fast Active-Matrix Near-Infrared Imaging. Adv. Mater. Technol. 2019, 5, 1900752. [Google Scholar] [CrossRef]
  36. Xuan, Y.; Lu, Y.; Honda, S.; Arie, T.; Akita, S.; Takei, K. Active-Matrix-Based Flexible Optical Image Sensor. Adv. Mater. Technol. 2021, 6, 2100259. [Google Scholar] [CrossRef]
  37. Ji, S.; Jang, J.; Hwang, J.C.; Lee, Y.; Lee, J.H.; Park, J.U. Amorphous Oxide Semiconductor Transistors with Air Dielectrics for Transparent and Wearable Pressure Sensor Arrays. Adv. Mater. Technol. 2019, 5, 1900928. [Google Scholar] [CrossRef]
  38. Baek, S.; Lee, Y.; Baek, J.; Kwon, J.; Kim, S.; Lee, S.; Strunk, K.P.; Stehlin, S.; Melzer, C.; Park, S.M.; et al. Spatiotemporal Measurement of Arterial Pulse Waves Enabled by Wearable Active-Matrix Pressure Sensor Arrays. ACS Nano 2022, 16, 368–377. [Google Scholar] [CrossRef] [PubMed]
  39. Ma, C.; Xu, D.; Wang, P.; Lin, Z.; Zhou, J.; Jia, C.; Huang, J.; Li, S.; Huang, Y.; Duan, X. Two-dimensional van der Waals thin film transistors as active matrix for spatially resolved pressure sensing. Nano Res. 2021, 14, 3395–3401. [Google Scholar] [CrossRef]
  40. Rustagi, S.; Singh, N.; Fang, W.; Buddharaju, K.; Omampuliyur, S.; Teo, S.; Tung, C.; Lo, G.; Balasubramanian, N.; Kwong, D. CMOS inverter based on gate-all-around silicon-nanowire MOSFETs fabricated using top-down approach. IEEE Electron Device Lett. 2007, 28, 1021–1024. [Google Scholar] [CrossRef]
  41. Singh, S. Fully solution-processed carbon nanotubes thin film transistors and PMOS inverters on glass substrate. Flexible and Printed Electronics 2023, 8, 015011. [Google Scholar] [CrossRef]
  42. Bhalerao, S.R.; Lupo, D.; Berger, P.R. Flexible, solution-processed, indium oxide (In2O3) thin film transistors (TFT) and circuits for internet-of-things (IoT). Mater. Sci. Semicond. Process. 2022, 139, 106354. [Google Scholar] [CrossRef]
  43. Li, X.; Ren, Y.; Wang, X.; Shao, S.; Li, H.; Wu, L.; Liu, X.; Zhao, J. A universal method for high-efficiency immobilization of semiconducting carbon nanotubes toward fully printed paper-based electronics. Adv. Electron. Mater. 2021, 7, 2001025. [Google Scholar] [CrossRef]
  44. Chen, D.; Xu, Y.; An, Z.; Li, Z.; Zhang, C. Thin-film transistors based on wide bandgap Ga2O3 films grown by aqueous-solution spin-coating method. Micro Nano Lett. 2019, 14, 1052–1055. [Google Scholar] [CrossRef]
  45. Ruzgar, S.; Caglar, M. The effect of Sn on electrical performance of zinc oxide based thin film transistor. J. Mater. Sci. Mater. Electron. 2019, 30, 485–490. [Google Scholar] [CrossRef]
  46. Shao, F.; Wan, Q. Recent progress on jet printing of oxide-based thin film transistors. J. Phys. D Appl. Phys. 2019, 52, 143002. [Google Scholar] [CrossRef]
  47. Singh, S.; Takeda, Y.; Matsui, H.; Tokito, S. Flexible PMOS inverter and NOR gate using inkjet-printed dual-gate organic thin film transistors. IEEE Electron Device Lett. 2020, 41, 409–412. [Google Scholar] [CrossRef]
  48. Zschieschang, U.; Klauk, H.; Borchert, J.W. High-Resolution Lithography for High-Frequency Organic Thin-Film Transistors. Adv. Mater. Technol. 2023, 8, 2201888. [Google Scholar] [CrossRef]
  49. Cui, N.; Ren, H.; Tang, Q.; Zhao, X.; Tong, Y.; Hu, W.; Liu, Y. Fully transparent conformal organic thin-film transistor array and its application as LED front driving. Nanoscale 2018, 10, 3613–3620. [Google Scholar] [CrossRef] [PubMed]
  50. Kim, S.; Seo, J.; Park, T.; Yoo, H. Interface Trap-Free, 100% Yield, Wafer-Scale, Non-Volatile Optically-Guided Memory Array from Cumulatively-Stacked Small Molecules/Fluoropolymer/Copper-Oxide Nanoparticles Heterostructure. Adv. Electron. Mater. 2022, 8, 2200752. [Google Scholar] [CrossRef]
  51. Kuribara, K.; Wang, H.; Uchiyama, N.; Fukuda, K.; Yokota, T.; Zschieschang, U.; Jaye, C.; Fischer, D.; Klauk, H.; Yamamoto, T. Organic transistors with high thermal stability for medical applications. Nat. Commun. 2012, 3, 723. [Google Scholar] [CrossRef]
  52. Cong, S.; Cao, Y.; Fang, X.; Wang, Y.; Liu, Q.; Gui, H.; Shen, C.; Cao, X.; Kim, E.S.; Zhou, C. Carbon nanotube macroelectronics for active matrix polymer-dispersed liquid crystal displays. ACS Nano 2016, 10, 10068–10074. [Google Scholar] [CrossRef]
  53. Ohara, H.; Sasaki, T.; Noda, K.; Ito, S.; Sasaki, M.; Endo, Y.; Yoshitomi, S.; Sakata, J.; Serikawa, T.; Yamazaki, S. 4.0-inch active-matrix organic light-emitting diode display integrated with driver circuits using amorphous In–Ga–Zn-Oxide thin-film transistors with suppressed variation. Jpn. J. Appl. Phys. 2010, 49, 03CD02. [Google Scholar] [CrossRef]
  54. Mizukami, M.; Oku, S.; Cho, S.-I.; Tatetsu, M.; Abiko, M.; Mamada, M.; Sakanoue, T.; Suzuri, Y.; Kido, J.; Tokito, S. A solution-processed organic thin-film transistor backplane for flexible multiphoton emission organic light-emitting diode displays. IEEE Electron Device Lett. 2015, 36, 841–843. [Google Scholar] [CrossRef]
  55. Zhao, T.-Y.; Zhang, D.-D.; Qu, T.-Y.; Fang, L.-L.; Zhu, Q.-B.; Sun, Y.; Cai, T.-H.; Chen, M.-L.; Wang, B.-W.; Du, J.-H. Flexible 64× 64 pixel AMOLED displays driven by uniform carbon nanotube thin-film transistors. ACS Appl. Mater. Interfaces 2019, 11, 11699–11705. [Google Scholar] [CrossRef] [PubMed]
  56. Lee, D.; Cho, E.-S.; Jeon, Y.; Kwon, S.J. Characterization of the material and electrical properties depending on the Mg: Ag ratio as a cathode for TEOLED. Mater. Chem. Phys. 2023, 303, 127742. [Google Scholar] [CrossRef]
  57. Park, J.H.; Shin, Y.J.; Kymissis, I.; Jeon, Y.; Kim, C.-H. Frequency-triggered circuit transition in organic light-emitting diodes probed by impedance spectroscopy. J. Mater. Chem. C 2023, 11, 9670–9677. [Google Scholar] [CrossRef]
  58. Hwangbo, S.; Hu, L.; Hoang, A.T.; Choi, J.Y.; Ahn, J.-H. Wafer-scale monolithic integration of full-colour micro-LED display using MoS2 transistor. Nat. Nanotechnol. 2022, 17, 500–506. [Google Scholar] [CrossRef] [PubMed]
  59. Ma, T.; Chen, J.; Chen, Z.; Liang, L.; Hu, J.; Shen, W.; Li, Z.; Zeng, H. Progress in Color Conversion Technology for Micro-LED. Adv. Mater. Technol. 2023, 8, 2200632. [Google Scholar] [CrossRef]
  60. Li, Y.; He, P.; Chen, S.; Lan, L.; Dai, X.; Peng, J. Inkjet-printed oxide thin-film transistors based on nanopore-free aqueous-processed dielectric for active-matrix quantum-dot light-emitting diode displays. ACS Appl. Mater. Interfaces 2019, 11, 28052–28059. [Google Scholar] [CrossRef] [PubMed]
  61. Choi, M.K.; Yang, J.; Kang, K.; Kim, D.C.; Choi, C.; Park, C.; Kim, S.J.; Chae, S.I.; Kim, T.-H.; Kim, J.H. Wearable red–green–blue quantum dot light-emitting diode array using high-resolution intaglio transfer printing. Nat. Commun. 2015, 6, 7149. [Google Scholar] [CrossRef]
  62. Peng, B.; Ren, X.; Wang, Z.; Wang, X.; Roberts, R.C.; Chan, P.K. High performance organic transistor active-matrix driver developed on paper substrate. Sci. Rep. 2014, 4, 6430. [Google Scholar] [CrossRef]
  63. Um, J.G.; Jeong, D.Y.; Jung, Y.; Moon, J.K.; Jung, Y.H.; Kim, S.; Kim, S.H.; Lee, J.S.; Jang, J. Active-Matrix GaN µ-LED Display Using Oxide Thin-Film Transistor Backplane and Flip Chip LED Bonding. Adv. Electron. Mater. 2019, 5, 1800617. [Google Scholar] [CrossRef]
  64. Choi, M.; Park, Y.J.; Sharma, B.K.; Bae, S.-R.; Kim, S.Y.; Ahn, J.-H. Flexible active-matrix organic light-emitting diode display enabled by MoS2 thin-film transistor. Sci. Adv. 2018, 4, eaas8721. [Google Scholar] [CrossRef]
  65. Baek, G.W.; Seo, S.G.; Hahm, D.; Kim, Y.J.; Kim, K.; Lee, T.; Kim, J.; Bae, W.K.; Jin, S.H.; Kwak, J. Optimum Design Configuration of Thin-Film Transistors And Quantum-Dot Light-Emitting Diodes for Active-Matrix Displays. Adv. Mater. 2023, 35, 2304717. [Google Scholar] [CrossRef]
  66. Kaminishi, D.; Ozaki, H.; Ohno, Y.; Maehashi, K.; Inoue, K.; Matsumoto, K.; Seri, Y.; Masuda, A.; Matsumura, H. Air-stable n-type carbon nanotube field-effect transistors with Si3N4 passivation films fabricated by catalytic chemical vapor deposition. Appl. Phys. Lett. 2005, 86, 113115. [Google Scholar] [CrossRef]
  67. Filiatrault, H.L.; Porteous, G.C.; Carmichael, R.S.; Davidson, G.J.; Carmichael, T.B. Stretchable light-emitting electrochemical cells using an elastomeric emissive material. Adv. Mater. 2012, 24, 2673–2678. [Google Scholar] [CrossRef] [PubMed]
  68. Liang, J.; Li, L.; Niu, X.; Yu, Z.; Pei, Q. Elastomeric polymer light-emitting devices and displays. Nat. Photonics 2013, 7, 817–824. [Google Scholar] [CrossRef]
  69. Liang, J.; Chu, M.; Zhou, Z.; Yan, Y.; Zhao, Y.S. Optically pumped lasing in microscale light-emitting electrochemical cell arrays for multicolor displays. Nano Lett. 2020, 20, 7116–7122. [Google Scholar] [CrossRef] [PubMed]
  70. Meier, S.B.; Tordera, D.; Pertegas, A.; Roldan-Carmona, C.; Orti, E.; Bolink, H.J. Light-emitting electrochemical cells: Recent progress and future prospects. Mater. Today 2014, 17, 217–223. [Google Scholar] [CrossRef]
  71. Liu, J.; Wang, J.; Zhang, Z.; Molina-Lopez, F.; Wang, G.-J.N.; Schroeder, B.C.; Yan, X.; Zeng, Y.; Zhao, O.; Tran, H. Fully stretchable active-matrix organic light-emitting electrochemical cell array. Nat. Commun. 2020, 11, 3362. [Google Scholar] [CrossRef]
  72. Hong, S.; Choi, S.H.; Park, J.; Yoo, H.; Oh, J.Y.; Hwang, E.; Yoon, D.H.; Kim, S. Sensory adaptation and neuromorphic phototransistors based on CsPb(Br1–xIx)3 perovskite and MoS2 hybrid structure. ACS Nano 2020, 14, 9796–9806. [Google Scholar] [CrossRef]
  73. Li, P.; Shan, X.; Lin, Y.; Meng, X.; Ma, J.; Wang, Z.; Zhao, X.; Li, B.; Liu, W.; Xu, H. Tin Doping Induced High-Performance Solution-Processed Ga2O3 Photosensor toward Neuromorphic Visual System. Adv. Funct. Mater. 2023, 33, 2303584. [Google Scholar] [CrossRef]
  74. Hong, S.; Baek, S.; Can, T.T.T.; Choi, W.S.; Kim, S. Fabrication of Highly Photosensitive MoS2 Photodetector Films Using Rapid Electrohydrodynamic-Jet Printing Process. Adv. Electron. Mater. 2022, 8, 2101063. [Google Scholar] [CrossRef]
  75. Li, M.; Zheng, J.; Wang, X.; Yu, R.; Wang, Y.; Qiu, Y.; Cheng, X.; Wang, G.; Chen, G.; Xie, K. Light-responsive self-strained organic semiconductor for large flexible OFET sensing array. Nat. Commun. 2022, 13, 4912. [Google Scholar] [CrossRef] [PubMed]
  76. Tang, Y.; Zhao, Y.; Liu, H. Room-temperature semiconductor gas sensors: Challenges and opportunities. ACS Sens. 2022, 7, 3582–3597. [Google Scholar] [CrossRef]
  77. Shaji, M.; Saji, K.; Jayaraj, M. Low temperature operated ZTO thin film transistor based gas sensor for selective detection of H2S. Mater. Sci. Semicond. Process. 2022, 150, 106927. [Google Scholar] [CrossRef]
  78. Kwon, H.; Yoo, H.; Nakano, M.; Takimiya, K.; Kim, J.-J.; Kim, J.K. Gate-tunable gas sensing behaviors in air-stable ambipolar organic thin-film transistors. RSC Adv. 2020, 10, 1910–1916. [Google Scholar] [CrossRef]
  79. Someya, T.; Sekitani, T.; Iba, S.; Kato, Y.; Kawaguchi, H.; Sakurai, T. A large-area, flexible pressure sensor matrix with organic field-effect transistors for artificial skin applications. Proc. Natl. Acad. Sci. USA 2004, 101, 9966–9970. [Google Scholar] [CrossRef] [PubMed]
  80. Hassinen, T.; Eiroma, K.; Mäkelä, T.; Ermolov, V. Printed pressure sensor matrix with organic field-effect transistors. Sens. Actuators A Phys. 2015, 236, 343–348. [Google Scholar] [CrossRef]
  81. Trung, T.Q.; Tien, N.T.; Kim, D.; Jang, M.; Yoon, O.J.; Lee, N.E. A flexible reduced graphene oxide field-effect transistor for ultrasensitive strain sensing. Adv. Funct. Mater. 2014, 24, 117–124. [Google Scholar] [CrossRef]
  82. Dubey, P.K.; Yogeswaran, N.; Liu, F.; Vilouras, A.; Kaushik, B.K.; Dahiya, R. Monolayer MoSe₂-based tunneling field effect transistor for ultrasensitive strain sensing. IEEE Trans. Electron Devices 2020, 67, 2140–2146. [Google Scholar] [CrossRef]
  83. Wang, H.; Tong, Y.; Zhao, X.; Tang, Q.; Liu, Y. Flexible, high-sensitive, and wearable strain sensor based on organic crystal for human motion detection. Org. Electron. 2018, 61, 304–311. [Google Scholar] [CrossRef]
  84. Zhu, C.; Chortos, A.; Wang, Y.; Pfattner, R.; Lei, T.; Hinckley, A.C.; Pochorovski, I.; Yan, X.; To, J.W.-F.; Oh, J.Y. Stretchable temperature-sensing circuits with strain suppression based on carbon nanotube transistors. Nat. Electron. 2018, 1, 183–190. [Google Scholar] [CrossRef]
  85. Seo, H.; Aihara, S.; Watabe, T.; Ohtake, H.; Sakai, T.; Kubota, M.; Egami, N.; Hiramatsu, T.; Matsuda, T.; Furuta, M.; et al. A 128 × 96 Pixel Stack-Type Color Image Sensor: Stack of Individual Blue-, Green-, and Red-Sensitive Organic Photoconductive Films Integrated with a ZnO Thin Film Transistor Readout Circuit. Jpn. J. Appl. Phys. 2011, 50, 024103. [Google Scholar] [CrossRef]
  86. Park, T.; Hur, J. Self-powered low-cost UVC sensor based on organic-inorganic heterojunction for partial discharge detection. Small 2021, 17, 2100695. [Google Scholar] [CrossRef]
  87. Zhang, Z.-X.; Li, C.; Lu, Y.; Tong, X.-W.; Liang, F.-X.; Zhao, X.-Y.; Wu, D.; Xie, C.; Luo, L.-B. Sensitive deep ultraviolet photodetector and image sensor composed of inorganic lead-free Cs3Cu2I5 perovskite with wide bandgap. J. Phys. Chem. Lett. 2019, 10, 5343–5350. [Google Scholar] [CrossRef] [PubMed]
  88. Zhang, Y.; Peng, M.; Liu, Y.; Zhang, T.; Zhu, Q.; Lei, H.; Liu, S.; Tao, Y.; Li, L.; Wen, Z. Flexible self-powered real-time ultraviolet photodetector by coupling triboelectric and photoelectric effects. ACS Appl. Mater. Interfaces 2020, 12, 19384–19392. [Google Scholar] [CrossRef]
  89. Li, D.; Jia, Z.; Tang, Y.; Song, C.; Liang, K.; Ren, H.; Li, F.; Chen, Y.; Wang, Y.; Lu, X.; et al. Inorganic-Organic Hybrid Phototransistor Array with Enhanced Photogating Effect for Dynamic Near-Infrared Light Sensing and Image Preprocessing. Nano Lett. 2022, 22, 5434–5442. [Google Scholar] [CrossRef]
  90. Li, D.; Du, J.; Tang, Y.; Liang, K.; Wang, Y.; Ren, H.; Wang, R.; Meng, L.; Zhu, B.; Li, Y. Flexible and Air-Stable Near-Infrared Sensors Based on Solution-Processed Inorganic–Organic Hybrid Phototransistors. Adv. Funct. Mater. 2021, 31, 2105887. [Google Scholar] [CrossRef]
  91. Takahashi, T.; Yu, Z.; Chen, K.; Kiriya, D.; Wang, C.; Takei, K.; Shiraki, H.; Chen, T.; Ma, B.; Javey, A. Carbon nanotube active-matrix backplanes for mechanically flexible visible light and X-ray imagers. Nano Lett. 2013, 13, 5425–5430. [Google Scholar] [CrossRef]
  92. Kim, J.; Jo, C.; Kim, M.G.; Park, G.S.; Marks, T.J.; Facchetti, A.; Park, S.K. Vertically Stacked Full Color Quantum Dots Phototransistor Arrays for High-Resolution and Enhanced Color-Selective Imaging. Adv. Mater. 2022, 34, 2106215. [Google Scholar] [CrossRef] [PubMed]
  93. Ruiz-Preciado, L.A.; Baek, S.; Strobel, N.; Xia, K.; Seiberlich, M.; Park, S.-m.; Lemmer, U.; Jung, S.; Hernandez-Sosa, G. Monolithically printed all-organic flexible photosensor active matrix. NPJ Flex. Electron. 2023, 7, 6. [Google Scholar] [CrossRef]
  94. Kim, S.; Park, H.; Choo, S.; Baek, S.; Kwon, Y.; Liu, N.; Yang, J.Y.; Yang, C.-W.; Yoo, G.; Kim, S. Active-matrix monolithic gas sensor array based on MoS2 thin-film transistors. Commun. Mater. 2020, 1, 86. [Google Scholar] [CrossRef]
  95. Shin, S.-H.; Ji, S.; Choi, S.; Pyo, K.-H.; Wan An, B.; Park, J.; Kim, J.; Kim, J.-Y.; Lee, K.-S.; Kwon, S.-Y. Integrated arrays of air-dielectric graphene transistors as transparent active-matrix pressure sensors for wide pressure ranges. Nat. Commun. 2017, 8, 14950. [Google Scholar] [CrossRef] [PubMed]
  96. Jang, J.; Kim, H.; Ji, S.; Kim, H.J.; Kang, M.S.; Kim, T.S.; Won, J.-e.; Lee, J.-H.; Cheon, J.; Kang, K. Mechanoluminescent, air-dielectric MoS2 transistors as active-matrix pressure sensors for wide detection ranges from footsteps to cellular motions. Nano Lett. 2019, 20, 66–74. [Google Scholar] [CrossRef] [PubMed]
  97. Zhao, Z.; Tang, J.; Yuan, J.; Li, Y.; Dai, Y.; Yao, J.; Zhang, Q.; Ding, S.; Li, T.; Zhang, R. Large-scale integrated flexible tactile sensor Array for sensitive smart robotic touch. ACS Nano 2022, 16, 16784–16795. [Google Scholar] [CrossRef] [PubMed]
  98. Karner-Petritz, E.; Petritz, A.; Uemura, T.; Namba, N.; Araki, T.; Sekitani, T.; Stadlober, B. Ultraflexible Organic Active Matrix Sensor Sheet for Tactile and Biosignal Monitoring. Adv. Electron. Mater. 2023, 9, 2201333. [Google Scholar] [CrossRef]
  99. Sun, Q.; Seung, W.; Kim, B.J.; Seo, S.; Kim, S.W.; Cho, J.H. Active matrix electronic skin strain sensor based on piezopotential-powered graphene transistors. Adv. Mater. 2015, 27, 3411–3417. [Google Scholar] [CrossRef]
  100. Oh, J.Y.; Son, D.; Katsumata, T.; Lee, Y.; Kim, Y.; Lopez, J.; Wu, H.-C.; Kang, J.; Park, J.; Gu, X. Stretchable self-healable semiconducting polymer film for active-matrix strain-sensing array. Sci. Adv. 2019, 5, eaav3097. [Google Scholar] [CrossRef]
  101. Li, C.-H.; Wang, C.; Keplinger, C.; Zuo, J.-L.; Jin, L.; Sun, Y.; Zheng, P.; Cao, Y.; Lissel, F.; Linder, C. A highly stretchable autonomous self-healing elastomer. Nat. Chem. 2016, 8, 618–624. [Google Scholar] [CrossRef]
  102. Li, M.; Chen, S.; Fan, B.; Wu, B.; Guo, X. Printed flexible strain sensor array for bendable interactive surface. Adv. Funct. Mater. 2020, 30, 2003214. [Google Scholar] [CrossRef]
  103. Zhao, J.; Wei, Z.; Li, Z.; Yu, J.; Tang, J.; Zhang, G.; Wang, Z. Skin-Inspired High-Performance Active-matrix Circuitry for Multimodal User-Interaction. Adv. Funct. Mater. 2021, 31, 2105480. [Google Scholar] [CrossRef]
  104. Hong, S.Y.; Lee, Y.H.; Park, H.; Jin, S.W.; Jeong, Y.R.; Yun, J.; You, I.; Zi, G.; Ha, J.S. Stretchable active matrix temperature sensor array of polyaniline nanofibers for electronic skin. Adv. Mater. 2016, 28, 930–935. [Google Scholar] [CrossRef] [PubMed]
  105. Ren, X.; Pei, K.; Peng, B.; Zhang, Z.; Wang, Z.; Wang, X.; Chan, P.K. A low-operating-power and flexible active-matrix organic-transistor temperature-sensor array. Adv. Mater. 2016, 28, 4832–4838. [Google Scholar] [CrossRef] [PubMed]
  106. Kim, J.S.; Jeong, M.W.; Nam, T.U.; Vo, N.T.P.; Jung, K.H.; Lee, T.I.; Oh, J.Y. Intrinsically Stretchable Subthreshold Organic Transistors for Highly Sensitive Low-Power Skin-Like Active-Matrix Temperature Sensors. Adv. Funct. Mater. 2023, 34, 2305252. [Google Scholar] [CrossRef]
Figure 1. Active-matrix array-based liquid crystal display. (a) Transparent and (b) opaque states of PDLC pixels depending on whether or not voltage is applied to the electrode. (c) Cross-sectional view of a unit pixel of a PDLC display driven by CNT-TFT. (d) Optical image of the CNT-TFT-based back panel before PDLC integration and SEM image of the CNT-TFT. (e) Transmittance for wavelengths in transparent and opaque states depending on whether voltage is applied. (f) Demonstration of displaying the numbers “1”, “2”, and “3” on blue, green, and red backgrounds, respectively, by selectively making pixels transparent (adapted from [52] with permission from American Chemical Society).
Figure 1. Active-matrix array-based liquid crystal display. (a) Transparent and (b) opaque states of PDLC pixels depending on whether or not voltage is applied to the electrode. (c) Cross-sectional view of a unit pixel of a PDLC display driven by CNT-TFT. (d) Optical image of the CNT-TFT-based back panel before PDLC integration and SEM image of the CNT-TFT. (e) Transmittance for wavelengths in transparent and opaque states depending on whether voltage is applied. (f) Demonstration of displaying the numbers “1”, “2”, and “3” on blue, green, and red backgrounds, respectively, by selectively making pixels transparent (adapted from [52] with permission from American Chemical Society).
Electronics 13 00241 g001
Figure 2. Active-matrix array-based light emitting diodes. (a) Cross-sectional view of a unit pixel of an OTFT-based active-matrix driver integrated with an LED. (b) Circuit scheme constructed to drive the array and (c) demonstration of display of text images (adapted from [62] with permission from Springer Nature). (d) Cross-sectional schematic diagram of a TFT-μ-LED pixel integrated through flip chip bonding (adapted from [63] with permission from John Wiley and Sons). (e) Driving a flexible OLED display based on a backplane circuit composed of MoS2 attached to a human wrist (adapted from [64] with permission from American Association for the Advancement of Science). (f) Four different configurations to compare the characteristics of each combination of QLED structure and CNT TFT type (adapted from [65] with permission from John Wiley and Sons).
Figure 2. Active-matrix array-based light emitting diodes. (a) Cross-sectional view of a unit pixel of an OTFT-based active-matrix driver integrated with an LED. (b) Circuit scheme constructed to drive the array and (c) demonstration of display of text images (adapted from [62] with permission from Springer Nature). (d) Cross-sectional schematic diagram of a TFT-μ-LED pixel integrated through flip chip bonding (adapted from [63] with permission from John Wiley and Sons). (e) Driving a flexible OLED display based on a backplane circuit composed of MoS2 attached to a human wrist (adapted from [64] with permission from American Association for the Advancement of Science). (f) Four different configurations to compare the characteristics of each combination of QLED structure and CNT TFT type (adapted from [65] with permission from John Wiley and Sons).
Electronics 13 00241 g002
Figure 3. Active-matrix array-based light emitting electrochemical cell. (a) Cross-linking process with the dielectric layer through a cross-linker mixed in the semiconductor conjugated polymer to prevent the semiconductor layer from delamination the gate dielectric under stretching. (b) Photograph of a fully stretchable OTFT active-matrix array. (c) Strain testing from 0% to 100% of the stretchable transistor array. (d) Schematic diagram of vertical integration of active-matrix OTFT array and LEC array. Demonstration of operation of an active matrix OLEC array (e) with and (f) without twisting, respectively (adapted from [71] with permission from Springer Nature).
Figure 3. Active-matrix array-based light emitting electrochemical cell. (a) Cross-linking process with the dielectric layer through a cross-linker mixed in the semiconductor conjugated polymer to prevent the semiconductor layer from delamination the gate dielectric under stretching. (b) Photograph of a fully stretchable OTFT active-matrix array. (c) Strain testing from 0% to 100% of the stretchable transistor array. (d) Schematic diagram of vertical integration of active-matrix OTFT array and LEC array. Demonstration of operation of an active matrix OLEC array (e) with and (f) without twisting, respectively (adapted from [71] with permission from Springer Nature).
Electronics 13 00241 g003
Figure 4. Active-matrix array-based photosensors. (a) Flexible imager integrated with a GOS film that converts X-rays into green light and (b) a cross-sectional view of one pixel. (c) Spatial mapping via an X-ray imager for irradiation of a circular X-ray source (adapted from [91] with permission from American Chemical Society). (d) Illustration of a hybrid phototransistor structure for NIR detection. (e) Scheme image in which the phototransistor array is irradiated with NIR and the corresponding (f) normalized photocurrent mapping of the phototransistor array (adapted from [89] with permission from American Chemical Society). (g) Schematic diagram of the structure of the vertically stacked QD/a-IGZO phototransistor and (h) transfer curve when irradiated with red, green, and blue light (adapted from [92] with permission from John Wiley and Sons). (i) A 10 × 10 active-matrix array based on an all-organic flexible photosensor, and (j) schematic of transmission mode-based PPG sensing and (k) measured pulse response (adapted from [93] with permission from Springer Nature).
Figure 4. Active-matrix array-based photosensors. (a) Flexible imager integrated with a GOS film that converts X-rays into green light and (b) a cross-sectional view of one pixel. (c) Spatial mapping via an X-ray imager for irradiation of a circular X-ray source (adapted from [91] with permission from American Chemical Society). (d) Illustration of a hybrid phototransistor structure for NIR detection. (e) Scheme image in which the phototransistor array is irradiated with NIR and the corresponding (f) normalized photocurrent mapping of the phototransistor array (adapted from [89] with permission from American Chemical Society). (g) Schematic diagram of the structure of the vertically stacked QD/a-IGZO phototransistor and (h) transfer curve when irradiated with red, green, and blue light (adapted from [92] with permission from John Wiley and Sons). (i) A 10 × 10 active-matrix array based on an all-organic flexible photosensor, and (j) schematic of transmission mode-based PPG sensing and (k) measured pulse response (adapted from [93] with permission from Springer Nature).
Electronics 13 00241 g004
Figure 5. Active-matrix array-based gas sensor. (a) Image of MoS2-based 7 × 6 active-matrix gas sensor, and (b) circuit schematic diagram and image consisting of switching TFT and sensing TFT. (c) Change in transfer curve of MoS2 TFT according to NO2 concentration. (d) Comparison of responses of nine MoS2 gas sensors to the same NO2 concentration (adapted from [94] with permission from Springer Nature).
Figure 5. Active-matrix array-based gas sensor. (a) Image of MoS2-based 7 × 6 active-matrix gas sensor, and (b) circuit schematic diagram and image consisting of switching TFT and sensing TFT. (c) Change in transfer curve of MoS2 TFT according to NO2 concentration. (d) Comparison of responses of nine MoS2 gas sensors to the same NO2 concentration (adapted from [94] with permission from Springer Nature).
Electronics 13 00241 g005
Figure 6. Active-matrix array-based pressure sensors. (a) Image of transparent pressure sensor based on air-dielectric graphene transistors. (b) Transfer curve of air-dielectric graphene transistors under various pressures and (c) measurement of change in drain current in response to real-time pressure change (adapted from [95] with permission from Springer Nature). (d) Illustration of MoS2-based pressure sensor array with integrated phosphor particles with mechanoluminescence behavior. (e) Comparison of drain current change of MoS2-based pressure sensor for pressure with and without phosphor particles (adapted from [96] with permission from American Chemical Society). (f) 64 × 64 active-matrix pressure sensor array composed of CNT TFTs and (g) cross-sectional schematic of unit pixels constituting the active-matrix pressure sensor array. (h) Current mapping through sensed pressure when an artificial bee is placed on a pressure sensor array (adapted from [97] with permission from American Chemical Society). (i) Schematic diagram of a pressure sensor array in which each pixel consists of a P(VDF:TrFE) ferroelectric transducer and DNTT OTFT. (j) Demonstration of pulse wave measurement by attaching a ferroelectric pressure sensor array to human skin (adapted from [98] with permission from John Wiley and Sons).
Figure 6. Active-matrix array-based pressure sensors. (a) Image of transparent pressure sensor based on air-dielectric graphene transistors. (b) Transfer curve of air-dielectric graphene transistors under various pressures and (c) measurement of change in drain current in response to real-time pressure change (adapted from [95] with permission from Springer Nature). (d) Illustration of MoS2-based pressure sensor array with integrated phosphor particles with mechanoluminescence behavior. (e) Comparison of drain current change of MoS2-based pressure sensor for pressure with and without phosphor particles (adapted from [96] with permission from American Chemical Society). (f) 64 × 64 active-matrix pressure sensor array composed of CNT TFTs and (g) cross-sectional schematic of unit pixels constituting the active-matrix pressure sensor array. (h) Current mapping through sensed pressure when an artificial bee is placed on a pressure sensor array (adapted from [97] with permission from American Chemical Society). (i) Schematic diagram of a pressure sensor array in which each pixel consists of a P(VDF:TrFE) ferroelectric transducer and DNTT OTFT. (j) Demonstration of pulse wave measurement by attaching a ferroelectric pressure sensor array to human skin (adapted from [98] with permission from John Wiley and Sons).
Electronics 13 00241 g006
Figure 8. Active-matrix array-based temperature sensors. (a) Schematic diagram of a polyaniline nanofiber temperature sensor array with CNT TFT backplane consisting of a total of four layers. (b) Change in the normalized resistance of polyaniline nanofiber-based temperature sensor according to temperature. (c) Change in the normalized drain current of the sensor after integration of the CNT TFT and temperature sensor (adapted from [104] with permission from John Wiley and Sons). (d) Circuit diagram and device structure of the unit temperature sensor constituting the flexible temperature sensor array. (e) Schematic diagram of heated temperature distribution measurement and 2D temperature mapping via Peltier heater located on top of the temperature sensor array (adapted from [105] with permission from John Wiley and Sons). (f) Scheme image of measuring the surface temperature of a metal ball using a stretchable temperature sensor array. Three-dimensional mapping through the distribution of normalized drain current extracted from the temperature sensor when the metal ball is (g) cold and (h) hot, respectively (adapted from [106] with permission from John Wiley and Sons).
Figure 8. Active-matrix array-based temperature sensors. (a) Schematic diagram of a polyaniline nanofiber temperature sensor array with CNT TFT backplane consisting of a total of four layers. (b) Change in the normalized resistance of polyaniline nanofiber-based temperature sensor according to temperature. (c) Change in the normalized drain current of the sensor after integration of the CNT TFT and temperature sensor (adapted from [104] with permission from John Wiley and Sons). (d) Circuit diagram and device structure of the unit temperature sensor constituting the flexible temperature sensor array. (e) Schematic diagram of heated temperature distribution measurement and 2D temperature mapping via Peltier heater located on top of the temperature sensor array (adapted from [105] with permission from John Wiley and Sons). (f) Scheme image of measuring the surface temperature of a metal ball using a stretchable temperature sensor array. Three-dimensional mapping through the distribution of normalized drain current extracted from the temperature sensor when the metal ball is (g) cold and (h) hot, respectively (adapted from [106] with permission from John Wiley and Sons).
Electronics 13 00241 g008
Table 1. Previous research on active-matrix array-based displays.
Table 1. Previous research on active-matrix array-based displays.
ApplicationTransistor MaterialsEmission MaterialsWavelength BandQuantum EfficiencyArray ScaleRef.
LCDCNT---5 × 5[52]
micro-LEDDNTTmicro-LED moduleWhite-8 × 8[62]
micro-LEDa-IGZOGaN456 nm-384 × 128[63]
OLEDMoS2Alq3/C545T--6 × 6[64]
QDLEDCNTCdSe/CdxZn1−xSe/ZnS QDs609 nm19.1%5 × 5[65]
OLECPII2TSY/ionic elastomer--5 × 5[71]
Table 2. Previous research on various sensor arrays based on active-matrix.
Table 2. Previous research on various sensor arrays based on active-matrix.
Sensor TypeTransistor MaterialsSensor MaterialsSensitivity|Operating Voltage|Array ScaleRef.
PhotoCNTP3HT:PCBM0.15 AW−15 V18 × 18[91]
PhotoIn2O3PTB7-Th/BTPV-4F1393 AW−110 V16 × 16[89]
Photoa-IGZOCdS QD, CdSe QD5.2 × 103 AW−115 V12 × 12[92]
PhotoDPP-DTTP3HT-IDTBR0.356 AW−110 V10 × 10[93]
GasMoS2MoS2-10 V7 × 6[94]
PressureGrapheneGraphene2.05 × 10−4 kPa−130 V12 × 12[95]
PressureMoS2MoS20.045 Mpa−160 V20 × 20[96]
PressureCNTMWCNTs/TPU composite385 kPa−13 V64 × 64[97]
PressureDNTTP(VDF:TrFE)0.12 nC N−13 V12 × 12[98]
StrainGrapheneP(VDF:TrFE)Gauge factor (GF): 691 V4 × 4[99]
StrainDPP-PDCA-PDMSDPP-PDCA-PDMSGF: 5.75 × 10560 V5 × 5[100]
Strain-AgNWs--4 × 9[102]
StrainMoS2GrapheneGF: 4125 V10 × 10[103]
TemperatureCNTpolyaniline nanofiber1.0% °C −110 V5 × 5[104]
Temperaturea-IGZOMxene-4 V16 × 16[105]
TemperatureDPPT-TTDPPT-TT9% °C−11 V5 × 5[106]
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

Kim, S.; Yoo, H. Active-Matrix Array Based on Thin-Film Transistors Using Emerging Materials for Application: From Lab to Industry. Electronics 2024, 13, 241. https://doi.org/10.3390/electronics13010241

AMA Style

Kim S, Yoo H. Active-Matrix Array Based on Thin-Film Transistors Using Emerging Materials for Application: From Lab to Industry. Electronics. 2024; 13(1):241. https://doi.org/10.3390/electronics13010241

Chicago/Turabian Style

Kim, Seongjae, and Hocheon Yoo. 2024. "Active-Matrix Array Based on Thin-Film Transistors Using Emerging Materials for Application: From Lab to Industry" Electronics 13, no. 1: 241. https://doi.org/10.3390/electronics13010241

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