Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 2241 2022-07-20 10:16:09 |
2 Reference format revised. Meta information modification 2241 2022-07-21 04:15:40 | |
3 added a period Meta information modification 2241 2022-07-21 04:16:47 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Jeong, C.;  Anwer, A.;  Khan, N.;  Ansari, M.Z.;  Baek, S.;  Yi, H.;  Kim, S.;  Noh, S.M. Touch Sensors for Flexible Wearable Devices. Encyclopedia. Available online: https://encyclopedia.pub/entry/25321 (accessed on 05 July 2024).
Jeong C,  Anwer A,  Khan N,  Ansari MZ,  Baek S,  Yi H, et al. Touch Sensors for Flexible Wearable Devices. Encyclopedia. Available at: https://encyclopedia.pub/entry/25321. Accessed July 05, 2024.
Jeong, Changyoon, Abdul Anwer, Nishat Khan, Mohd Zahid Ansari, Sang-Soo Baek, Hoon Yi, Soeun Kim, Seung Man Noh. "Touch Sensors for Flexible Wearable Devices" Encyclopedia, https://encyclopedia.pub/entry/25321 (accessed July 05, 2024).
Jeong, C.,  Anwer, A.,  Khan, N.,  Ansari, M.Z.,  Baek, S.,  Yi, H.,  Kim, S., & Noh, S.M. (2022, July 20). Touch Sensors for Flexible Wearable Devices. In Encyclopedia. https://encyclopedia.pub/entry/25321
Jeong, Changyoon, et al. "Touch Sensors for Flexible Wearable Devices." Encyclopedia. Web. 20 July, 2022.
Touch Sensors for Flexible Wearable Devices
Edit

The traditional touch sensors are bulky, complicated, inflexible, and difficult-to-wear devices made of stiff materials. The touch screen is gaining further importance with the trend of current IoT technology flexibly and comfortably used on the skin or clothing to affect different aspects of human life. Flexible sensors have advantages over conventional rigid devices. These sensors are nontoxic and lightweight and can be worn comfortably on the body.

flexible sensor wearable devices touch sensor

1. Touch-Sensor Working Principles

There are three types of sensing actions: touch, movement, and pressure. The distance over which actions are performed differentiates fingertip touch and movement, and fingertip pressure level provides a third sensing action. Regular touch occurs when the finger is at a zero distance from the sensors and no force is applied. With increasing pressure, the amount of force increases (pressure sensing). Sensors that detect movement can use a combination of multiple sensing points or proximity sensors (the current distance exceeds zero).

1.1. Single Mechanism

1.1.1. Capacitive Touch Sensors

Capacitive sensor technologies can be broadly divided into two main categories: mutual-capacitive and self-capacitive technologies [1]. For a self-capacitive system, capacitive changes are measured relative to earth ground. It operates on the basis of the parallel-layer model, where the electrode and bottom (or user’s finger) act as the two layers of the capacitor. As capacitance is added with each “touch”, the capacitance on the electrode of the self-capacitive system increases [2]. Contrarily, any intended or unintended capacitance between two “charge-holding objects” can be termed mutual capacitance [3]. A mutual capacitance between intersecting elements of columns and rows can be intentionally created via projected capacitance touch sensors. With the help of such system electronics, a single scan can detect multiple touches by measuring each node (intersection) individually. The surface capacitance measurement is used by some touch devices to determine the human body (finger). The sensor is contacted by a finger, forming a capacitor dynamically [4].

1.1.2. Resistive Touch Sensors

Resistive touch sensors respond to pressure [5]. These sensors are composed of several layers, with the most critical layer formed by two electrodes separated by another layer with poor conduction. Herein, a light pressure resulting from contact can predominantly change the resistance. Apart from a finger, other stimuli, such as a stylus, can also be sensed by these sensors. Resistive touch sensors were the earliest used sensors, with the advantages of low cost, fast response, a linear output, and high durability. However, such sensors require power, making them unsuitable for wearable systems (i.e., low-power systems) [6].

1.1.3. Piezoelectric and Triboelectric Touch Sensors

Piezoelectric materials translate deformation into electrical energy, enabling the development of a piezoelectric sensor [7]. Applying slight mechanical pressure (light touch) to a so-called piezoelectric material results in the separation of the electric charges because of electrical dipole moments, generating electrical voltage. When the external force is withdrawn, the polarization phenomenon disappears. By detecting the change of the electrical signal, it can be used to realize pressure sensing. Pressure sensing can be achieved by detecting the changing electrical signal. Some of the commonly used piezoelectric materials are BaTiO3, PZT, ZnO, PVDF, PbTiO3, and polypropylene (PP) [8]. Piezoelectric sensor performance may hinder as a result of cracking due to the pressure applied to traditional piezoelectric materials to overcome the problem of cracks and hole defects.

1.1.4. Triboelectric Touch Sensors

Triboelectric sensors (or contact-electrification) is employed by a triboelectric nanogenerator (TENG) [9][10], where a physical contact (a touch) generates a potential electrical signal without reliance on an external power supply.
Triboelectric nanogenerators (TENGs) form the basis of triboelectric sensors. The inductive charges are generated on the friction layers of TENGs under external forces owing to the electron affinity differences. Upon the removal of the external force, current starts to flow between electrodes as a result of the formation of internal potential. Moreover, the current signals are analyzed for the sensing function. Sensors using electrostatic induction for the detection of human motions have been available since the introduction of triboelectric nanogenerators in 2012 [11]. The triboelectric nanogenerators may transform the received signals of body movements into electrical impulses, allowing the sensor to function without the need for external power. Commonly used as friction layer materials, they can easily lose or gain electrons, such as PTFE, PDMS, PI, nylon [12] copper, and silver.

2. Hybrid Multisensor Mechanisms

2.1. Two-Principle Integration Mechanism

A wearable sensor based on a single-principle mechanism is unable to detect multiple features and could not meet the requirements for multifunctional detection. Therefore, it is necessary to develop hybrid mechanisms to detect multiple properties and enable the fabrication of multifunctional sensors.
Two-principle mechanisms consist of various integrated mechanisms, such as triboelectric–piezoelectric mechanisms or capacitive–piezoelectric mechanisms. These two-principle mechanisms are used extensively in wearable devices. Tang et al. demonstrated a working mechanism based on simultaneous triboelectric and piezoelectric mechanisms in self-generated powered sensors [13]. The two-principle mechanisms operate in different directions because the triboelectric components consist of four working electrodes that operate in the xy plane, whereas the piezoelectric element functions in the z plane. This integration has shown extensive applicability in the human–machine interaction and other wearable electronic devices [13].

2.2. Three-Principle Integration

Hybridization involving three integration mechanisms will play a tremendous role in future wearable devices for use in wide applications, such as the human–machine interface, smart robots, and wearable electronics. To further improve the performance of sensing materials, it is important to introduce different mechanistic parameters into sensing mechanisms to expand their applicability in wearable and other devices. Various polymeric materials, such as PVDF, PDMS, PET, polyurethane (PU), polypropylene (PP), and other nanomaterials, are required to improve the cracking performance of materials subjected to great pressure or temperature changes, thereby requiring the integration of more than one mechanism for enhancing the sensing ability of the developed materials. Zhao et al. [14], for example, developed high-performance sensors by combining three effects: triboelectricity, piezoresistivity, and piezoelectricity. They fabricated a nanofilm of carbonized polyacrylonitrile/barium titanate (PAN-C/BTO) using the electrospinning method. The resultant nanofilms could independently and simultaneously detect pressure and curvature with high and enhanced sensitivity. These sensors could be used multifariously in detecting human motion, such as finger tapping and swallowing.

3. Nanocomposite Material for Flexible Touch Sensors

Flexible sensors should be made of a lightweight material that is biocompatible, comfortable, and does not cause irritation. Nanocomposite materials mostly include metallic thin films [15][16], carbon nanotubes (CNTs) [17][18], metal nanowires (NWs) [19], metal nanoparticles (NPs) [20], and conductive polymers [21][22]. Good electrochemical activity, high electrical conductivity, and a large active area make the NWs, CNTs, and conductive polymers preferred choices for sensors. In addition, NW/CNT composites can be coated/printed directly onto a substrate [23] to produce a sensor with high sensitivity, stretchability, and durability. Conductive polymers, such as poly(3,4-ethylene dioxythiophene) (PEDOT), and especially its complex with poly(styrene sulfonate) (PEDOT:PSS), can be synthesized by chemical or electrochemical deposition [24] and exhibit high conductivity, good transmission of light, good processability in water, and high flexibility. For the mass production of sensors, printable conductive materials are particularly advantageous because they enable nearly all sensor parts to be printed. Several metal conductive inks have been shown to be promising because of their tendency to disperse in solvents and their compatibility with different printing technologies. Ag nanoparticle-based inks and nanowires have been extensively studied as flexible electrodes or conductors for polyethylene (PE) [24]. However, Cu nanoparticle-based ink, which is inexpensive and highly conductive, has attracted particular attention [25]. Inks based on carbon nanomaterials (e.g., CNTs and graphene) have also been shown to be printable and stretchable for flexible sensors [26][27]. The integration of Ag nanowires (AgNWs) [28], ITO [29], graphene [30], PDMS [31], CNTs [31], and PEDOT:PSS [32] has led to transparent stretchable electrodes suitable for certain touch displays and photovoltaic applications. The fabrication of transparent electrodes/conductors faces a major challenge in the trade-off relationship among flexibility, transparency, and conductivity, which depends on the conductive filler concentration [33]. This challenge is particularly relevant to touch sensors such as touch screens and fingerprint sensors.
Liquid metals and liquid ionic materials are intrinsically flexible. Ionic additives can improve the conductivity and stretchability of PEDOT:PSS, leading to conductivity as high as 4100 S cm−1 at 100% strain, as demonstrated by Wang et al. [21], and enabling the fabrication of soft sensors that can detect both positive and negative pressures from −60 to 20 kPa [34]. Liquid metals—specifically, eutectic gallium–indium alloys—can be used to fabricate flexible circuits via the integration of room-temperature liquid metals (RTLMs) and water-soluble poly(vinyl alcohol) (PVA) because of their high conductivity, intrinsic stretchability, and low piezoresistivity. Thus, RTLMs and PVA can be integrated into flexible circuits [35], asymmetric force sensors can be prepared from hydrophilic polymer networks [36], and soft sensors can be produced by a 3D-printed rigid micro-bump [37][38]. In dynamic applications, liquid components introduce reliability problems. An excellent solution would be to embed the liquid metals into elastomers, such as PVA.
Furthermore, a polymer matrix combined with a wide range of nanofillers can produce various polymer nanocomposites. Nanofillers can have either two- or three-dimensional structures, similar to CNTs or graphene. The properties of CNTs, such as their high mechanical strength, high aspect ratio, and outstanding electrical properties, make them a particularly promising material. CNTs can be embedded in various polymer matrices to produce materials with diverse properties. Unfortunately, the high Van der Waals forces of CNTs cause the particles to tend to cluster or agglomerations. Thus, the distribution of CNTs in polymers strongly affects the performance of nanocomposites. There have been many efforts to achieve a homogeneously dispersed CNTs within the polymer matrix. This is required to fabrication sensors and establish the foundation for high repeatability. Numerous approaches have been developed to tailor and improve the dispersion of CNTs/polymer composites via solution mixing, including direct mixing, in situ polymerization and melt processing.

4. Manufacturing Technologies

In current studies, spinning, coating, printing, and transferring are indispensable methods that strongly influence the electromechanical and flexible properties of touch sensors because they govern the material and the manufacturing process used.
Yarns or surfaces can be coated with conductive coatings. In this context, graphene nanoplatelets and carbon nanofibers painted along both sides of a rubber piece result in flexible and stretchable capacitive touch sensors with a low sheet resistance (~10 Ω sq−1). Chen et al. [39] developed a flexible touch sensor using electron-induced perpendicular graphene sheets implanted on porous carbon films (Figure 5a). These sheets displayed high sensitivity (0.13 kPa−1 at pressures less than 0.1 kPa and 4.41 MPa−1 at pressures greater than 10 kPa) as well as a fast response time of 66 ms when the substrate was ~0.5 mm Si.
Conductive printing techniques enable the fabrication of flexible touch sensors using printing technology. A transparent capacitive touch sensor fabricated using the inkjet printing technique has been proposed for use in freestanding nanoparticle arrays made of ultrafine polydopamine with a controlled line-to-line separation (similar to pitch) 5b [40]. A self-powered touch sensor with a polyethylene terephthalate substrate with an area of 10 × 10 mm2 and a thickness of 380 µm was found to be capable of powering diodes, supplying power to a device, or charging a capacitor. It is a paper-based harvester screen-printed using a mesh [41]. Recently, researchers have focused extensively on 3D-printing technology because of its high commercialization potential, ease of integration, and large-scale manufacturing capability.

5. Applications of the Touch Types

5.1. E-Skins-Based Approach

An electronic skin is an electronic device that mimics human skin properties by being flexible, stretchable, and self-healing. E-skins have multiple applications, including prosthetics, robotics, and skin-attachable devices. E-skins made of ultrathin materials, such as PET or poly(ethylene naphthalate), are a good choice for a small, sustainable deformations. In addition, elastomer substrates, such as polyurethane (PU), latex, and poly(polydimethylsiloxane) (PDMS) may be used in applications that require stretchability [42]. Asghar et al. [43] used magnetically generated microstructures (MPs/PDMSs) to show a piezo-capacitive flexible pressure sensor. In the device, pressure can be sensed over a wide range (0 to 145 kPa) with a fast response (50 ms), as well as high cyclic stability (>9000 cycles).

5.2. E-Textiles-Based Approach

As one of many wearable sensors, e-textiles in stretchy fabrics have drawn attention from both the academic and industrial worlds because of their supreme wearability in allowing seamless fitting across different body sizes and shapes [44]. Traditionally, e-textiles have been created by knitting, weaving, or embroidering functional fibers into fabrics in a twisted or coaxial structure, or embedding functional nanoparticles directly into fabrics. Through adding electronic elements, e-textiles, such as conductive fibers or fabrics, can be used in wearable devices, the human–machine interface, or for controlling/monitoring applications [45].

5.3. E-Healthcare-Based Approach

For healthcare implementation devices must be low energy consumers and biocompatible to prevent skin irritation. In e-healthcare, wearable bio-chemical/chemical sensors are featured as examples of a flexible touch-sensing. Their greatest issue is maintaining precision in a variable working environment with multiple impacts, such as temperature and humidity [46]. Normally, these biochemical or chemical sensors are often connected to human skin or are woven into garments and fabrics in order to detect external toxins or monitor the level of specific (blood) compounds 7/7 [47].
With the addition of an electronic unit, e-textiles, such as conductive nanofibers or nanofabrics, can be used in wearable devices, data collection, control, and for monitoring applications. A temperature sensor e-textile made up of rGO flakes that have been applied to a bleached cotton yarn used batch dyeing in a high-throughput manner (1000 kg/h) [48]. In addition to the overcoat, cotton yarns were knitted into scaffold shapes via an automatic knitting machine to ensure high mechanical stability against a cyclic analysis at 25–55 °C. The great sensitivity and rapid response time of carbon-based nanomaterials, such as CNTs, have also made them an attractive alternative to rGO flakes for temperature sensors [49].

References

  1. Park, B.J.; Oh, S.; Kim, F.S.; Chang, S.T. Pixel-free capacitive touch sensor using a single-layer ion gel. J. Mater. Chem. C 2019, 7, 10264–10272.
  2. Jeon, G.J.; Lee, S.H.; Lee, S.H.; Shim, J.B.; Ra, J.H.; Park, K.W.; Yeom, H.I.; Nam, Y.; Kwon, O.K.; Park, S.H.K. Highly Sensitive Active-Matrix Driven Self-Capacitive Fingerprint Sensor based on Oxide Thin Film Transistor. Sci. Rep. 2019, 9, 3216.
  3. Heo, S.; Park, K.; Choi, E.H.; Bien, F. Differential Coded Multiple Signaling Method with Fully Differential Receiver for Mutual Capacitive Fingerprint TSP. IEEE Trans. Circuits Syst. I Regul. Pap. 2020, 67, 74–85.
  4. Kweon, O.Y.; Samanta, S.K.; Won, Y.; Yoo, J.H.; Oh, J.H. Stretchable and Self-Healable Conductive Hydrogels for Wearable Multimodal Touch Sensors with Thermoresponsive Behavior. ACS Appl. Mater. Interfaces 2019, 11, 26134–26143.
  5. Chen, W.; Yan, X. Progress in achieving high-performance piezoresistive and capacitive flexible pressure sensors: A review. J. Mater. Sci. Technol. 2020, 43, 175–188.
  6. Chang, H.; Kim, S.; Jin, S.; Lee, S.W.; Yang, G.T.; Lee, K.Y.; Yi, H. Ultrasensitive and Highly Stable Resistive Pressure Sensors with Biomaterial-Incorporated Interfacial Layers for Wearable Health-Monitoring and Human-Machine Interfaces. ACS Appl. Mater. Interfaces 2018, 10, 1067–1076.
  7. Gao, X.; Zheng, M.; Yan, X.; Fu, J.; Zhu, M.; Hou, Y. The alignment of BCZT particles in PDMS boosts the sensitivity and cycling reliability of a flexible piezoelectric touch sensor. J. Mater. Chem. C 2019, 7, 961–967.
  8. Xin, Y.; Liu, T.; Sun, H.; Xu, Y.; Zhu, J.; Qian, C.; Lin, T. Recent progress on the wearable devices based on piezoelectric sensors. Ferroelectrics 2018, 531, 102–113.
  9. Gogurla, N.; Roy, B.; Park, J.Y.; Kim, S. Skin-contact actuated single-electrode protein triboelectric nanogenerator and strain sensor for biomechanical energy harvesting and motion sensing. Nano Energy 2019, 62, 674–681.
  10. Jeong, C.; Lee, S.; Roh, H.D.; Feng, M.Q.; Park, Y. Bin Hierarchically structured ZnO nanorod-carbon fiber composites as ultrathin, flexible, highly sensitive triboelectric sensors. Smart Mater. Struct. 2020, 29, 025002.
  11. Wang, Z.L.; Wu, W. Nanotechnology-enabled energy harvesting for self-powered micro-/nanosystems. Angew. Chemie-Int. Ed. 2012, 51, 11700–11721.
  12. Qian, J.; He, J.; Qian, S.; Zhang, J.; Niu, X.; Fan, X.; Wang, C.; Hou, X.; Mu, J.; Geng, W.; et al. A Nonmetallic Stretchable Nylon-Modified High Performance Triboelectric Nanogenerator for Energy Harvesting. Adv. Funct. Mater. 2020, 30, 1907414.
  13. Tang, G.; Shi, Q.; Zhang, Z.; He, T.; Sun, Z.; Lee, C. Hybridized wearable patch as a multi-parameter and multi-functional human-machine interface. Nano Energy 2021, 81, 105582.
  14. Zhao, G.; Zhang, X.; Cui, X.; Wang, S.; Liu, Z.; Deng, L.; Qi, A.; Qiao, X.; Li, L.; Pan, C.; et al. Piezoelectric Polyacrylonitrile Nanofiber Film-Based Dual-Function Self-Powered Flexible Sensor. ACS Appl. Mater. Interfaces 2018, 10, 15855–15863.
  15. Ansari, M.Z.; Janicek, P.; Nandi, D.K.; Slang, S.; Bouska, M.; Oh, H.; Shong, B.; Kim, S.H. Low-temperature growth of crystalline Tin(II) monosulfide thin films by atomic layer deposition using a liquid divalent tin precursor. Appl. Surf. Sci. 2021, 565, 150152.
  16. Kim, K.J.; Lu, P.; Culp, J.T.; Ohodnicki, P.R. Metal-Organic Framework Thin Film Coated Optical Fiber Sensors: A Novel Waveguide-Based Chemical Sensing Platform. ACS Sensors 2018, 3, 386–394.
  17. Yu, S.; Wang, X.; Xiang, H.; Zhu, L.; Tebyetekerwa, M.; Zhu, M. Superior piezoresistive strain sensing behaviors of carbon nanotubes in one-dimensional polymer fiber structure. Carbon N. Y. 2018, 140, 1–9.
  18. Jeong, C.; Park, Y. Bin Exfoliated Graphite Nanoplatelet-Carbon Nanotube Hybrid Composites for Compression Sensing. ACS Omega 2020, 5, 2630–2639.
  19. Duan, S.; Wang, Z.; Zhang, L.; Liu, J.; Li, C. A Highly Stretchable, Sensitive, and Transparent Strain Sensor Based on Binary Hybrid Network Consisting of Hierarchical Multiscale Metal Nanowires. Adv. Mater. Technol. 2018, 3, 1800020.
  20. Zhao, Y.; Yang, Y.; Cui, L.; Zheng, F.; Song, Q. Electroactive nanoparticles driven electrochemical sensor for endogenous H2S detection. Biosens. Bioelectron. 2018, 117, 53–59.
  21. Wang, Y.; Zhu, C.; Pfattner, R.; Yan, H.; Jin, L.; Chen, S.; Molina-Lopez, F.; Lissel, F.; Liu, J.; Rabiah, N.I.; et al. A highly stretchable, transparent, and conductive polymer. Sci. Adv. 2017, 3, e1602076.
  22. Wang, Y.; Ding, Y.; Guo, X.; Yu, G. Conductive polymers for stretchable supercapacitors. Nano Res. 2019, 12, 1978–1987.
  23. Min, S.H.; Lee, G.Y.; Ahn, S.H. Direct printing of highly sensitive, stretchable, and durable strain sensor based on silver nanoparticles/multi-walled carbon nanotubes composites. Compos. Part B Eng. 2019, 161, 395–401.
  24. Chu, T.Y.; Zhang, Z.; Tao, Y. Printing Silver Conductive Inks with High Resolution and High Aspect Ratio. Adv. Mater. Technol. 2018, 3, 1700321.
  25. Zhong, Z.; Lee, H.; Kang, D.; Kwon, S.; Choi, Y.M.; Kim, I.; Kim, K.Y.; Lee, Y.; Woo, K.; Moon, J. Continuous Patterning of Copper Nanowire-Based Transparent Conducting Electrodes for Use in Flexible Electronic Applications. ACS Nano 2016, 10, 7847–7854.
  26. Liang, X.; Li, H.; Dou, J.; Wang, Q.; He, W.; Wang, C.; Li, D.; Lin, J.M.; Zhang, Y. Stable and Biocompatible Carbon Nanotube Ink Mediated by Silk Protein for Printed Electronics. Adv. Mater. 2020, 32, e2000165.
  27. Bellani, S.; Petroni, E.; Del Rio Castillo, A.E.; Curreli, N.; Martín-García, B.; Oropesa-Nuñez, R.; Prato, M.; Bonaccorso, F. Scalable Production of Graphene Inks via Wet-Jet Milling Exfoliation for Screen-Printed Micro-Supercapacitors. Adv. Funct. Mater. 2019, 29, 1807659.
  28. An, B.W.; Heo, S.; Ji, S.; Bien, F.; Park, J.U. Transparent and flexible fingerprint sensor array with multiplexed detection of tactile pressure and skin temperature. Nat. Commun. 2018, 9, 2458.
  29. Yuan, Z.; Zhou, T.; Yin, Y.; Cao, R.; Li, C.; Wang, Z.L. Transparent and Flexible Triboelectric Sensing Array for Touch Security Applications. ACS Nano 2017, 11, 8364–8369.
  30. Liu, N.; Chortos, A.; Lei, T.; Jin, L.; Kim, T.R.; Bae, W.G.; Zhu, C.; Wang, S.; Pfattner, R.; Chen, X.; et al. Ultratransparent and stretchable graphene electrodes. Sci. Adv. 2017, 3, e170015.
  31. Chen, J.; Zhu, Y.; Jiang, W. A stretchable and transparent strain sensor based on sandwich-like PDMS/CNTs/PDMS composite containing an ultrathin conductive CNT layer. Compos. Sci. Technol. 2020, 186, 107938.
  32. Teo, M.Y.; Kim, N.; Kee, S.; Kim, B.S.; Kim, G.; Hong, S.; Jung, S.; Lee, K. Highly stretchable and highly conductive PEDOT:PSS/Ionic liquid composite transparent electrodes for solution-processed stretchable electronics. ACS Appl. Mater. Interfaces 2017, 9, 819–826.
  33. Li, W.; Yang, S.; Shamim, A. Screen printing of silver nanowires: Balancing conductivity with transparency while maintaining flexibility and stretchability. npj Flex. Electron. 2019, 3, 13.
  34. Shi, H.; Al-Rubaiai, M.; Holbrook, C.M.; Miao, J.; Pinto, T.; Wang, C.; Tan, X. Screen-Printed Soft Capacitive Sensors for Spatial Mapping of Both Positive and Negative Pressures. Adv. Funct. Mater. 2019, 29, 1809116.
  35. Teng, L.; Ye, S.; Handschuh-Wang, S.; Zhou, X.; Gan, T.; Zhou, X. Liquid Metal-Based Transient Circuits for Flexible and Recyclable Electronics. Adv. Funct. Mater. 2019, 29, 1808739.
  36. Peng, H.; Xin, Y.; Xu, J.; Liu, H.; Zhang, J. Ultra-stretchable hydrogels with reactive liquid metals as asymmetric force-sensors. Mater. Horizons 2019, 6, 618–625.
  37. Kim, K.; Choi, J.; Jeong, Y.; Cho, I.; Kim, M.; Kim, S.; Oh, Y.; Park, I. Highly Sensitive and Wearable Liquid Metal-Based Pressure Sensor for Health Monitoring Applications: Integration of a 3D-Printed Microbump Array with the Microchannel. Adv. Healthc. Mater. 2019, 8, e1900978.
  38. Ansari, M.Z.; Parveen, N.; Nandi, D.K.; Ramesh, R.; Ansari, S.A.; Cheon, T.; Kim, S.H. Enhanced activity of highly conformal and layered tin sulfide (SnSx) prepared by atomic layer deposition (ALD) on 3D metal scaffold towards high performance supercapacitor electrode. Sci. Rep. 2019, 9, 10225.
  39. Chen, S.; Wang, Y.; Yang, L.; Karouta, F.; Sun, K. Electron-Induced Perpendicular Graphene Sheets Embedded Porous Carbon Film for Flexible Touch Sensors. Nano-Micro Lett. 2020, 12, 136.
  40. Liu, L.; Pei, Y.; Ma, S.; Sun, X.; Singler, T.J. Inkjet Printing Controllable Polydopamine Nanoparticle Line Array for Transparent and Flexible Touch-Sensing Application. Adv. Eng. Mater. 2020, 22, 1901351.
  41. Ferreira, G.; Goswami, S.; Nandy, S.; Pereira, L.; Martins, R.; Fortunato, E. Touch-Interactive Flexible Sustainable Energy Harvester and Self-Powered Smart Card. Adv. Funct. Mater. 2020, 30, 1908994.
  42. Yang, J.C.; Mun, J.; Kwon, S.Y.; Park, S.; Bao, Z.; Park, S. Electronic Skin: Recent Progress and Future Prospects for Skin-Attachable Devices for Health Monitoring, Robotics, and Prosthetics. Adv. Mater. 2019, 31, 1904765.
  43. Asghar, W.; Li, F.; Zhou, Y.; Wu, Y.; Yu, Z.; Li, S.; Tang, D.; Han, X.; Shang, J.; Liu, Y.; et al. Piezocapacitive Flexible E-Skin Pressure Sensors Having Magnetically Grown Microstructures. Adv. Mater. Technol. 2020, 5, 1900934.
  44. Heo, J.S.; Eom, J.; Kim, Y.H.; Park, S.K. Recent Progress of Textile-Based Wearable Electronics: A Comprehensive Review of Materials, Devices, and Applications. Small 2018, 14, 1703034.
  45. Ismar, E.; Kurşun Bahadir, S.; Kalaoglu, F.; Koncar, V. Futuristic Clothes: Electronic Textiles and Wearable Technologies. Glob. Challenges 2020, 4, 1900092.
  46. Tessarolo, M.; Gualandi, I.; Fraboni, B. Recent progress in wearable fully textile chemical sensors. Adv. Mater. Technol. 2018, 3, 1700310.
  47. Kamarudin, S.F.; Mustapha, M.; Kim, J.K. Green Strategies to Printed Sensors for Healthcare Applications. Polym. Rev. 2021, 61, 116–156.
  48. Afroj, S.; Karim, N.; Wang, Z.; Tan, S.; He, P.; Holwill, M.; Ghazaryan, D.; Fernando, A.; Novoselov, K.S. Engineering Graphene Flakes for Wearable Textile Sensors via Highly Scalable and Ultrafast Yarn Dyeing Technique. ACS Nano 2019, 13, 3847–3857.
  49. Zaporotskova, I.V.; Boroznina, N.P.; Parkhomenko, Y.N.; Kozhitov, L.V. Carbon nanotubes: Sensor properties. A review. Mod. Electron. Mater. 2016, 2, 95–105.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , , , , , ,
View Times: 1.7K
Revisions: 3 times (View History)
Update Date: 21 Jul 2022
1000/1000
Video Production Service