Next Article in Journal
A Heartbeat Classifier for Continuous Prediction Using a Wearable Device
Previous Article in Journal
Underwater Acoustic Signal Detection Using Calibrated Hidden Markov Model with Multiple Measurements
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Progress of Research into Flexible Sensors in the Field of Smart Wearables

1
College of Textile, Zhongyuan University of Technology, Zhengzhou 450007, China
2
Henan Province Collaborative Innovation Center of Textile and Garment Industry, Zhengzhou 450007, China
3
College of Automation Engineering, University of Electronic Science and Technology, Chengdu 611731, China
*
Author to whom correspondence should be addressed.
Sensors 2022, 22(14), 5089; https://doi.org/10.3390/s22145089
Submission received: 20 May 2022 / Revised: 2 July 2022 / Accepted: 3 July 2022 / Published: 6 July 2022
(This article belongs to the Section Electronic Sensors)

Abstract

:
In modern society, technology associated with smart sensors made from flexible materials is rapidly evolving. As a core component in the field of wearable smart devices (or ‘smart wearables’), flexible sensors have the advantages of excellent flexibility, ductility, free folding properties, and more. When choosing materials for the development of sensors, reduced weight, elasticity, and wearer’s convenience are considered as advantages, and are suitable for electronic skin, monitoring of health-related issues, biomedicine, human–computer interactions, and other fields of biotechnology. The idea behind wearable sensory devices is to enable their easy integration into everyday life. This review discusses the concepts of sensory mechanism, detected object, and contact form of flexible sensors, and expounds the preparation materials and their applicability. This is with the purpose of providing a reference for the further development of flexible sensors suitable for wearable devices.

1. Introduction

With the development of technology, intelligent wearable devices that can be integrated into the human body and perform multiple functions have come from science fiction to people’s real life. With the increasing demand in the field of smart wearables, ordinary sensors have gradually failed to meet the requirements. Flexible sensors with the characteristics of flexibility, foldability, and wearability have become a vital breakthrough [1]. Using flexible sensors, researchers can convert the external force into electrical signals, perform signal processing, and use flexible sensors in wearable products to monitor human body indicators in real-time and accurately [2,3,4,5].
Flexible sensors and traditional sensors are produced with entirely different materials and processes. Flexible sensors must be applied to flexible fabrics, plastic films, etc. They cannot use high-temperature processes, have high requirements for electrical conductivity, and their performance must remain stable under folding, bending, and water vapor infiltration. The properties of flexible sensors vary with the preparation method, and the working mechanism is also different from that of ordinary sensors, including changing the conductive network between overlapping nanomaterials, the tunneling effect, crack propagation, and more [6,7,8]. Due to the various preparation methods of conductive materials, they have greater compatibility with different substrate materials. For example, a wide range of processes, such as dip coating, in situ assembly, mechanical press, transfer printing technology, and chemical deposition, etc., can be used to combine with the substrate [9,10,11,12]. Cyclic stretching or compression is usually used to test the selection of a suitable polymer matrix to design a special conductive network to stabilize the conductive network, so that the sensitive components have a stable sensing behavior. In addition, by means of their bionic function and structure [13], such as artificial nanostructures derived from highly sensitive structures (insect cracks or leaves, etc.) and stretchable structures (wrinkles, textured structures, etc.) that occur in nature, it provides additional feasible strategies for generating good sensitivity and stretchability.
In addition to their continuous mapping capabilities, flexible sensors are characterized by high sensitivity, flexibility, and simple manufacturing processes [14]. However, flexible sensors currently face significant challenges as they are still in the initial stages of development. Flexible sensors must be integrated into object surfaces with compatibility, durability, and wear resistance. In contrast, existing commercial smart wearable devices are mainly implemented by packaging integrated circuits with solid substrates. However, solid substrates are mechanically incompatible with the soft human body, resulting in unreliable measurement results due to measurement position changes and unreliable skin contact.
This paper develops the discussion of flexible sensors in the field of smart wearable from the research history of flexible sensors. Then, the different classifications of flexible sensors in terms of the sensing mechanism, detection mode, and contact form are elaborated. Next, six common materials for flexible sensors are introduced. Next, the frontier applications of flexible sensors in the field of smart wearables are also discussed. Finally, the current challenges and future opportunities of flexible sensors are summarized to provide a reference for the development of wearable, flexible sensors.

Research History of Flexible Sensors

Since the late 1990s, smart wearables have been developed in various ways. More user-friendly and human-centered designs have been developed, such as diverse research on physical factors such as comfort, usability, ergonomics, and wearer design, and smart wearables have been developed in all aspects and are widely used in areas such as electronic skin, health monitoring, sports monitoring, smart medicine, and human–computer interaction. In 2000, Jiang et al. [15] proposed a new microfabrication technique that has been developed and applied to the development of flexible shear–stress sensor arrays. In 2003, Kerpa et al. [16] proposed the development of a flexible tactile sensor system for humanoid robots. In 2006, Manunza et al. [17] fabricated the first fully flexible field effect device for chemical detection based on organic field-effect transistors, which opened the way for the production of flexible chemical and strain sensors. A wearable yarn-based piezoresistive sensor was put forward by Huang et al. in 2008 [18], demonstrating the feasibility of a yarn-based sensor, and the results showed that the yarn-based sensor could accurately track breathing signals. In 2010, Figueiredo et al. [19] developed a low-power wireless acquisition module that does not require the preparation of electrolyte gels, and it provides continuous monitoring of the user’s electrocardiogram (ECG) and activity. The design of textile touch sensing interaction was explored by Roh et al. [20] in 2013 with the new metal composite embroidery yarns (MCEYs) and a simple and easy fabrication technique. In 2016, Ji et al. [21] presented a flexible capacitive tactile sensor for robot skin, and a tactile feedback and signal acquisition system was built for the application of robotic obstacle avoidance. In 2017, Li et al. first fabricated [22] a highly flexible multifunctional smart coating using spray-coating multiwalled carbon nanotubes dispersed in a thermoplastic elastomer solution, followed by treatment with ethanol. In 2019, Wang et al. [23] comprehensively summarized a comprehensive review of the latest progress concerning smart wearable sensors, with a focus on bio-multifunctional (biocompatible, biodegradable, and self-healing) device designs. In 2020, Baeg et al. [24] summarized recent progresses in flexible electronic systems, composed of transistor-based circuitry and active matrices built on plastic or textile substrates. In 2021, Guo et al. [25] developed the design and production of dual microstructures of surface micro-bumps and internal hollow pores in the conductive material, MXene, to obtain a multifunctional high-performance pressure sensor, showing an ultra-sensitive ability to extract gesture behavioral information and physiological information from the sensor signals, and its implications for human health. In 2022, Zhang et al. [26] developed an on-skin ultrathin and stretchable multifunctional sensor for smart medical wearables that paves a promising path for future wearables for smart skin and healthcare applications. The above discussion briefly outlines the timeline of major milestones in the development of flexible sensors in the smart wearable field (Figure 1).

2. Sensory Mechanisms of Flexible Sensors

Flexible sensors are the focus of intelligent wearable research. They are flexible sensors with different sensing mechanisms, different detected objects, and different contact forms (Figure 2). Flexible sensors can convert external activity (such as mechanical deformation) into directly measurable electrical signals. Depending on their sensory mechanism, they can be divided into resistive [27], piezoelectric [28], capacitive [29], and other types of friction [30]. When the external pressure changes, the appropriate resistance value, capacitance value, or voltage value of the flexible sensor changes accordingly. This means sensors can be alternately categorized as physical or chemical sensors as appropriate to the detected object, and further divided into contact and non-contact sensors according to whether physical connection is required. Of these, contact sensors are the traditional type, in which they cause their own deformation through some physical touch and applied mechanical force, which triggers changes to the electrical signal. This is a step in the mode of operation that non-contact sensors can realize with long-distance reading to collect the same information of the measured object, thus achieving the desired function without physical contact.

2.1. Resistive Flexible Sensor

A flexible resistive sensor is a device that converts external pressure values into resistance or current values. In accordance with the different sensory mechanisms of the device, it can be defined as a piezoresistive or strain type. For flexible piezoresistive sensors, when subjected to external pressure, the conductivity of the conductor material is changed with some small changes in itself [23], i.e., the piezoresistive effect. By the same token, when flexible strain sensors are stretched or compressed by external pressure (that is, when mechanical deformation occurs), the cross-sectional area of the conductor changes, resulting in a corresponding change in the conductive area and thus the recorded resistance [37]. Figure 3 is a schematic diagram of the signal conversion mechanism of a resistive flexible sensor.
A piezoresistive flexible sensor is a sensor based on the piezoresistive effect [38]. When subjected to external pressure, the conductivity of the conductor material changes with minor changes. Flexible piezoresistive sensors have the advantages of a high sensitivity, good linearity, and measurable static force, which has stimulated much research in recent years. Moreover, they can provide a convenient, timely, and portable solution for the detection of motion when worn; however, the development of cost-effective pressure sensor materials with high compressibility and sensitivity has proven to be challenging. Chowdhury et al. [39] developed a flexible and stretchable CNF/PDMS nanocomposite with piezoresistive sensing capabilities via the use of ultrasonic fabrication technology. Excellent electrical conductivity and durable and reliable piezoresistive sensing responses could be obtained by dispersing CNFs into PDMS. The versatility and ease of fabrication of the developed nanocomposites can make up for the lack of high sensitivity and applicability, flaws common to current flexible sensors. This has an undoubted impact on the practical applications of sensors in smart wearables. Yu et al. [40] constructed a highly sensitive, flexible, and wearable flexible piezoresistive sensor based on the structures of pine and needles in nature, in which the sensing layer was composed of graphene-attached tissue fibers and zinc oxide (ZnO). The original 3D structure of the optical fiber was improved by growing ZnO crystals in situ and coupling its piezoelectric capability, and, at the same time, the zinc oxide could be used as a component in the photoelectric sensor. With an ultra-wide testable pressure range (0–100 KPa), the sensor is capable of monitoring subtle pressure changes caused by respiration or the wearer’s pulse (i.e., at the wrist) and large pressure changes that occur during activities such as walking, wrist movements (e.g., from sports), and running. Dynamic or static pressure sensing is achieved thanks to the properties of ZnO and graphene fibers, which also endow wearable sensors with good optical responses. Yang et al. [13], inspired by the surface microstructure of rose petals, developed a layered microstructure as the core of a biomimetic flexible piezoresistive sensor, consisting of a PANI/PVDF nanofiber membrane. This sensor could fit well on the skin, demonstrating it capability for monitoring human physiological signals and motion status, such as wrist pulse, throat activity, spinal posture, and gait recognition. This study provides a promising research scheme for the rapid development of next-generation wearable bioelectronic devices.
As shown by these examples, flexible piezoresistive sensors have attracted great attention in the fields of electronic skin, intelligent robots, and human–computer interaction due to their portability, flexibility, and excellent sensory performance. Li et al. [41] fabricated flexible piezoresistive sensors by treating the backbone of PU with CS to obtain their positively charged equivalents before dip-coating the negatively charged MXene material Ti3C2Tx. The sensors of Ti3C2Tx with CS and PU were observed to provide a versatile sensory platform with the potential to detect large and small pressure signals. Ding et al. [42] also used the dip-coating method to prepare PEDOT:PSS sponges for flexible piezoresistive sensors with good compressibility and a stable piezoresistive response. These were intended for use in the detection of various human actions, including speaking, bending fingers or elbows, and walking. This is indicative of the increasingly important role flexible piezoresistive sensors play in smart wearables, with potential roles in other applications, including virtual reality, entertainment technology, human–machine interface, personal healthcare, and sports science.
Flexible strain sensors can be attached to clothing or human skin for real-time monitoring of human activity. As the core of smart wearable sensing devices, they have generated a great deal of interest among the scientific research community. Among the work carried out in developing them, Zhou et al. [43] developed a sensor with high elasticity and durability by spraying CNT ink onto pre-stretched TPU fiber mats (Figure 4a). This sensor was shown to have a fast response time (70 ms), excellent durability, and good sensory performance when responding to bending. Due to its ultra-high sensitivity, it can easily monitor various subtle movements across the human body. Elsewhere, Yang et al. [44] fabricated a graphene fabric sensor with a negative resistance change by thermal reduction of GO (Figure 4b). Such sensors can be directly knitted onto clothing for monitoring both large and subtle movements. The same team also developed an AgNPs-bridged graphene variant (Figure 4c) for real-time detection of subtle and intense human motion [45]. It demonstrated an excellent range of strain detection with a gage factor (GF) of 475, fast response/recovery speed, and good linearity, showing good durability and long-term stability in stretching/repetitive cycling, compared to the performance of other graphene flexible strain sensors. Larimi et al. [46] fabricated a flexible strain sensor by infusing graphene nanoflakes onto a rubber-like adhesive pad (Figure 4d), resulting in a highly stretchable product capable of withstanding up to 350% strain. Strong and stable electrical responses were maintained even after 10,000 stretching cycles and the sensor showed an ability to measure the human heartbeat and track a wide range of human motion. In other testing, it could also be used in robotic haptic applications to control robotic limbs and extremities.

2.2. Piezoelectric Flexible Sensors

This type of sensor is designed based on the so-called piezoelectric effect principle of appropriate materials [47]. The piezoelectric effect refers to the deformation (including bending and expansion) of some dielectrics under the action of an external force in a particular direction. In this case, electrical polarization occurs internally, and charges of opposite signs are generated on the two surfaces at the same time. When the external force is removed, the dielectric returns to an uncharged state and when the force’s direction changes, the polarity of the charge also interacts with the changes. The amount of charge generated by the dielectric is proportional to the magnitude of the externally applied force. In this process, the magnitude of the external force can be calculated by the magnitude of the current (Figure 5).
Self-powered physical sensors for wearable applications have been in great demand in recent years, and flexible piezoresistive sensors have taken an important step forward as self-powered devices. Such sensors have the characteristics of self-power, high sensitivity, and good responsiveness [48], and are an important part of the physical movement energy in the human body. Chen et al. [49] developed a flexible triaxial tactile sensor with piezoelectrically enhanced P(VDF-TrFE) micropillars, which has great potential in advanced robotics, wearable electronics, and various human–machine interface implementations. To realize a large-area flexible pressure sensor array applicable for smart wearable products, Zhang et al. [50] developed a PVDF-TrFE/AgNWs piezoelectric capacitor and a-IGZOTFT piezoelectric sensor integrated on a PI flexible substrate in sequence. The doped AgNWs enhanced the piezoelectric properties of PVDF-TrFE while avoiding damage to TFTs, which would have been caused by high-voltage polarization. In addition, combined with the flexible TFT amplification circuit, the corresponding signal was further amplified.
Ultrathin sensing devices utilizing piezoelectric materials have become an important part of developing piezoelectric sensors for continuous bio-signal monitoring systems that can conform to the contours of the skin. Kim et al. [51] developed a transparent and biocompatible BNNS piezoelectric sensor using BNNS and PDMS as a piezoelectric active and flexible element, respectively (Figure 6a). These sensors can generate electrical energy with various signal forms for monitoring body movement. Moreover, Wang et al. [52] proposed an energy-saving and fully flexible piezoelectric sensor (Figure 6b), which was possible through the integration of PVDF nanorod arrays with excellent mechanical force–electrical pulse conversion. These are based on piezoelectric effect-polarized PVDF arrays, which generate sufficient charge to drive the device and realize self-signal amplification, meaning the fabricated sensor has a good pressure sensitivity, detection limit, and response time. Yang et al. [53] introduced PDA as a surface modifier to BTO and blended it with a PVDF matrix in different proportions to form homogeneous PDA/BTO/PVDF composites. From this, a flexible piezoelectric pressure sensor was fabricated using the surface solution casting technique (Figure 6c). The sensor was shown to detect joint bending and human motion exhibited by different movement styles in different signal curves. Additionally, Jiang et al. [54] produced PVDF nanocomposites containing BaTiO3 nanoparticles and incorporated them, via electrospinning, into flexible piezoelectric pressure tactile sensors (Figure 6d). These sensors demonstrated a good flexibility and linear response to external mechanical force, and were able to detect different musical sounds.

2.3. Flexible Capacitive Sensors

When a flexible capacitive sensor is subjected to external pressure (the magnitude of which can be calculated by measuring the changes in the electrical signals), the capacitance value between the electrode plates also changes [55], causing changes in other electrical signals. Generally, flexible capacitive sensors are dielectric materials with a microstructure added to the middle of the flexible electrode. This microstructure changes significantly under the action of external pressure, which results in changes in the dielectric properties of the dielectric material and thus the capacitance value as well [56]. The sensitivity of flexible capacitive sensors can generally be improved by substituting the material of the dielectric layer, performing related structural modifications, or by compounding the substrate with a substance with a high dielectric constant [57]. Figure 7 shows the signal conversion mechanism of the capacitive flexible sensor.
With respect to wearable electronic devices, capacitive pressure sensors benefit from low energy consumption, good stability, simple structure, and adjustable sensitivity. Unsurprisingly, such developments have attracted much attention, giving rise to further innovations. Xiong et al. [58] made an ultra-high sensitivity flexible capacitive pressure sensor based on convex microarrays and ultrathin dielectric layers. When external pressure was applied to the sensor, the sensitivity of the sensor was recorded to greatly improve due to the synergistic effect of the microarray on the electrode surface and the ultra-thin dielectric layer, resulting in a significant increase in the contact area and a similar reduction in the distance between the two electrodes. This flexible sensor has been successfully used to monitor various human bio-signals and robotic hand movements, paving the way for applications in areas such as smart medical care, automatic speech recognition, and language recognition. Moreover, Li et al. [59] developed a highly sensitive, stretchable, and tough capacitive strain sensor using double-crosslinked zinc alginate-alginate/PAM hydrogels as ionic conductors to repeatedly adhere a PVA/PAA borax organo-gel, which served as the dielectric layer [60]. By combining the mechanical advantages of water/organo-gels, the assembled capacitive strain sensor exhibited high elasticity and excellent sensitivity. These synergistic effects enabled capacitive strain sensors to be used in highly reliable wearable sensory devices to monitor various weak physiological signals and a wide range of human movements, such as finger, wrist, elbow, and knee joint movements, smiling, and talking. Additionally, Keum et al. [61] developed a fabric-based high-sensitivity sensor using silver-coated multifilament fibers woven into a conductive fabric as electrodes and flexible ion-gel membranes with high dielectric constants as sensing elements. The composition of this capacitive system of this sensor allowed it to achieve a multi-point detection capability on large-area fabrics.

2.4. Flexible Friction/Triboelectric Sensors

Flexible triboelectric sensors are typically composed of upper and lower electrodes, each lined with material on the inner face responsible for triboelectric generation, with a thin layer of air between the two. When pressure is applied, two materials with different charged sequences come into contact with each other, and a triboelectric phenomenon occurs, generating opposite charges on both sides of the interface. When the pressure is released, the two surfaces with equal and opposite charges are automatically separated. Due to the phenomenon of electrostatic induction, compensating charges are proportionately generated on the surface of the electrodes, but the air layer between the materials cannot completely neutralize the charges on the two surfaces, resulting in a potential difference [62]. This mechanism enables flexible triboelectric sensors to generate signals when pressure is applied and released (Figure 8).
Mechanical energy, if harnessed efficiently, can not only make a significant contribution to global electrical demands but also serve as an independent and sustainable energy source for mobile electronic devices. However, triboelectricity is difficult to accumulate and utilize, so its value is often overlooked. In recent years, the high efficiency, low cost, and environmental friendliness of triboelectric nanogenerators have attracted attention among researchers. In one instance, Ning et al. [63] constructed a wearable and self-powered triboelectric sensor by introducing a helical structure braided fiber on a stretchable base. Due to this structure, the sensor exhibited high sensitivity, with a single helical fiber-optic strain sensor (10 cm in length) outputting 0.5 V even at 1% tensile strain. For its practical applications, this helical-structured sensor functions by responding to instances of contraction and relaxation of the thoracic cavity and abdomen caused by the heartbeat and breathing. Based on this sensor, a wearable self-powered real-time respiration monitoring system has been developed. Further, Zhang et al. [64] designed a high-performance flexible self-powered sensor with SnS2 restoring GO-based humidity, driven by a triboelectric nanogenerator. Under the effects of external mechanical motion, the nanogenerator generated a large voltage through the triboelectric effect between PTFE and a copper film. Applying this concept to a flexible humidity sensor, an SnS2/GO composite was fabricated on a PET substrate by means of screen printing. By integrating the two components in this manner, a self-powered triboelectric humidity sensor was realized. The sensor featured a high stable output voltage, fast response/recovery time, long-term stability, and ultra-low power consumption, making it suitable for sensing humidity over a wide relative range. In addition, it was capable of monitoring human breathing, coughing, and the approach of objects to the respiratory tract, enabling multifunctional applications in various fields.
The rapid development of robotics and virtual reality technology has put forward higher requirements with respect to advances in the human–machine interface in terms of achieving efficient parallel control. Relevant to this, Zhu et al. [65] proposed a triboelectric bidirectional (TBD) sensor based on the switch and basic grating structures. It had a bidirectional sensory capability for the detection of rotational and linear motions from different joints, including multi-DOF rotation of the shoulder and twists of the wrist. At the same time, the sensor exhibited a good performance irrespective of the speed (good results from 10 to 300 rpm), proving its feasibility in different scenarios. Elsewhere, Chen et al. [66] proposed and studied a self-powered flexible triboelectric sensor patch applied for trajectory and fingertip motion. It was made from environmentally friendly materials, namely starch-based hydrogel, PDMS, and silicone rubber. By designing a grid structure on top of the triboelectric functional layer, only four inductions were needed to detect the trajectory, velocity, and acceleration of the fingertip movements on the 2D-SFTS, and the one-dimensional stretching displacement and velocity could be easily measured. The combination of the two patches achieved the 3D motion control of the robotic manipulator.
Flexible resistive sensors can detect static force and dynamic force with a simple structure but suffer from substantial signal drift, although it is not easily disturbed by external fields. However, they have the advantages of being self-powered and a fast response time. Conversely, analogous capacitive sensors have a simple structure and small signal drift but are easily disturbed by external fields, and their sensitivity is limited by the compressibility of the dielectric layer. Additionally, the flexible triboelectric sensor benefits from a low production cost, simple structure, high output voltage, and self-driving ability. On the measured object, the general resistive and capacitive sensors have a wider range of applications, albeit with exceptions. In some unfavorable environments, the power supply needs to be replaced after some time, which is more labor and material resource intensive, while flexible piezoelectric and triboelectric sensors can detect signals in the field for a greater period of time, saving in costs, thanks to being self-powered. A comparison of sensors with different sensory mechanisms is shown in Table 1.

3. Detection Methods of Flexible Sensors

3.1. Physical Sensors

Physical sensors are those which use certain conversion element’s physical properties and effects such as dissociation [86], polarization [87], thermoelectricity [88], photoelectricity [89], magnetoelectricity [90], and others. In recent years, these classifications of sensors have been used for signal detection in areas related to smart wearable sensing devices, such as recording body movements, physiological signals, body temperature, and similar such information. External environmental characteristics, for example, ambient temperature and humidity, airflow, and light, are also applicable areas for signal detection.
With the development of advanced materials and fabrication technologies, flexible sensors with multifunctional sensory capabilities are becoming increasingly popular in the field of smart wearables. Furthermore, the combination of elastic substrates with conductive fillers, such as graphene [91], carbon nanotubes [92], metals/semiconductors [93], and conductive polymers [94], has been widely investigated. However, most of these sensors suffer from the disadvantages of a low elasticity, insufficient flexibility, and poor durability due to the rigidity of their materials. This has increased the need for stretchable, flexible, and highly sensitive sensors capable of detecting various mechanically induced deformations. Khalili et al. [95] produced a nanofiber network based on physically cross-linked PLA and TPU blends by electrostatic spinning and depositing a uniform layer of SWCNT on the fibers using a spraying method in order to make the fiber surface conductive. The resulting sensor possessed shape memory properties, wherein the fibers were able to retain their original shape after stretching and thermal recovery. This behavior allows sensors to overcome their plastic deformation and greatly improve their elasticity. Elsewhere, Li et al. [96] developed a multidimensional flexible sensor fabricated from Ag/PDMS composites. The development and production of this type of flexible sensor is simple and fast, and the Ag/PDMS flexible sensor has a variety of applications, including real-time dynamic monitoring of the human body and sound intensity detection. In yet further work, Kumar et al. [97] developed a printable PU nanocomposite ink using PEG as a thermal material, CNT as a heat-resistant conductive filler, and PU as a flexible structural material, using physical mixing and chemical cross-linking methods to develop low-cost, flexible temperature sensors.

3.2. Chemical Sensors

Chemical sensors use electrochemical reactions to convert the composition and concentration of inorganic and organic chemical substances into electrical signals. The most commonly used active components are ion-selective electrodes, which serve in this function by measuring the pH of a solution or the concentration of certain ions, such as K+, Na+, Ca+, etc. [98,99]. Currently, chemical sensors are most commonly used in wearable and flexible sensing technologies, which offer advantages such as high sensitivity, miniaturization of the sensors, and direct measurement without markers [100].
Some traditional methods for preparing electrochemical sensors have been applied to the field of smart wearable or flexible platforms. Chemical sensors detect the magnitude of electric currents generated during the oxidation or reduction process of analytes with activity. In this manner, they have been used for the continuous detection of lactic acid, uric acid, glucose, etc., with the help of specific enzymes (e.g., glucose oxidase, lactate oxidase, and urea oxidase) that catalyze chemical reactions of the target metabolite to generate an electric current, allowing the representation of the target analyte’s concentration as the intensity of the generated current. For example, Chen et al. [101] designed and fabricated flexible electrochemical luminescent sensors by immobilizing highly luminescent nanospheres on an AuNT network and coating them with elastic molecularly imprinted polymers. The prepared sensors exhibited continuous and desirable mechanical compliance while generating very stable electrochemically luminescent signals during deformation, helping to detect physiologically relevant chemicals in the body in a highly selective manner. They successfully integrated these sensors into flexible wearable sensing devices, thus providing a promising new method for non-invasive monitoring of metabolites for healthcare and biomedical research. Further, Nakata et al. [99] developed a wearable sweat chemo-sensor for measuring pH in human sweat consisting of an ion-sensitive field-effect transistor integrated with a flexible temperature sensor. Liu et al. [102] prepared a flexible electrochemical sensor for non-invasive urea monitoring with high selectivity in human sweat by electro-polymerizing EDOT monomers on a hierarchical network of CNT and AuNT to imprint template molecules. The sensor successfully analyzed urea levels in external solutions and human sweat with satisfactory sensitivity and high selectivity, providing an effective and promising method for non-invasive monitoring of urea levels in human body fluids.
Chemical sensors are sensitive to certain gases, reacting with the tested gas and generating an electrical signal proportional to the gas concentration. The further development and applicability of such functionality have attracted the most interest. Figure 9 shows the application of flexible chemical sensors. Ho Seo et al. [103] developed a self-powered H2 gas sensor using a photovoltaic cell and a chemically mechanically deformed Pd-PUA nano-grid structure. The sensor’s mechanism was based on the change in the light transmittance of the nano-grating structure caused by the expansion of Pd upon reaction with H2 gas, which was then detected by the change in the photovoltaic cell’s output current. The sensor required no external power source thanks to acquiring its necessary energy from ambient light. It was capable of detecting gaseous hydrogen at concentrations as low as 0.1%; was selective for CO, H2S, and NO2; and demonstrated good stability against changes in the humidity. Additionally, the sensor exhibited reliability and reproducibility, ensuring its long-term use in practical situations. Meanwhile, Bezdek et al. [104] based SWCNTs on non-covalently functionalized P4VP, a carbon nanotube that doped (Pt-POM) CH4 oxidation catalyst onto the sensor through P4VP coordination, resulting in a significant improvement in all sensor properties. Additionally, Shin et al. [105] performed a simple co-functionalization of WO3 nanofibers with alkali metal (Na) and noble metal (Pt) catalysts by the simple addition of NaCl and Pt nanoparticles followed by electrostatic spinning, which produced a highly sensitive and flexible H2S gas sensor. Without any collection or filtration equipment, this sensor offered the possibility of direct, reliable, and fast detection of H2S in human breath.
With growing numbers of nanotechnological innovations, the application of material science to flexible electronics has promoted the rapid development of physical and chemical sensors. This has provided a lot of potential for the non-invasive detection of biomarkers in body fluids such as human sweat and saliva, and the diagnosis of physical health-related conditions, a vital part of maintaining and improving people’s standard of living and well-being. A comparison of flexible sensors using different detection methods is shown in Table 2.

4. Contact Forms of Flexible Sensors

4.1. Contact Flexible Sensors

Contact flexible sensors function primarily via physical contact caused by its own deformation to achieve the detection of stress, strain, and pressure. To obtain and collect this information, deformation can only be monitored in discrete locations and specific routes, meaning the measurement capacity of small-scale strain is limited. They require a certain measuring force to ensure the detection component of the sensor is in contact with the object to be measured and to be fully subject to the changes brought about by the deformation.
In recent years, contact flexible sensors notable for being highly sensitive, stretchable, wearable, and easily integrated into devices have shown great potential and prospects for application in the study of intelligent robots, artificial skin, prosthetic limbs, and portable medical devices. However, they are still limited to recognizing mechanical or mechanical deformations and can currently only sense simple physiological signals such as human pressure, pulse, and vocal cord vibration. Pertinently, Naim et al. [117] developed a novel low-cost, active, dry-contact surface electromyography (sEMG) sensor for a bionic arm. The sensor used stainless-steel electrodes to acquire signals from the surface of the skin, and could be connected to a variety of amplifiers, satisfying the need for an economical and wearable sEMG acquisition system for the development of prostheses. This sEMG sensor and its optimal arrangement was useful for achieving highly accurate individual finger control for the bionic arm and did so relatively cheaply. In a separate development work, Yao et al. [118] proposed a flexible thin-film hand sensor, designed to reliably measure the contact pressure distribution at the viscoelastic handle interface and detect the contact force generated by the limb grasping the tool at that location. Innovatively, Wang et al. [52] constructed a high-performance, energy-efficient, and fully flexible piezoelectric haptic sensor by integrating piezoelectric materials with the electromechanical conversion function and signal of PVDF. Additionally, Peng et al. [119] prepared a flexible p-GaN film/n-ZnO nanowire via laser peeling and direct hydrothermal growth of the nanowires on a film light-emitting diode sensor array, which yielded a mapped distribution of two-dimensional pressure values. It did so by reading the light intensity of all light-emitting diode pixels in parallel. At the same time, the local compressive strain enhanced the intensity of each pixel composed of GaN/ZnO nanowire heterostructures of light-emitting diodes thanks to the piezoelectric photoelectric effect.

4.2. Non-Contact Flexible Sensors

Non-contact flexible sensors are defined as those for which the detection of physical quantities is fulfilled by means of non-direct contact. Related technology is widely used in smart wearables, penetrating all aspects of sensor power supplies, and relevant to signals, their generation, modulation, detection, and transmission, etc. Commonly used non-contact methods include optical and thermal detection, and wireless transmission. They currently use the electronic properties of the atomic structure inherent to the surface on certain nanomaterials (such as carbon nanotubes and graphene), or use high-resolution optical sensors, which can not only achieve high spatial resolution and sensitivity but also full-field sensing within a certain range. Thus, they can effectively improve the comprehensiveness and continuity of monitoring and surveillance structures.
These non-contact flexible sensors have received a great deal of attention due to their great potential in human–computer interaction, health monitoring, and healthcare, and their influence has grown alongside the expansion of technological services (i.e., smartphones and internet of things), with increasing demand in light of the coronavirus epidemic. Chen et al. [120] developed a speckle map analysis for a new fiber-optic non-contact displacement sensor with an accuracy and repeatability of 10 nm and a range of 3–4 m. The fiber-optic sensor had a bandwidth of ±5 Hz and could precisely record target positions at 100 Hz. High accuracy and repeatability were achieved using the sensitivity of the fiber optic scattergram and rapid change in the electric field in the near-field region. Taking an alternate course, Guo et al. [121] developed an artificially innervated foam encased piezoresistive non-tactile sensor with self-healing inspired by the mechanosensory based on innervation. The sensor used a new elastic polymer with a metal particle composite sensitive to the direction of any applied contact force, and to the proximity of the human body. In a further innovation, Zhang et al. [122] formed the first flexible sensor based on a flower-like PANI-coated filter paper using a spraying method. Upon preparation, the paper formed a special interwoven flower-like structure, which provided high permeability and bending stability, enabling rapid non-contact detection of nitro-aromatic explosives. Moreover, the results of Tang et al. [123] showed high humidity measuring behavior in both rigid and flexible substrates with a sensor based on SnS nanoflakes. The sensor was able to accurately monitor human breathing patterns and rapidly recognize fingertip movements, showing great promise for applications in physiological evaluation, portable diagnostic systems, and non-contact interface localization.
In recent years, the ability to exchange information between humans and machines has relied heavily on contact-based mechanisms. However, such interactions can lead to serious problems such as unavoidable material degradation through use and potential pathogenic cross-contamination between users. Therefore, a revolution in contactless human–machine interfacing is essential. In this vein, Lu et al. [124] made a flexible, high-sensitivity humidity sensor and array via electrospun PA. The sensor detected the symptoms of asthma in a non-contact manner by monitoring the respiratory rate in real-time. Coupled with a remote alarm system, it could provide a non-contact interface for administering medication to bedridden patients. Consequently, the study provided an effective approach to further development of smart electronics for non-contact scenarios. Tang et al. [125] designed a frictional electric touchless screen sensor from a graphene/ITO/PET triple-layer structure for recognizing various gestures in non-contact operation mode using human body electric charges. The sensor was lightweight, flexible, transparent, could be easily integrated into a smart electronic screen, and was effective at detecting various gestures such as the raising or lowering of a finger, fist clenching, opening of the palm, and palm flipping in different directions and speeds. The results compared positively against conventional capacitive sensors. This innovative frictional electric touchless screen sensor is expected to make a significant breakthrough in intelligent human–computer interaction.
Compared to contact sensing, the non-contact type is more convenient in a variety of situations. Non-contact sensors are equally capable of monitoring the human heart rate, respiration, and other physiological parameters without contacting the human body. Further, this method of monitoring can allow the detection of breathing, pulse, heart rate, and other signals without disturbing the normal activities of the human body. At the same time, because nothing needs to be worn by the test subjects, they may not be aware of the measurement process itself during an extensive data collection period, thereby eliminating the influence of the subject’s psychological responses on the test results; therefore, more objective information can be obtained. Table 3 presents a comparison of the numerous flexible sensors that use different materials and methods of preparation. It analyzes several aspects of their practical performance, including the sensitivity factor, effective linear range, and cycling stability. The results demonstrate that there are still challenges in obtaining optimal flexible sensors (i.e., those with high sensitivity and a good effective linear range at the same time).

5. Commonly Used Materials for Flexible Sensors

5.1. Flexible Substrates

In order to meet the requirements of flexible electronic devices, materials that are thin, light, transparent, flexible, highly stretchable, insulated, and corrosion resistant have become ideal as applicable substrates. Among the common substrate materials are PET, PMDS, PAN, PU, PVA, PI, and textile materials. Flexible substrates are applied based on how they are deformed (common types are bending, plasticity, and uniaxial, biaxial, and ray-like stretching), and how they store and collect energy. Chen et al. [133] proposed an electrochemically assisted deposition method to deposit GO onto crossed finger electrode patterned PET substrates, and after reducing the resulting films using hydrazine vapor, reduced oxide GO films on bridged PET substrates were obtained. This GO oxide film showed excellent electronic properties and an outstanding gas-sensitive performance. Shang et al. [134] prepared a flexible substrate fluxgate based on MEMS technology and used the prepared flexible substrate fluxgate as a current sensor for online measurement. Jia et al. [135] introduced PAN to TPU and, via electrostatic spinning, fabricated a flexible TPU/PAN film. This was followed up with the preparation of a highly conductive and stretchable Ti3C2 MXene/TPU/PAN film by a simple dip coating process. By introducing PAN to the flexible substrate, this group was able to effectively improve the interaction between Ti3C2 MXene and the substrate without affecting the electrical conductivity of Ti3C2 MXene. Li et al. [136] used in situ chemical oxidation polymerization to synthesize PANI and GO-PANIHs hybrids with a rambutan-like hollow nanosphere structure. Based on the GO-PANIH and flexible PET substrates, the prepared sensors could detect NH3 at room temperature without additional electrodes. Moreover, Lin et al. [137] successfully fabricated a flexible and highly sensitive LSG/Cu-NPs glucose sensor using DVD laser-scribed graphene as the conductive substrate. Song et al. [138] synthesized a sponge-like sugar cube PDMS scaffold as a template and developed a porous structure CNT/PDMS sponge with high electrical conductivity and mechanical properties through the drip-drying process of CNT ink dispersion. The CNT–PDMS sponge has high sensitivity as a piezoresistance sensor, which is capable of detecting various stress and has high sensitivity. Moreover, he also developed a highly compressible integrated system comprising a piezoresistance sensor and a compressible supercapacitor based on the wearable CNT/PDMS sponge. Wu et al. [139] first mixed CNF and sugar cube particles as a template and waited until the PDMS solidified after pouring it to remove the sugar cubes to obtain a PDMS/CNF composite. This method transfers the CNF that is originally on the surfaces of the sugar particles and embeds it in the pore walls of the porous PDMS, which ensures good adhesion between the CNFs and the polymer. In 10,000 compression-release cycles of 30% strain, PDMS/CNF showed a better performance of good repeatability. Meanwhile, PDMS/CNF also has stable piezoresistive properties, with good linearity under 70% strain. Using PDMS for swelling in organic solvents, it can be used to monitor the movement of human joints.

5.2. Metallic Materials

Relevant metallic materials are generally conductors such as gold, silver, and copper, which are most commonly used for electrodes and wires. For the modern printing process, conductive materials are used for nano-inks, including nanoparticles and nanowires, etc. In addition to good electrical conductivity, metal nanoparticles can also be sintered into thin films or wires. Liquid metals are pertinent materials with low melting points, often at or near room temperature. Due to their high electrical conductivity, low toxicity, and excellent mobility, they are becoming highly desirable candidates as flexible sensors. Kawasetsu et al. [140] developed a flexible and soft inductive triaxial haptic sensor using liquid metals as sensing targets. Liu et al. [141] produced a flexible polymer-based photoelectrochemical sensor by layer-by-layer self-assembly on graphene, which more than tripled its photocurrent. Chen et al. [142] investigated the use of liquid metals with different means of adhesion on various material surfaces to print a circuit with certain patterns on thermal transfer paper. It was then transferred to a flexible substrate, leading to a stretchable flexible pressure sensor. Recently, the discovery of infiltrating metal oxide nanostructures with noble metal nanoparticles has opened the door to the fabrication of biosensors for the detection of volatile organic compounds. Almog et al. [143] used CS’s affinity for heavy metal ions and the functional properties of graphene oxide to generate an environmentally benign organic nanocomposite for flexible electrochemical sensors.
Metal nanowires, especially silver nanowires, are considered the best materials for flexible sensors. Even in highly twisted and multi-cycle test experiments, the excellent aspect ratio and good electrical conductivity of metal nanowires can guarantee a good conduction effect. Yin et al. [144] prepared biodegradable conductive AgNW/CNF hybrid nanoparticles by surface solution blending and vacuum filtration. Due to the amphiphilic property of cellulose, AgNW can be uniformly dispersed, and an effective conductive network can be constructed. In order to study the application of AgNW/CNF nanoparticles in strain sensors, a sandwich strain sensor with a typical microcrack structure was prepared using the pre-strain technique. Due to the different crack densities of structures under different prestrains, interesting and typical prestrain-dependent strain sensing behaviors are observed. Mixed nanoparticles as temperature sensors also exhibit stable and reproducible negative temperature-sensing behavior. This study guides the fabrication of flexible and biodegradable sensors. Shi et al. [145] manufactured a flexible transparent capacitive pressure sensor based on a patterned microstructure AgNWs/PDMS composite dielectric layer material. The dielectric layer material was modified to improve the sensitivity of the device. By mixing AgNW with PDMS at different concentrations, the optimal doping ratio of AgNW was 0.12 wt%. Then, the dielectric layer was microstructured, and a symmetrical PET/ITO film electrode was used. The patterned microstructured sensor had a higher sensitivity (0.831 kPa−1), a lower detection limit (<0.5 kPa), and good stability and durability compared to a pure PDMS dielectric layer with a planar structure.

5.3. Organic Materials

Large-scale pressure sensor arrays are important for the development of future wearable sensors. Those based on piezoresistive and capacitive signal mechanisms suffer from signal crosstalk, which leads to inaccurate measurements, one of the biggest challenges for wearable flexible sensors. To address this, high-performance organic crystalline materials are considered as strong candidates for the next generation of flexible electronics to be used in displays, image sensors, and artificial skins, amongst others [146]. Not only do they offer flexibility, molecular diversity, low cost, solution processability, and inherent compatibility with flexible substrates, but they also have few grain boundaries and defects, ensuring excellent and uniform electronic properties [147]. Moreover, such materials can be used as a powerful tool to explore the intrinsic electronic and mechanical properties of organic matter, reveal the physical properties of flexible devices, and provide further guidance for the future of the materials themselves and the design of the devices that use them. Lai et al. [148] developed the organic charge-modulated field-effect transistor, a new structure for sensors capable of operating at low voltages, which was shown to be fully applicable across a large swathe of technologies, e.g., from inkjet printing to chemical vapor deposition. Meanwhile, Matsui et al. [149] investigated recent advances and developments in flexible and printed organic thin film transistor devices, including organic materials, fabrication processes, electronic devices, and integrated circuits, highlighting their applications in medical sensors.

5.4. Inorganic Semiconductors

Inorganic semiconductor materials have excellent electrical properties but suffer from being brittle and exhibit poor machinability and deformability. Conversely, organic semiconductors are flexible but have generally poorer electrical properties. Researchers have been able to achieve a certain degree of flexibility by preparing inorganic semiconductor films on substrates to reduce the material stiffness, when combined with a suitable structural design. However, this method does not change the intrinsic brittleness of inorganic semiconductors, and there are still major limitations in the processing, fabrication, and applications. Therefore, inorganic semiconductor materials represented by ZnO and ZnS have shown promising applications in the field of wearable flexible electronic sensors due to their excellent piezoelectric properties. Zhang et al. [150] developed an ultrasensitive flexible room temperature gas sensor for lung cancer monitoring and diagnosis based on Na-doped ZnO/reduced GO. Shirley et al. [151] devised a textile pressure sensor using a sandwich model of the semiconductor material ZnO placed between two conductive fabric electrodes. The sensor could be used for a variety of applications in the clinical setting and for tracking human motion. Depending on the target application, the sensor pressure range and sensitivity could be adjusted and optimized while the use of textiles provided a strategy for the effective development of further flexible fabric-based health monitoring sensors. Sholehah et al. [152] successfully fabricated ethylene gas sensors using ZnO-Ag layers on flexible PET-ITO substrates, deposited via electrochemical techniques. Guan et al. [153] built on the flexible ZnO/PAN/nonwoven nanocomposites to produce a gas-phase poly combined with a high-sensitivity NH3 gas sensor. Additionally, Chen et al. [154] developed a watch-type wearable NO2 sensor whose core material was ZnS nanoparticles/nitrogen-doped reduced oxide GO. It had a low power consumption (0.52 μW), a low detection limit (69 ppb), and was capable of maintaining its mechanical durability. This sensor provided timely monitoring and early warning of NO2 leakage by reading and analyzing signal data via wireless Bluetooth transmission.

5.5. Carbon Materials

Carbon materials commonly used for flexible wearable electronic sensors include graphene and carbon nanotubes. The former is light, thin, transparent, has good electrical and thermal conductivity, and has extremely important and promising applications in sensing technology, mobile communication, information technology, and electric vehicles. Mamleyev et al. [155] used optimized laser radiation to pattern graphitic carbon structures onto micron-scale PI sheets and induced rapid local pyrolysis to produce a flexible device. The resulting laser carbon films displayed an excellent electrical conductivity and high surface area with a graded porosity distribution along their cross-section. Qi et al. [156] prepared self-supported flexible enzyme-free sensors by electro-depositing PPy on the surface of flexible carbon paper. This was followed by hydrothermal growth of CoS nanospheres, leading to a final sensor that could be applied to glucose detection in humans. Patil et al. [157] used fine gold nanoparticles doped with porous carbon as electrode materials to fabricate flexible biosensors with high accuracy and reliable operation for the detection of pH and uric acid levels in body fluids. Zhang et al. [158] developed a freestanding flexible film made of graphene/EC nanocomposite, which was integrated with a PET substrate. This gave rise to a highly sensitive flexible gas sensor with an ultra-low strain response. Li et al. [159] prepared flexible fork finger electrodes using silver nanowires, with reduced GO serving as the gas-sensitive material. Li et al. [160] fabricated a novel heterostructure enhanced sensitivity stress/strain sensor consisting of 2D graphene and 1D PZT nanowires as sensitive materials. The constituent carbon nanotubes were characterized by a high crystallinity, good electrical conductivity, large specific surface area, microporous size that could be controlled by the synthetic process, and specific surface utilization up to 100%. Concerning functional nanomaterials, MWCNTs are widely used strain-sensing polymer composites. Due to their unique one-dimensional structure, they are expected to exhibit high anisotropy when aligned along a certain direction, which has a positive effect on improving the sensory performance of pristine polymer composites. Building on this, Laing et al. [161] demonstrated a scalable contact that could be easily integrated into a resistance-based pressure sensor based on CNT conductive networks and photo-resistant insulating layers. Additionally, Hao et al. [162] developed a flexible strain sensor with nanocomposites filled with three-dimensional CNT foam while Huang et al. [163] prepared a flexible lightweight CNT/TPU pressure sensor with a novel aligned porous structure, which demonstrated excellent compressibility and stability.

5.6. Two-Dimensional Materials

Two-dimensional nanomaterials have many unique physical properties, high mechanical flexibility, a large specific surface area, and good electrical conductivity, making two-dimensional nanomaterials a good choice as conductive sensitive materials for piezoresistive sensors. Two-dimensional nanomaterials can be widely used in flexible sensors by adjusting their structure, lamellar thickness, and surface modification [164]. These include TMDs and MXene. TMDs refer to compounds formed by transition metals and sulfur atoms and are widely used in molybdenum disulfide (MoS2), vanadium disulfide (VS2), tungsten disulfide (WS2), tungsten diselenide (WSe2), etc. Park et al. [165] studied a large-area tactile sensor based on an active matrix and utilization of the semiconductor and mechanical properties of MoS2. The sensor can measure pressure from 1 to 120 KPa, which is far better than the sensing range of human skin. In addition, the sensor has the advantage of multi-point susceptible detection, which can accurately identify the object’s shape grasped by the human hand by simultaneously monitoring the external pressure in different positions. Yang et al. [166] prepared a high-performance pressure sensor based on 1TMoS2 in PDMS foam as the conductive active layer and layered micromesh veins as spacers. Due to the excellent electrical conductivity and novel structural design, 1Tmos2-PDMs perform better than previously reported devices. It has a high sensitivity of 1036.04 KPa in a wide linear range of 1 to 23 KPa, a fast response time (<50 ms), an ultra-low detectable pressure limit of 6.2 Pa, and excellent repeatable loading and unloading stability (10,000 cycles).
MXene has excellent electrical conductivity, chemical stability, mechanical flexibility, and hydrophilicity. The electrical and optical properties of MXene can be further optimized by appropriate surface modification and functionalization to expand the application range of MXene in flexible wearable sensors. Yue et al. [167] developed a 3D hybrid porous MXene sponge piezoresistive sensor using a dip-coating process and PVA nanowires as spacers. The large contact area between the MXene and sponge and the strong van der Waals force mean that MXene could be fixed on the sponge well. Due to the sponge’s rich porous structure, the sensor showed good compressibility and could be fully restored to its original state after more than 95% of the volume compression. The MXene sponge sensor had high sensitivity over a wide pressure range, low detection limits, fast response times, and excellent durability over 10,000 cycles. Chen et al. [168] prepared lightweight MXene-based aerogel by connecting MXene(Ti3C2) nanosheets into continuous wavy sheets using bacterial cellulose fiber as a nano-binder. The aerogel’s ultra-high structural stability allowed it to withstand 99% extremely high strain over 100 cycles and long-term compression at 50% strain over at least 100,000 cycles. In addition, it has high sensitivity, high linearity, and a low detection limit for minor strains and pressures. Yang et al. [169] spun PU dispersion into stretchable PU felt. Then, the MXene solution was dropped on PU to form the Mxene-PU interlocking conductive network, and then the double-sided carbon adhesive was used to lead the wires at both ends of the network to prepare the MXene/PU strain sensor. When subjected to tensile stress, adjacent MXene sheets covering PU slide between each other, and the internal conductive path is disconnected or connected, resulting in a change in the output resistance. The strain detection range of the sensor is 1~100%, which can be used to detect a pulse, heart rate, finger bending, arm bending, and other human activity signals.

6. Applications of Flexible Sensors

Electronic devices of the future will become even more miniaturized, lightweight, and versatile. Without interfering with human movement, they can be centralized in our clothes or on our skin to detect changes in health-related conditions and human movement. These applications are highly coordinated, thus stimulating research interest in making them widely applicable to electronic skin, health monitoring, motion monitoring, smart healthcare, and human–computer interaction.

6.1. Electronic Skin

Electronic skin (or e-skin) is a new type of flexible wearable sensor that is transparent, soft, and thin like human skin. By attaching e-skins to the fingers and arms of robots, robots gain the ability to sense external physical contact. Inspired by human skin, many have worked to reproduce haptic functions, but e-skin is needed not only to gain tactile perception but, more importantly, to also differentiate multiple stimuli simultaneously to further gain multidimensional perception. E-skin is the core of the future wearable electronic device network, which has gained wide attention in research into human–machine interface, robot smart skin, medical monitoring, and bio-integrated devices due to its multi-functionality, ultra-thinness, low or even zero energy consumption, flexibility, and biocompatibility. Compared with traditional rigid, brittle, silicon-based sensors, e-skins, which do not suffer such flaws, can effectively capture high-quality signals on curved surfaces due to their elasticity. E-skins are expected to play an important role in the era of intelligence ahead.
Dong et al. [170] designed a stretchable and washable skin-inspired friction nanogenerator for biomechanical energy accumulation and multifunctional pressure sensing. Elsewhere, Zeng et al. [171] developed a tunable pressure sensor with ultra-high sensitivity based on folded microstructures (Figure 10a). By changing the aspect ratio and amplitude of the microstructure in a controlled manner, the sensitivity and operating range could be adjusted, respectively. The sensor achieved a maximum sensitivity of 1.713 KPa−1. With a highly compressible microstructure, it also exhibited a relatively low detectable pressure limit (1.5 Pa), fast response time (<50 ms), wide pressure range, and excellent cycling stability. Lee et al. [172] developed a multilayer flexible pressure sensor by simultaneously controlling the tactile stress delivered to the active sensing region and the corresponding current through the gradient structure (Figure 10b). The sensor had an ultra-high level of sensitivity (3.8 × 105 Kpa−1) and a very wide pressure range, spanning up to 100 KPa. Additionally, the multilayer pressure sensor was observed to have a fast response time (0.016 ms) and a low minimum detectable pressure (0.025 Pa). This sensor successfully demonstrated diverse applicability, from detecting sounds to weak air pressures, and from mapping pressure in general to monitoring individual health.

6.2. Health Monitoring

Lightweight and wearable flexible electronic devices are essential for personal healthcare systems and are not constrained by time and space. In order to monitor human biosignals in a non-invasive manner, highly sensitive, reliable, and sustainable healthcare monitoring devices are required. Highly sensitive sensors can detect small skin strains due to blood flow, pulse, and respiration, and digitize this information for an analysis the subject’s health, such as blood pressure, respiratory rate, and heart rate.
Kano et al. [173] developed a fast-responding nanocrystal-based humidity sensor, which varied the current reading by 5 orders of magnitude over an 8–83% relative humidity range (Figure 11a). The sensor exhibited a fast response to humidity changes with a response/recovery time of 40 ms at 3 Hz. The research team demonstrated its use in monitoring human breathing and moisture evaporation from the skin in real-time. Furthermore, Shi et al. [69] proposed a simple and low-cost method for fabricating flexible piezoresistive pressure sensors with a large area layered structure. Sun et al. [174] assembled a conductive nanocomposite hydrogel composed of oxidized MWCNTs and PAM (Figure 11b). Xu et al. [175] studied and produced a multifunctional flexible sensor based on two different graphene films, which can detect physiological signals and biomarkers simultaneously without mutual interference. The sensor detects different human physiological signals, such as the pulse and respiratory rate. It has an excellent detection ability by directly attaching to different human body parts (Figure 11c). He et al. [73] developed a cost-effective nylon mesh for a simple capacitive sensor (Figure 11d), which incorporated a nylon mesh with regularly distributed microscale square holes. This sensor exhibited high sensitivity up to 0.33 KPa−1 in the low-pressure range and continued to do so after more than 1050 repetitive cycles at 400 and 1000 Pa. This confirmed its long operational durability and an ultra-low detection limit below 3.3 Pa, which can be applied to real-time monitoring of the human pulse.

6.3. Motion Monitoring

The data obtained by those flexible sensors applied to human motion monitoring is relatively simple, predominantly based on basic flexion of the limbs or extension movements. Consequently, the data analysis in sports science or fitness has not been fully brought up to date with real scientific requirements, and the data mining, comprehensive analysis, and means of utilization need to be continuously improved. In addition, motion monitoring is a time-consuming process in terms of output and feedback, especially for the module of data collection and transmission. Their working mode is often constrained by the battery’s power reserve, with an insufficient energy supply adversely affecting the accuracy and stability of the collected information.
Wang et al. [176] devised a flexible resistive strain sensor with a special three-dimensional conductive network using a flexible TPU fiber mat decorated with reduced GO (Figure 12a). This strain sensor demonstrated excellent electrical properties, hyper-elasticity, high sensitivity, good durability, and fast responsiveness. The full range of human motion detection (including large and fine human motions such as walking, jumping, finger bending, coughing, vocalization, and fine muscle movements of the cheeks) indicated this sensor had a wide range of applications. Using the solution method and ultrasonic dispersion technology, Yang et al. [177] prepared a room-temperature RTV silicone rubber based on CF and CB with excellent elasticity and high electrical conductivity of rubber composites. The sensitivity properties and strain sensing mechanism of CF/CB-RTV silicone rubber sensors were investigated, and it was concluded that the prepared composites have great potential as smart monitoring strain sensors and flexible electrodes. This study thus provided an important alternative solution for the fabrication of low-cost and high-performance strain sensors and electrodes based on conductive elastomer composites. Ergen et al. [178] developed an aerogel-based strain and sweat sensor able to efficiently extract real-time information by combining involuntary human motion and chemical signals. These sensors possessed good mechanical integrity, allowed high-density electrical energy generation during fine body movements, and could be applied to continuous human motion monitoring by measuring the pH, ion concentration, and volume of secreted sweat. Pu et al. [179] fabricated sensors for targeted detection by oxidation of a Ti3C2Tx MXene surface to form nanocrystalline TiO2, which were surmised to function by forming a sensor with strongly interacting sensing layers. The strain off-domain was efficiently converted into strain localization by introducing a loose isolated network structure in the sensing layer and two distinct sensing layers with microcrack and through-crack patterns were easily fabricated. The sensors exhibited an excellent performance, including flexibility, fuzzy drift, frequency stability, fast response and relaxation time, and long-term durability. It was concluded that it is highly feasible to achieve a full range of human motion monitoring demonstrations by selectively using sensors for targeted deformation detection. Ma et al. [180] manufactured a strain sensor with high sensitivity and an extensive working strain range based on 3D printing technology and plasma processing (Figure 12b). The concept of a self-compensating second-order structure is proposed for the first time to balance the sensitivity within the strain range (0–350%). At strains below 20%, the sliding and breaking of the graphene sheet coated on the grid surface dramatically improved the sensor’s sensitivity, up to 20 times as much as that of the non-second-order structure sensor. This unique structural design enables sensors to detect the full range of human motion, from pulse and acoustic vibration to breathing and arm bending, providing a new approach for the development of a new generation of full-range strain.

6.4. Smart Healthcare

The creation of new intelligent medical wearable products that feature flexible sensors is an important step for promoting the continuous development of advanced healthcare. Through big data, cloud computing, and other technological applications, sensors collect a large amount of data concerning users’ health, behavior, and habits in real-time, providing important support for medical analysis. Matching this, internet-based medical and health data systems are rapidly evolving, allowing doctors to develop timely and effective healthcare plans based on the analysis of routine patient examinations and clinical studies. To obtain this data accurately, continuous and repeated measurements must be taken with wearable devices, which can be used for monitoring patients when they exercise or sleep. Combining medical systems with wearable sensors [181] and applying big data to the construction of medical platforms will allow for the effective analysis of large quantities of recorded information.
Xu et al. [182] designed an adjustable external fixator for the lower limbs, which was able to realize timed intelligent temperature control and meet local fixed-point cooling. The adjustable structure and application of an intelligent air cushion was hypothesized to meet the adequate fixation of the lower limbs of individuals with varying statures and solve the problems of emergency transportation and subsequent treatment of related injuries, such as those commonly suffered by winter sports athletes. Liu et al. [183] introduced a wireless sensor network into a healthcare system and proposed a new form of imaging system. Feng et al. [184] developed a motion monitoring technique for flexible integrated sensor systems that is able to independently measure bending strain and pressure coupled with a low-cost and simple manufacturing process. Yu et al. [185] prepared ultra-flexible and lead-free piezoelectric nanogenerators as highly sensitive self-powered sensors for monitoring human motion, which displayed promise for applications in smart healthcare.

6.5. Human–Computer Interaction

Human–computer interaction refers to the exchange of information, via a number of methods, between a user and a digital system, whether it be a machine or software. With the development of computer technology, the interactive devices people use are no longer limited to just keyboards or mice but can also be smart wearable sensing devices.
Wan et al. [186] posited a multimodal artificial sensory memory system consisting of sensors to generate bionic visual, auditory, and tactile inputs, and flexible carbon nanotube synaptic transistors with synaptic-like signal processing and memorization behavior. Zhao et al. [187] developed pressure sensors based on fabricated flexible organic friction transistors with a high sensitivity, fast response time, and excellent stability. Liu et al. [188] proposed a multifunctional dual-network composite hydrogel with good biocompatibility, tensile, and self-healing properties, which showed great potential for human–computer interactions.
In recent years, the field of human–computer interaction has faced great challenges, and high sensitivity, accuracy, reproducibility, mechanical flexibility, and low cost have become requirements for new products that might be in general use. Recent advances in soft materials and system integration technologies offer a unique opportunity for their use in advanced healthcare and human–machine interface design. The hybrid integration of biocompatible materials with miniaturized, wireless wearable systems is an attractive prospect.

7. Conclusions

Over the past few years, many researchers have researched and made significant progress in the improvement of flexible sensors. Challenges in terms of accuracy, reliability, high power consumption, rigidity or large volume factors, and difficulties in data interpretation have limited their wider application. Flexible sensors that manage to meet the challenges of high sensitivity, high accuracy, high reproducibility, high mechanical flexibility, and low cost are urgently needed for smart wearable sensing devices to enable their practical application in areas such as e-skin, health monitoring, motion monitoring, smart healthcare, and human–computer interaction. On this basis, the focus should be placed on three aspects of flexible sensors: sensory mechanisms, the subject of detection, and the form of contact, with the hope of providing references for the research and construction of new high-performance structural sensors.
However, despite great advances in the research on flexible sensors, new sensing mechanisms and types of microstructure design should be further explored to develop them with an even better overall performance. At present, particular performance issues (i.e., achieving high sensitivity and a large scale of monitoring at the same time) still need to be solved. In addition, their practical application on the human body or in harsh environments is more limited, so the development of more environmentally friendly sensors is the focus of future research. As more research is conducted, more functions will open for flexible sensors, including self-healing and biocompatibility, etc. The main problem in the future will be how to integrate multiple sensing functions and power sources into a single intricate device. An example of this is a multifunctional system similar to human skin, which requires the incorporation of multiple electronic components and sensors onto very small soft substrates, thus requiring large-area, low-cost integration and manufacturing techniques. At the same time, multifunctional sensors will generate a large amount of electrical signal information. Thus, the matter of collecting and maintaining the accumulated data, and ensuring the functionality and long-term stability of integrated sensors need to be further explored. Therefore, it may be safely concluded that such sensors will become mainstream in the next generation of smart wearable sensing devices.

Author Contributions

Literature research and discussion, Y.Y. and C.G.; writing, Y.Y. and C.G.; methodology, H.Y. and F.X.; supervision, Y.Y. and D.C.; project administration and funding acquisition, H.L. and D.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Key Research and Development Program of China (No. 2021YFC2009100), the National Natural Science Foundation of China (No. 62172081), the Advanced Textile Equipment Technology Provincial and Ministerial Co-construction Collaborative Innovation Center (No. 410012310003), the College Teaching Master Team of Henan Provincial (No. 32500094), and the Research Foundation for Young Doctor Teachers of Zhongyuan University of Technology (No. 34110557).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

Chitosan: CS; Polydimethylsiloxane: PDMS; Carbon Nanofibers: CNF; Two-dimension transition metal carbides, nitrides, and carbonitrides: MXene; Thermoplastic Polyurethane: TPU; Carbon Nanotubes: CNT; Graphene Oxide: GO; Isopropyl alcohol: IPA; Silicon Dioxide: SiO2; Polyaniline: PANI; Poly(Lactic-co-glycolic Acid): PLGA; Carbon Conductive Paste: CCP; Sodium Chloride: NaCl; Polystyrene: PS; Graphene Nano: GNP; Multi-walled Carbon Nanotubes: MWCNT; Silver Nanowires: AgNWs; Polyethylene Terephthalate: PET; Zinc Oxide: ZnO; Polyvinyl Alcohol: PVA; Polyethyleneimine: PEI; Polyurethane: PU; Activated Carbon: AC; Tetramethylammonium Hydroxide: TMAH; Polyvinylidene fluoride-trifluoroethylene: P(VDF-TrFE); Single-walled Carbon Nanotubes: SWCNTs; Fluorinated ethylene propylene: FEP; Polyacrylonitrile: PAN; Barium Titanate: BTO; Polyethylene Glycol: PEG; Poly(ethylene glycol) diacrylate: PEGDA; Silver Nanoparticles: AgNPs; Thin Film Transistors: TFT; Boron Nitride Nanosheets: BNNS; Polyacrylamide: PAM; Polyacrylic Acid: PAA; Polytetrafluoroethylene: PTFE; Polylactic Acid: PLA; 3,4-Ethylenedioxythiophene: EDOT; Poly(4-Vinylpyridine): P4VP; Platinum-polyoxometalate: (Pt-POM)CH4; Tungsten Trioxide: WO3; Indium Tin Oxide: ITO; Poly(ionic liquid)s: PVIM; Cobalt Polyoxometalate: Co5POM; Poly(3,4-ethylenedioxythiophene): PEDOT; Polystyrene Sulfonate: PSS; Ethylene Glycol: EG; Toluene Diisocyanate: TDI; 2,2,6,6-tetramethylpiperidinooxy: TEMPO; Fluorododecyl Trichlorosilane: SigmaAldrich; Gallium Mononitride: GaN; Polyethylene Naphthalate: PEN; Polyvinylidene Fluoride: PVDF; Polyvinylpyrrolidone: PVP; Carbonyl Iron Particles: CIP; Ethyl Cellulose: EC; Polypyrole: PPy; Acrylamide: AAm; Sodium Dodecyl Sulfate: SDS; Potassium Persulfate: KPS; Tetramethylethylenediamine: TMEDA; 1-Ethyl-3-methylimidazolium Chloride: [EMIM]Cl; Poly(ethylene glycol) Diacrylate: PEGDA; Poly (2-hydroxyethyl methacrylate): PHEA; Lead Zirconate Titanate: PZT; Transition metal chalcogenides: TMDs; Poly(ethylene glycol) Diacrylate: PEGDA; Room-temperature-vulcanizing Silicone: RTV; Carbon Fiber: CF; Carbon Black: CB.

References

  1. Wang, K.; Lou, Z.; Wang, L.; Zhao, L.; Zhao, S.; Wang, D.; Shen, G. Bioinspired interlocked structure-induced high deformability for two-dimensional titanium carbide (mxene)/natural microcapsule-based flexible pressure sensors. ACS Nano 2019, 13, 9139–9147. [Google Scholar] [CrossRef] [PubMed]
  2. Yan, T.; Wang, Z.; Pan, Z.J. Flexible strain sensors fabricated using carbon-based nanomaterials: A review. Curr. Opin. Solid State Mater. Sci. 2018, 22, 213–228. [Google Scholar] [CrossRef]
  3. Zheng, H.; Lin, N.; He, Y.; Zuo, B. Self-healing, self-adhesive silk fibroin conductive hydrogel as a flexible strain sensor. ACS Appl. Mater. Interfaces 2021, 13, 40013–40031. [Google Scholar] [CrossRef] [PubMed]
  4. Gholami, M.; Napier, C.; Patiño, A.G.; Cuthbert, T.J.; Menon, C. Fatigue monitoring in running using flexible textile wearable sensors. Sensors 2020, 20, 5573. [Google Scholar] [CrossRef]
  5. Moiş, G.D.; Sanislav, T.; Folea, S.C.; Zeadally, S. Performance evaluation of energy-autonomous sensors using power-harvesting beacons for environmental monitoring in internet of things (IOT). Sensors 2018, 18, 1709. [Google Scholar] [CrossRef] [Green Version]
  6. Lou, Z.; Li, L.; Wang, L.; Shen, G. Recent progress of self-powered sensing systems for wearable electronics. Small 2017, 13, 1701791. [Google Scholar] [CrossRef]
  7. Kim, S.; Amjadi, M.; Lee, T.I.; Jeong, Y.; Kwon, D.; Kim, M.S.; Kim, K.; Kim, T.S.; Oh, Y.S.; Park, I. Wearable, ultrawide-range, and bending-insensitive pressure sensor based on carbon nanotube network-coated porous elastomer sponges for human interface and healthcare devices. ACS Appl. Mater. Interfaces 2019, 11, 23639–42368. [Google Scholar] [CrossRef]
  8. Liu, G.; Tan, Q.; Kou, H.; Zhang, L.; Wang, J.; Lv, W.; Dong, H.; Xiong, J. A flexible temperature sensor based on reduced graphene oxide for robot skin used in internet of things. Sensors 2018, 18, 1400. [Google Scholar] [CrossRef] [Green Version]
  9. Zhang, S.; Liu, H.; Yang, S.; Shi, X.; Zhang, D.; Shan, C.; Mi, L.; Liu, C.; Shen, C.; Guo, Z. Ultrasensitive and highly compressible piezoresistive sensor based on polyurethane sponge coated with a cracked cellulose nanofibril/silver nanowire layer. ACS Appl. Mater. Interfaces 2019, 11, 10922–10932. [Google Scholar] [CrossRef]
  10. Choi, H.; Jeong, H.Y.; Lee, D.S.; Choi, C.G.; Choi, S.Y. Flexible NO2 gas sensor using multilayer graphene films by chemical vapor deposition. Carbon Lett. 2013, 14, 186–189. [Google Scholar] [CrossRef] [Green Version]
  11. Bariya, M.; Shahpar, Z.; Park, H.; Sun, J.; Jung, Y.; Gao, W.; Nyein, H.Y.; Liaw, T.S.; Tai, L.C.; Ngo, Q.P.; et al. Roll-to-roll gravure printed electrochemical sensors for wearable and medical devices. ACS Nano 2018, 12, 6978–6987. [Google Scholar] [CrossRef]
  12. Ahn, H.Y.; Kim, J.G.; Gong, M.S. Preparation of flexible resistive humidity sensors with different electrode gaps by screen printing and their humidity-sensing properties. Macromol. Res. 2012, 20, 174–180. [Google Scholar] [CrossRef]
  13. Yang, T.; Deng, W.; Chu, X.; Wang, X.; Hu, Y.; Fan, X.; Song, J.; Gao, Y.; Zhang, B.; Tian, G.; et al. Hierarchically microstructure-bioinspired flexible piezoresistive bioelectronics. ACS Nano 2021, 15, 11555–11563. [Google Scholar] [CrossRef]
  14. Sun, P.; Wu, D.; Liu, C. High-sensitivity tactile sensor based on Ti2C-PDMS sponge for wireless human-computer interaction. Nanotechnology 2021, 32, 295506. [Google Scholar] [CrossRef]
  15. Jiang, F.; Lee, G.B.; Tai, Y.C.; Ho, C.M. A flexible micromachine-based shear-stress sensor array and its application to separation-point detection. Sens. Actuators A Phys. 2000, 79, 194–203. [Google Scholar] [CrossRef]
  16. Kerpa, O.; Weiss, K.; Worn, H. Development of a flexible tactile sensor system for a humanoid robot. In Proceedings of the 2003 IEEE/RSJ International Conference on Intelligent Robots and Systems (IROS 2003), Las Vegas, NV, USA, 27 October–1 November 2003; pp. 1–6. [Google Scholar]
  17. Manunza, I.; Sulis, A.; Bonfiglio, A. Organic semiconductor field effect transistors for unconventional applications: Flexible sensors and wearable devices. In Proceedings of the International Workshop on Wearable and Implantable Body Sensor Networks (BSN’06), Cambridge, MA, USA, 3–5 April 2006. [Google Scholar]
  18. Huang, C.T.; Shen, C.L.; Tang, C.F.; Chang, S.H. A wearable yarn-based piezo-resistive sensor. Sens. Actuators A Phys. 2008, 141, 396–403. [Google Scholar] [CrossRef]
  19. Figueiredo, C.P.; Becher, K.; Hoffmann, K.P.; Mendes, P.M. Low power wireless acquisition module for wearable health monitoring systems. In Proceedings of the 2010 Annual International Conference of the IEEE Engineering in Medicine and Biology, Buenos Aires, Argentina, 31 August–4 September 2010; pp. 704–707. [Google Scholar]
  20. Roh, J.-S. Textile touch sensors for wearable and ubiquitous interfaces. Text. Res. J. 2013, 84, 739–750. [Google Scholar] [CrossRef]
  21. Ji, Z.; Zhu, H.; Liu, H.; Chen, T.; Sun, L. A flexible capacitive tactile sensor for robot skin. In Proceedings of the 2016 International Conference on Advanced Robotics and Mechatronics (ICARM), Macau, China, 18–20 August 2016; pp. 207–212. [Google Scholar]
  22. Li, L.; Bai, Y.; Li, L.; Wang, S.; Zhang, T. A superhydrophobic smart coating for flexible and wearable sensing electronics. Adv. Mater. 2017, 29, 1702517. [Google Scholar] [CrossRef]
  23. Wang, L.; Lou, Z.; Jiang, K.; Shen, G. Bio-multifunctional smart wearable sensors for medical devices. Adv. Intell. Syst. 2019, 1, 19000040. [Google Scholar] [CrossRef] [Green Version]
  24. Baeg, K.J.; Lee, J. Flexible electronic systems on plastic substrates and textiles for smart wearable technologies. Adv. Mater. Technol. 2020, 5, 2000071. [Google Scholar] [CrossRef]
  25. Guo, L.; Li, Z.; Hu, W.; Liu, T.; Zheng, Y.; Yuan, M.; Dai, Y.; Ning, R.; Zhu, Y.; Tao, K.; et al. A flexible dual-structured mxene for ultra-sensitive and ultra-wide monitoring of anatomical and physiological movements. J. Mater. Chem. A 2021, 9, 26867–26874. [Google Scholar] [CrossRef]
  26. Zhang, S.; Chhetry, A.; Zahed, M.; Sharma, S.; Park, C.; Yoon, S.; Park, J.Y. On-skin ultrathin and stretchable multifunctional sensor for smart healthcare wearables. NPJ Flex. Electron. 2022, 6, 11. [Google Scholar] [CrossRef]
  27. Wang, Z.L. Triboelectric nanogenerators as new energy technology and self-powered sensors-principles, problems and perspectives. Faraday Discuss. 2014, 176, 447–458. [Google Scholar] [CrossRef]
  28. Vollmer, M. Infrared. Eur. J. Phys. 2013, 34, s49–s50. [Google Scholar] [CrossRef] [Green Version]
  29. Chen, J.J.; Xiang, T. Robot soft grabbing with new piezoresistive tactile sensor. Adv. Mater. Res. 2013, 744, 501–504. [Google Scholar] [CrossRef]
  30. Qin, J.; Yin, L.J.; Hao, Y.N.; Zhong, S.L.; Zhang, D.L.; Bi, K.; Zhang, Y.X.; Zhao, Y.; Dang, Z.M. Flexible and stretchable capacitive sensors with different microstructures. Adv. Mater. 2021, 33, e2008267. [Google Scholar] [CrossRef]
  31. Chen, J.; Zheng, J.; Gao, Q.; Zhang, J.; Zhang, J.; Omisore, O.M.; Wang, L.; Li, H. Polydimethylsiloxane (pdms)-based flexible resistive strain sensors for wearable applications. Appl. Sci. 2018, 8, 345. [Google Scholar] [CrossRef] [Green Version]
  32. Dejke, V.; Eng, M.P.; Brinkfeldt, K.; Charnley, J.; Lussey, D.; Lussey, C. Development of prototype low-cost qtss wearable flexible more enviro-friendly pressure, shear, and friction sensors for dynamic prosthetic fit monitoring. Sensors 2021, 21, 3764. [Google Scholar] [CrossRef] [PubMed]
  33. Zhang, H.; Han, W.; Xu, K.; Zhang, Y.; Lu, Y.; Nie, Z.; Du, Y.; Zhu, J.; Huang, W. Metallic sandwiched-aerogel hybrids enabling flexible and stretchable intelligent sensor. Nano Lett. 2020, 20, 3449–3458. [Google Scholar] [CrossRef] [PubMed]
  34. Ammu, S.; Dua, V.; Agnihotra, S.R.; Surwade, S.P.; Phulgirkar, A.; Patel, S.; Manohar, S.K. Flexible, all-organic chemiresistor for detecting chemically aggressive vapors. J. Am. Chem. Soc. 2012, 134, 4553–4556. [Google Scholar] [CrossRef]
  35. Goud, K.Y.; Sandhu, S.S.; Teymourian, H.; Yin, L.; Tostado, N.; Raushel, F.M.; Harvey, S.P.; Moores, L.C.; Wang, J. Textile-based wearable solid-contact flexible fluoride sensor: Toward biodetection of g-type nerve agents. Biosens. Bioelectron. 2021, 182, 113172. [Google Scholar] [CrossRef]
  36. Makushko, P.; Oliveros Mata, E.S.; Cañón Bermúdez, G.S.; Hassan, M.; Laureti, S.; Rinaldi, C.; Fagiani, F.; Barucca, G.; Schmidt, N.; Zabila, Y.; et al. Flexible magnetoreceptor with tunable intrinsic logic for on-skin touchless human-machine interfaces. Adv. Funct. Mater. 2021, 31, 2101089. [Google Scholar] [CrossRef]
  37. Lee, J.; Lim, M.; Yoon, J.; Kim, M.S.; Choi, B.; Kim, D.M.; Kim, D.H.; Park, I.; Choi, S.J. Transparent, flexible strain sensor based on a solution-processed carbon nanotube network. ACS Appl. Mater. Interfaces 2017, 9, 26279–26285. [Google Scholar] [CrossRef]
  38. Ma, Y.; Liu, N.; Li, L.; Hu, X.; Zou, Z.; Wang, J.; Luo, S.; Gao, Y. A highly flexible and sensitive piezoresistive sensor based on mxene with greatly changed interlayer distances. Nat. Commun. 2017, 8, 1207. [Google Scholar] [CrossRef] [Green Version]
  39. Chowdhury, S.A.; Saha, M.C.; Patterson, S.; Robison, T.; Liu, Y. Highly conductive polydimethylsiloxane/carbon nanofiber composites for flexible sensor applications. Adv. Mater. Technol. 2019, 4, 1800398. [Google Scholar] [CrossRef] [Green Version]
  40. Yu, Z.; Xu, J.; Gong, H.; Li, Y.; Li, L.; Wei, Q.; Tang, D. Bioinspired self-powered piezoresistive sensors for simultaneous monitoring of human health and outdoor UV light intensity. ACS Appl. Mater. Interfaces 2022, 14, 5101–5111. [Google Scholar] [CrossRef]
  41. Li, X.P.; Li, Y.; Li, X.; Song, D.; Min, P.; Hu, C.; Zhang, H.B.; Koratkar, N.; Yu, Z.Z. Highly sensitive, reliable and flexible piezoresistive pressure sensors featuring polyurethane sponge coated with mxene sheets. J. Colloid Interface Sci. 2019, 542, 54–62. [Google Scholar] [CrossRef]
  42. Ding, Y.; Yang, J.; Tolle, C.; Zhu, Z. Conductive sponge prepared via one-step dip coating as piezoresistive pressure sensor for human motion detection. ACS Appl. Mater. Interfaces 2018, 10, 16077–16086. [Google Scholar] [CrossRef]
  43. Zhou, Y.; Zhan, P.; Ren, M.; Zheng, G.; Dai, K.; Mi, L.; Shen, C. Significant stretchability enhancement of a crack-based strain sensor combined with high sensitivity and superior durability for motion monitoring. ACS Appl. Mater. Interfaces 2019, 11, 7405–7414. [Google Scholar] [CrossRef]
  44. Yang, Z.; Pang, Y.; Han, X.L.; Yang, Y.; Ling, J.; Jian, M.; Ren, T.L. Graphene textile strain sensor with negative resistance variation for human motion detection. ACS Nano 2018, 12, 9134–9141. [Google Scholar] [CrossRef]
  45. Yang, Z.; Wang, D.Y.; Pang, Y.; Li, Y.X.; Wang, Q.; Zhang, T.Y.; Ren, T.L. Simultaneously detecting subtle and intensive human motions based on a silver nanoparticles bridged graphene strain sensor. ACS Appl. Mater. Interfaces 2018, 10, 3948–3954. [Google Scholar] [CrossRef]
  46. Larimi, S.R.; Nejad, H.R.; Oyatsi, M.; O’Brien, A.; Hoorfar, M.; Najjaran, H. Low-cost ultra-stretchable strain sensors for monitoring human motion and bio-signals. Sens. Actuators A Phys. 2018, 271, 182–191. [Google Scholar] [CrossRef]
  47. Jung, Y.H.; Hong, S.K.; Wang, H.S.; Han, J.H.; Pham, T.X.; Park, H.; Lee, K.J. Flexible piezoelectric acoustic sensors and machine learning for speech processing. Adv. Mater. 2020, 32, e1904020. [Google Scholar] [CrossRef]
  48. Curry, E.J.; Ke, K.; Chorsi, M.T.; Wrobel, K.S.; Miller, A.N., III; Patel, A.; Nguyen, T.D. Biodegradable piezoelectric force sensor. Proc. Natl. Acad. Sci. USA 2018, 115, 909–914. [Google Scholar] [CrossRef] [Green Version]
  49. Chen, X.; Shao, J.; Tian, H.; Li, X.; Tian, Y.; Wang, C. Flexible three-axial tactile sensors with microstructure-enhanced piezoelectric effect and specially-arranged piezoelectric arrays. Smart Mater. Struct. 2018, 27, 025018. [Google Scholar] [CrossRef]
  50. Zhang, Z.; Chen, L.; Yang, X.; Li, T.; Chen, X.; Li, X.; Zhang, J. Enhanced flexible piezoelectric sensor by the integration of P (VDF-TrFE)/AgNWs film with a-IGZO TFT. IEEE Electron. Device Lett. 2018, 40, 111–114. [Google Scholar] [CrossRef]
  51. Kim, K.B.; Jang, W.; Cho, J.Y.; Woo, S.B.; Jeon, D.H.; Ahn, J.H.; Sung, T.H. Transparent and flexible piezoelectric sensor for detecting human movement with a boron nitride nanosheet (BNNS). Nano Energy 2018, 54, 91–98. [Google Scholar] [CrossRef]
  52. Wang, J.; Jiang, J.; Zhang, C.; Sun, M.; Han, S.; Zhang, R.; Liu, H. Energy-efficient, fully flexible, high-performance tactile sensor based on piezotronic effect: Piezoelectric signal amplified with organic field-effect transistors. Nano Energy 2020, 76, 105050. [Google Scholar] [CrossRef]
  53. Yang, Y.; Pan, H.; Xie, G.; Jiang, Y.; Chen, C.; Su, Y.; Tai, H. Flexible piezoelectric pressure sensor based on polydopamine-modified BaTiO3/PVDF composite film for human motion monitoring. Sens. Actuators A Phys. 2020, 301, 111789. [Google Scholar] [CrossRef]
  54. Jiang, J.; Tu, S.; Fu, R.; Li, J.; Hu, F.; Yan, B.; Chen, S. Flexible piezoelectric pressure tactile sensor based on electrospun BaTiO3/polyv (vinylidene fluoride) nanocomposite membrane. ACS Appl. Mater. Interfaces 2020, 12, 33989–33998. [Google Scholar] [CrossRef] [PubMed]
  55. Yu, N.; Mugo, S.M. A flexible-imprinted capacitive sensor for rapid detection of adrenaline. Talanta 2019, 204, 602–606. [Google Scholar]
  56. Golabzaei, S.; Khajavi, R.; Shayanfar, H.A.; Yazdanshenas, M.E.; Talebi, N. Fabrication and characterization of a flexible capacitive sensor on pet fabric. Int. J. Cloth. Sci. Technol. 2018, 30, 687–697. [Google Scholar] [CrossRef]
  57. Zhou, Q.; Ji, B.; Wei, Y.; Hu, B.; Gao, Y.; Xu, Q.; Zhou, B. A bio-inspired cilia array as the dielectric layer for flexible capacitive pressure sensors with high sensitivity and a broad detection range. J. Mater. Chem. A 2019, 7, 27334–27346. [Google Scholar] [CrossRef]
  58. Xiong, Y.; Shen, Y.; Tian, L.; Hu, Y.; Zhu, P.; Sun, R.; Wong, C.P. A flexible, ultra-highly sensitive and stable capacitive pressure sensor with convex microarrays for motion and health monitoring. Nano Energy 2020, 70, 104436. [Google Scholar] [CrossRef]
  59. Li, M.; Liang, J.; Wang, X.; Zhang, M. Ultra-sensitive flexible pressure sensor based on microstructured electrode. Sensors 2020, 20, 371. [Google Scholar] [CrossRef] [Green Version]
  60. Mo, F.; Huang, Y.; Li, Q.; Wang, Z.; Jiang, R.; Gai, W.; Zhi, C. A highly stable and durable capacitive strain sensor based on dynamically super-tough hydro/organo-gels. Adv. Funct. Mater. 2021, 31, 2010830. [Google Scholar] [CrossRef]
  61. Keum, K.; Eom, J.; Lee, J.H.; Heo, J.S.; Park, S.K.; Kim, Y.H. Fully-integrated wearable pressure sensor array enabled by highly sensitive textile-based capacitive ionotronic devices. Nano Energy 2021, 79, 105479. [Google Scholar] [CrossRef]
  62. Wang, Z.L.; Chen, J.; Lin, L. Progress in triboelectric nanogenerators as a new energy technology and self-powered sensors. Energy Environ. Sci. 2015, 8, 2250–2282. [Google Scholar] [CrossRef]
  63. Ning, C.; Cheng, R.; Jiang, Y.; Sheng, F.; Yi, J.; Shen, S.; Zhang, Y.; Peng, X.; Dong, K.; Wang, Z.L. Helical fiber strain sensors based on triboelectric nanogenerators for self-powered human respiratory monitoring. ACS Nano 2022, 16, 2811–2821. [Google Scholar] [CrossRef]
  64. Zhang, D.; Xu, Z.; Yang, Z.; Song, X. High-performance flexible self-powered tin disulfide nanoflowers/reduced graphene oxide nanohybrid-based humidity sensor driven by triboelectric nanogenerator. Nano Energy 2020, 67, 104251. [Google Scholar] [CrossRef]
  65. Zhu, M.; Sun, Z.; Chen, T.; Lee, C. Low cost exoskeleton manipulator using bidirectional triboelectric sensors enhanced multiple degree of freedom sensory system. Nat. Commun. 2021, 12, 2692. [Google Scholar] [CrossRef]
  66. Chen, T.; Shi, Q.; Zhu, M.; He, T.; Sun, L.; Yang, L.; Lee, C. Triboelectric self-powered wearable flexible patch as 3d motion control interface for robotic manipulator. ACS Nano 2018, 12, 11561–11571. [Google Scholar] [CrossRef]
  67. Yan, J.; Ma, Y.; Li, X.; Zhang, C.; Cao, M.; Chen, W.; Luo, S.; Zhu, M.; Gao, Y. Flexible and high-sensitivity piezoresistive sensor based on mxene composite with wrinkle structure. Ceram. Int. 2020, 46, 23592–23598. [Google Scholar] [CrossRef]
  68. Li, S.; Li, R.; González, O.G.; Chen, T.; Xiao, X. Highly sensitive and flexible piezoresistive sensor based on c-mwcnts decorated tpu electrospun fibrous network for human motion detection. Compos. Sci. Technol. 2021, 203, 108617. [Google Scholar] [CrossRef]
  69. Shi, J.; Wang, L.; Dai, Z.; Zhao, L.; Du, M.; Li, H.; Fang, Y. Multiscale hierarchical design of a flexible piezoresistive pressure sensor with high sensitivity and wide linearity range. Small 2018, 14, e1800819. [Google Scholar] [CrossRef]
  70. Zhang, C.; Chen, R.; Xiao, C.; Zhao, H.; Wang, Y.; Geng, D.; Zhou, W. Laser direct writing of highly ordered two-level hierarchical microstructures for flexible piezoresistive sensor with enhanced sensitivity. Adv. Mater. Interfaces 2021, 9, 2101596. [Google Scholar] [CrossRef]
  71. Xu, H.; Gao, L.; Wang, Y.; Cao, K.; Hu, X.; Wang, L.; Lu, Y. Flexible waterproof piezoresistive pressure sensors with wide linear working range based on conductive fabrics. Nanomicro Lett. 2020, 12, 159. [Google Scholar] [CrossRef]
  72. Xu, M.; Cai, H.; Liu, Z.; Chen, F.; Wang, Y.; Dai, F.; Li, Z. Skin-friendly corrugated multilayer microspherical sensor fabricated with silk fibroin, poly (lactic-co-glycolic acid), polyaniline, and kappa-carrageenan for wide range pressure detection. Int. J. Biol. Macromol. 2022, 194, 755–762. [Google Scholar] [CrossRef]
  73. He, Z.; Chen, W.; Liang, B.; Liu, C.; Yang, L.; Lu, D.; Gui, X. Capacitive pressure sensor with high sensitivity and fast response to dynamic interaction based on graphene and porous nylon networks. ACS Appl. Mater. Interfaces 2018, 10, 12816–12823. [Google Scholar] [CrossRef]
  74. Yang, J.; Luo, S.; Zhou, X.; Li, J.; Fu, J.; Yang, W.; Wei, D. Flexible, tunable, and ultrasensitive capacitive pressure sensor with microconformal graphene electrodes. ACS Appl. Mater. Interfaces 2019, 11, 14997–15006. [Google Scholar] [CrossRef]
  75. Wei, P.; Guo, X.; Qiu, X.; Yu, D. Flexible capacitive pressure sensor with sensitivity and linear measuring range enhanced based on porous composite of carbon conductive paste and polydimethylsiloxane. Nanotechnology 2019, 30, 455501. [Google Scholar] [CrossRef]
  76. Qiu, J.; Guo, X.; Chu, R.; Wang, S.; Zeng, W.; Qu, L.; Xing, G. Rapid-response, low detection limit, and high-sensitivity capacitive flexible tactile sensor based on three-dimensional porous dielectric layer for wearable electronic skin. ACS Appl. Mater. Interfaces 2019, 11, 40716–40725. [Google Scholar] [CrossRef]
  77. Ma, L.; Shuai, X.; Hu, Y.; Liang, X.; Zhu, P.; Sun, R.; Wong, C.P. A highly sensitive and flexible capacitive pressure sensor based on a micro-arrayed polydimethylsiloxane dielectric layer. J. Mater. Chem. C 2018, 6, 13232–13240. [Google Scholar] [CrossRef]
  78. Luo, S.; Zhou, X.; Tang, X.; Li, J.; Wei, D.; Tai, G.; Yang, J. Microconformal electrode-dielectric integration for flexible ultrasensitive robotic tactile sensing. Nano Energy 2021, 80, 105580. [Google Scholar] [CrossRef]
  79. Mo, L.; Meng, X.; Zhao, J.; Pan, Y.; Sun, Z.; Guo, Z.; Li, L. Full printed flexible pressure sensor based on microcapsule controllable structure and composite dielectrics. Flex. Print. Electron. 2021, 6, 014001. [Google Scholar] [CrossRef]
  80. Wang, L.; Liu, Y.; Liu, Q.; Zhu, Y.; Wang, H.; Xie, Z.; Zi, Y. A metal-electrode-free, fully integrated, soft triboelectric sensor array for self-powered tactile sensing. Microsyst. Nanoeng. 2020, 6, 59. [Google Scholar] [CrossRef]
  81. Yang, P.; Shi, Y.; Li, S.; Tao, X.; Liu, Z.; Wang, X.; Chen, X. Monitoring the degree of comfort of shoes in-motion using triboelectric pressure sensors with an ultrawide detection range. ACS Nano 2022, 16, 4654–4665. [Google Scholar] [CrossRef]
  82. Yu, J.; Hou, X.; Cui, M.; Shi, S.; He, J.; Sun, Y.; Chou, X. Flexible pdms-based triboelectric nanogenerator for instantaneous force sensing and human joint movement monitoring. Sci. China Mater. 2019, 62, 1423–1432. [Google Scholar] [CrossRef] [Green Version]
  83. Fang, Y.; Zou, Y.; Xu, J.; Chen, G.; Zhou, Y.; Deng, W.; Chen, J. Ambulatory cardiovascular monitoring via a machine-learning-assisted textile triboelectric sensor. Adv. Mater. 2021, 33, e2104178. [Google Scholar] [CrossRef]
  84. Zhao, G.; Zhang, X.; Cui, X.; Wang, S.; Liu, Z.; Deng, L.; Li, L. Piezoelectric polyacrylonitrile nanofiber film-based dual-function self-powered flexible sensor. ACS Appl. Mater. Interfaces 2018, 10, 15855–15863. [Google Scholar] [CrossRef]
  85. Zhang, T.; Wen, Z.; Lei, H.; Gao, Z.; Chen, Y.; Zhang, Y.; Sun, X. Surface-microengineering for high-performance triboelectric tactile sensor via dynamically assembled ferrofluid template. Nano Energy 2021, 87, 106215. [Google Scholar] [CrossRef]
  86. Banstola, B.; Murray, K.K. Pulsed valve matrix-assisted ionization. Analyst 2017, 142, 1672–1675. [Google Scholar] [CrossRef] [PubMed]
  87. Chen, F.; Huang, H.; Guo, L.; Zhang, Y.; Ma, T. The role of polarization in photocatalysis. Angew. Chem. Int. Ed. Engl. 2019, 58, 10061–10073. [Google Scholar] [CrossRef] [PubMed]
  88. Chen, Z.G.; Shi, X.; Zhao, L.D.; Zou, J. High-performance snse thermoelectric materials: Progress and future challenge. Prog. Mater. Sci. 2018, 97, 283–346. [Google Scholar] [CrossRef] [Green Version]
  89. Ouyang, W.; Teng, F.; He, J.H.; Fang, X. Enhancing the photoelectric performance of photodetectors based on metal oxide semiconductors by charge-carrier engineering. Adv. Funct. Mater. 2019, 29, 1807672. [Google Scholar] [CrossRef]
  90. Yan, Y.; Geng, L.D.; Tan, Y.; Ma, J.; Zhang, L.; Sanghadasa, M.; Priya, S. Colossal tunability in high frequency magnetoelectric voltage tunable inductors. Nat. Commun. 2018, 9, 4998. [Google Scholar] [CrossRef]
  91. Pang, Y.; Jian, J.; Tu, T.; Yang, Z.; Ling, J.; Li, Y.; Ren, T.L. Wearable humidity sensor based on porous graphene network for respiration monitoring. Biosens. Bioelectron. 2018, 116, 123–129. [Google Scholar] [CrossRef]
  92. Park, C.H.; Schroeder, V.; Kim, B.J.; Swager, T.M. Ionic liquid-carbon nanotube sensor arrays for human breath related volatile organic compounds. ACS Sens. 2018, 3, 2432–2437. [Google Scholar] [CrossRef]
  93. Falina, S.; Syamsul, M.; Rhaffor, N.A.; Sal Hamid, S.; Mohamed Zain, K.A.; Abd Manaf, A.; Kawarada, H. Ten years progress of electrical detection of heavy metal ions (hmis) using various field-effect transistor (fet) nanosensors: A review. Biosensors 2021, 11, 478. [Google Scholar] [CrossRef]
  94. Tessarolo, M.; Possanzini, L.; Campari, E.G.; Bonfiglioli, R.; Violante, F.S.; Bonfiglio, A.; Fraboni, B. Adaptable pressure textile sensors based on a conductive polymer. Flex. Print. Electron. 2018, 3, 034001. [Google Scholar] [CrossRef]
  95. Khalili, N.; Asif, H.; Naguib, H.E. Towards development of nanofibrous large strain flexible strain sensors with programmable shape memory properties. Smart Mater. Struct. 2018, 27, 055002. [Google Scholar] [CrossRef]
  96. Li, H.; Zhang, J.; Chen, J.; Luo, Z.; Zhang, J.; Alhandarish, Y.; Wang, L. A supersensitive, multidimensional flexible strain gauge sensor based on Ag/PDMS for human activities monitoring. Sci. Rep. 2020, 10, 4639. [Google Scholar] [CrossRef]
  97. Kumar, A.; Hsieh, P.Y.; Shaikh, M.O.; Kumar, R.R.; Chuang, C.H. Flexible temperature sensor utilizing mwcnt doped peg-pu copolymer nanocomposites. Micromachines 2022, 13, 4639. [Google Scholar] [CrossRef]
  98. Sempionatto, J.R.; Mishra, R.K.; Martín, A.; Tang, G.; Nakagawa, T.; Lu, X.; Wang, J. Wearable ring-based sensing platform for detecting chemical threats. ACS Sens. 2017, 2, 1531–1538. [Google Scholar] [CrossRef] [Green Version]
  99. Nakata, S.; Shiomi, M.; Fujita, Y.; Arie, T.; Akita, S.; Takei, K. A wearable ph sensor with high sensitivity based on a flexible charge-coupled device. Nat. Electron. 2018, 1, 596–603. [Google Scholar] [CrossRef]
  100. Sun, S.; Zhang, X.; Song, X.; Liang, S.; Wang, L.; Yang, Z. Bottom-up assembly of hierarchical Cu2O nanospheres: Controllable synthesis, formation mechanism and enhanced photochemical activities. Crystengcomm 2012, 14, 3545–3553. [Google Scholar] [CrossRef]
  101. Chen, M.M.; Cheng, S.B.; Ji, K.; Gao, J.; Liu, Y.L.; Wen, W.; Huang, W.H. Construction of a flexible electrochemiluminescence platform for sweat detection. Chem. Sci. 2019, 10, 6295–6303. [Google Scholar] [CrossRef] [Green Version]
  102. Liu, Y.L.; Liu, R.; Qin, Y.; Qiu, Q.F.; Chen, Z.; Cheng, S.B.; Huang, W.H. Flexible electrochemical urea sensor based on surface molecularly imprinted nanotubes for detection of human sweat. Anal. Chem. 2018, 90, 13081–13087. [Google Scholar] [CrossRef]
  103. Seo, M.H.; Kang, K.; Yoo, J.Y.; Park, J.; Lee, J.S.; Cho, I.; Park, I. Chemo-mechanically operating palladium-polymer nanograting film for a self-powered H2 gas sensor. ACS Nano 2020, 14, 16813–16822. [Google Scholar] [CrossRef]
  104. Shin, H.; Kim, D.H.; Jung, W.; Jang, J.S.; Kim, Y.H.; Lee, Y.; Kim, I.D. Surface activity-tuned metal oxide chemiresistor: Toward direct and quantitative halitosis diagnosis. ACS Nano 2021, 15, 14207–14217. [Google Scholar] [CrossRef]
  105. Bezdek, M.J.; Luo SX, L.; Ku, K.H.; Swager, T.M. A chemiresistive methane sensor. Proc. Natl. Acad. Sci. USA 2021, 118, e2022515118. [Google Scholar] [CrossRef]
  106. Di, X.; Ma, Q.; Xu, Y.; Yang, M.; Wu, G.; Sun, P. High-performance ionic conductive poly (vinyl alcohol) hydrogels for flexible strain sensors based on a universal soaking strategy. Mater. Chem. Front. 2021, 5, 315–323. [Google Scholar] [CrossRef]
  107. Xia, S.; Song, S.; Gao, G. Robust and flexible strain sensors based on dual physically cross-linked double network hydrogels for monitoring human-motion. Chem. Eng. J. 2018, 354, 817–824. [Google Scholar] [CrossRef]
  108. Sun, J.; Lu, G.; Zhou, J.; Yuan, Y.; Zhu, X.; Nie, J. Robust physically linked double-network ionogel as a flexible bimodal sensor. ACS Appl. Mater. Interfaces 2020, 12, 14272–14279. [Google Scholar] [CrossRef]
  109. Li, F.; Wang, R.; Song, C.; Zhao, M.; Ren, H.; Wang, S.; Hao, Y. A skin-inspired artificial mechanoreceptor for tactile enhancement and integration. ACS Nano 2021, 15, 16422–16431. [Google Scholar] [CrossRef]
  110. Su, Q.; Zou, Q.; Li, Y.; Chen, Y.; Teng, S.Y.; Kelleher, J.T.; Wang, S. A stretchable and strain-unperturbed pressure sensor for motion interference–free tactile monitoring on skins. Sci. Adv. 2021, 7, eabi4563. [Google Scholar] [CrossRef]
  111. Bell, C.; Nammari, A.; Uttamchandani, P.; Rai, A.; Shah, P.; Moore, A.L. Flexible electronics-compatible non-enzymatic glucose sensing via transparent cuo nanowire networks on pet films. Nanotechnology 2017, 28, 245502. [Google Scholar] [CrossRef]
  112. Lim, J.M.; Ryu, M.Y.; Yun, J.W.; Park, T.J.; Park, J.P. Electrochemical peptide sensor for diagnosing adenoma-carcinoma transition in colon cancer. Biosens. Bioelectron. 2017, 98, 330–337. [Google Scholar] [CrossRef]
  113. Thakur, N.; Kumar, M.; Adhikary, S.D.; Mandal, D.; Nagaiah, T.C. Pvim-Co5POM/MNC composite as a flexible electrode for the ultrasensitive and highly selective non-enzymatic electrochemical detection of cholesterol. Chem. Commun. 2019, 55, 5021–5024. [Google Scholar] [CrossRef] [PubMed]
  114. Yoon, J.; Sim, M.; Oh, T.S.; Yoon, Y.S.; Kim, D.J. Flexible electrochemical sensor based on NiCu(OOH) for monitoring urea in human sweat. J. Electrochem. Soc. 2021, 168, 117510. [Google Scholar] [CrossRef]
  115. Li, R.; Qi, H.; Ma, Y.; Deng, Y.; Liu, S.; Jie, Y.; Yin, L. A flexible and physically transient electrochemical sensor for real-time wireless nitric oxide monitoring. Nat. Commun. 2020, 11, 3207. [Google Scholar] [CrossRef] [PubMed]
  116. Reid, D.O.; Smith, R.E.; Garcia-Torres, J.; Watts, J.F.; Crean, C. Solvent treatment of wet-spun pedot: Pss fibers for fiber-based wearable pH sensing. Sensors 2019, 19, 4213. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Naim, A.M.; Wickramasinghe, K.; De Silva, A.; Perera, M.V.; Lalitharatne, T.D.; Kappel, S.L. Low-cost active dry-contact surface emg sensor for bionic arms. In Proceedings of the 2020 IEEE International Conference on Systems, Man, and Cybernetics (SMC), Toronto, ON, Canada, 11–14 October 2020; pp. 3327–3332. [Google Scholar]
  118. Yao, Y.; Rakheja, S.; Marcotte, P. Relationship among hand forces imparted on a viscoelastic hand-handle interface. Measurement 2019, 145, 525–534. [Google Scholar] [CrossRef]
  119. Peng, Y.; Que, M.; Lee, H.E.; Bao, R.; Wang, X.; Lu, J.; Pan, C. Achieving high-resolution pressure mapping via flexible GaN/ZnO nanowire LEDs array by piezo-phototronic effect. Nano Energy 2019, 58, 633–640. [Google Scholar] [CrossRef]
  120. Chen, W.; Feng, F.; Chen, D.; Lin, W.; Chen, S.C. Precision non-contact displacement sensor based on the near-field characteristics of fiber specklegrams. Sens. Actuators A Phys. 2019, 296, 1–6. [Google Scholar] [CrossRef]
  121. Guo, H.; Tan, Y.J.; Chen, G.; Wang, Z.; Susanto, G.J.; See, H.H.; Tee, B.C. Artificially innervated self-healing foams as synthetic piezo-impedance sensor skins. Nat. Commun. 2020, 11, 5747. [Google Scholar] [CrossRef]
  122. Zhang, W.; Wu, Z.; Hu, J.; Cao, Y.; Guo, J.; Long, M.; Jia, D. Flexible chemiresistive sensor of polyaniline coated filter paper prepared by spraying for fast and non-contact detection of nitroaromatic explosives. Sens. Actuators B Chem. 2020, 304, 127233. [Google Scholar] [CrossRef]
  123. Tang, Y.; Zhou, H.; Sun, X.; Diao, N.; Wang, J.; Zhang, B.; Mao, Y. Triboelectric touch-free screen sensor for noncontact gesture recognizing. Adv. Funct. Mater. 2019, 30, 1907891. [Google Scholar] [CrossRef]
  124. Lu, L.; Jiang, C.; Hu, G.; Liu, J.; Yang, B. Flexible noncontact sensing for human-machine interaction. Adv. Mater. 2021, 33, e2100218. [Google Scholar] [CrossRef]
  125. Tang, H.; Li, Y.; Ye, H.; Hu, F.; Gao, C.; Tao, L.; Zhang, G. High-performance humidity sensor using schottky-contacted sns nanoflakes for noncontact healthcare monitoring. Nanotechnology 2020, 31, 055501. [Google Scholar] [CrossRef]
  126. Viola, F.A.; Spanu, A.; Ricci, P.C.; Bonfiglio, A.; Cosseddu, P. Ultrathin, flexible and multimodal tactile sensors based on organic field-effect transistors. Sci. Rep. 2018, 8, 8073. [Google Scholar] [CrossRef] [Green Version]
  127. Jia, W.; Zhang, Q.; Cheng, Y.; Zhao, D.; Liu, Y.; Zhang, W.; Sang, S. Flexible and highly sensitive piezoresistive pressure sensor with sandpaper as a mold. Nano 2019, 14, 1950081. [Google Scholar] [CrossRef]
  128. Zhou, Q.; Ji, B.; Hu, B.; Li, S.; Xu, Y.; Gao, Y.; Zhou, B. Tilted magnetic micropillars enabled dual-mode sensor for tactile/touchless perceptions. Nano Energy 2020, 78, 105382. [Google Scholar] [CrossRef]
  129. Zhao, P.; Zhang, R.; Tong, Y.; Zhao, X.; Zhang, T.; Tang, Q.; Liu, Y. Strain-discriminable pressure/proximity sensing of transparent stretchable electronic skin based on pedot:pss/swcnt electrodes. ACS Appl. Mater. Interfaces 2020, 12, 55083–55093. [Google Scholar] [CrossRef]
  130. Xu, J.; Pei, L.; Li, J.; Pang, H.; Li, Z.; Li, B.; Gong, X. Flexible, self-powered, magnetism/pressure dual-mode sensor based on magnetorheological plastomer. Compos. Sci. Technol. 2019, 183, 107820. [Google Scholar] [CrossRef]
  131. Guan, F.; Xie, Y.; Wu, H.; Meng, Y.; Shi, Y.; Gao, M.; Pei, Q. Silver nanowire-bacterial cellulose composite fiber-based sensor for highly sensitive detection of pressure and proximity. ACS Nano 2020, 14, 15428–15439. [Google Scholar] [CrossRef]
  132. Ge, J.; Wang, X.; Drack, M.; Volkov, O.; Liang, M.; Cañón Bermúdez, G.S.; Makarov, D. A bimodal soft electronic skin for tactile and touchless interaction in real time. Nat. Commun. 2019, 10, 4405. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Chen, Q.; Liu, D.; Lin, L.; Wu, J. Bridging interdigitated electrodes by electrochemical-assisted deposition of graphene oxide for constructing flexible gas sensor. Sens. Actuators B Chem. 2019, 286, 591–599. [Google Scholar] [CrossRef]
  134. Shanglin, Y.; Rong, F. Flexible-substrate fluxgate current sensor based on mems technology. Sens. Mater. 2020, 32, 3083–3094. [Google Scholar] [CrossRef]
  135. Jia, Z.; Li, Z.; Ma, S.; Zhang, W.; Chen, Y.; Luo, Y.; Kong, L. Constructing conductive titanium carbide nanosheet (mxene) network on polyurethane/polyacrylonitrile fibre framework for flexible strain sensor. J. Colloid Interface Sci. 2021, 584, 1–10. [Google Scholar] [CrossRef]
  136. Li, S.; Wang, T.; Yang, Z.; He, J.; Wang, J.; Zhao, L.; Lu, G. Room temperature high performance nh3 sensor based on go-rambutan-like polyaniline hollow nanosphere hybrid assembled to flexible pet substrate. Sens. Actuators B Chem. 2018, 273, 726–734. [Google Scholar] [CrossRef]
  137. Lin, S.; Feng, W.; Miao, X.; Zhang, X.; Chen, S.; Chen, Y.; Zhang, Y. A flexible and highly sensitive nonenzymatic glucose sensor based on dvd-laser scribed graphene substrate. Biosens. Bioelectron. 2018, 110, 89–96. [Google Scholar] [CrossRef]
  138. Song, Y.; Chen, H.; Su, Z.; Chen, X.; Miao, L.; Zhang, J.; Zhang, H. Highly compressible integrated supercapacitor-piezoresistance-sensor system with cnt-pdms sponge for health monitoring. Small 2017, 13, 1702091. [Google Scholar] [CrossRef]
  139. Wu, S.; Zhang, J.; Ladani, R.B.; Ravindran, A.R.; Mouritz, A.P.; Kinloch, A.J.; Wang, C.H. Novel electrically conductive porous pdms/carbon nanofiber composites for deformable strain sensors and conductors. ACS Appl. Mater. Interfaces 2017, 9, 14207–14215. [Google Scholar] [CrossRef]
  140. Al-Shedivat, M.; Naous, R.; Cauwenberghs, G.; Salama, K.N. Memristors empower spiking neurons with stochasticity. IEEE J. Emerg. Sel. Top. Circuits Syst. 2015, 5, 242–253. [Google Scholar] [CrossRef]
  141. Liu, J.; Ma, H.; Yang, Y.; Yang, W.; Jiao, Z.; Yu, Y. Study on direct writing of gallium metal for the flexible sensor. Adv. Mater. Sci. Eng. 2021, 2021, 9980534. [Google Scholar] [CrossRef]
  142. Chen, R. Liquid metal based flexible pressure sensor for tactile sensing of robots. J. Phys. Conf. Ser. 2021, 1885, 052025. [Google Scholar] [CrossRef]
  143. Almog, R.; Shashar, E.; Sverdlov, Y.; Shacham-Diamand, Y. Decorating metal oxide nanostructures with noble metal np for bio-sensing applications. In ECS Meeting Abstracts; IOP Publishing: Bristol, UK, 2021. [Google Scholar]
  144. Yin, R.; Yang, S.; Li, Q.; Zhang, S.; Liu, H.; Han, J.; Liu, C.; Shen, C. Flexible conductive ag nanowire/cellulose nanofibril hybrid nanopaper for strain and temperature sensing applications. Sci. Bull. 2020, 65, 899–908. [Google Scholar] [CrossRef]
  145. Shi, R.; Lou, Z.; Chen, S.; Shen, G. Flexible and transparent capacitive pressure sensor with patterned microstructured composite rubber dielectric for wearable touch keyboard application. Sci. China Mater. 2018, 61, 1587–1595. [Google Scholar] [CrossRef] [Green Version]
  146. Deng, W.; Zhang, X.; Jia, R.; Huang, L.; Zhang, X.; Jie, J. Organic molecular crystal-based photosynaptic devices for an artificial visual-perception system. NPG Asia Mater. 2019, 11, 77. [Google Scholar] [CrossRef] [Green Version]
  147. Wang, Y.; Sun, L.; Wang, C.; Yang, F.; Ren, X.; Zhang, X.; Dong, H.; Hu, W. Organic crystalline materials in flexible electronics. Chem. Soc. Rev. 2019, 48, 1492–1530. [Google Scholar] [CrossRef]
  148. Lai, S.; Viola, F.A.; Cosseddu, P.; Bonfiglio, A. Floating gate, organic field-effect transistor-based sensors towards biomedical applications fabricated with large-area processes over flexible substrates. Sensors 2018, 18, 688. [Google Scholar] [CrossRef]
  149. Matsui, H.; Takeda, Y.; Tokito, S. Flexible and printed organic transistors: From materials to integrated circuits. Org. Electron. 2019, 75, 105432. [Google Scholar] [CrossRef]
  150. Zhang, L.; Guo, Y.Y.H.; Liu, G.; Tan, Q. Na-doped zno and rgo composite-based flexible acetone gas sensor operated in room temperature. IEEE Access 2020, 8, 171568–171574. [Google Scholar] [CrossRef]
  151. Shirley, J.A.; Florence, S.E.; Sreeja, B.S.; Padmalaya, G.; Radha, S. Zinc oxide nanostructure-based textile pressure sensor for wearable applications. J. Mater. Sci. Mater. Electron. 2020, 31, 16519–16530. [Google Scholar] [CrossRef]
  152. Sholehah, A.; Karmala, K.; Huda, N.; Utari, L.; Septiani, N.L.; Yuliarto, B. Structural effect of ZnO-Ag chemoresistive sensor on flexible substrate for ethylene gas detection. Sens. Actuators A Phys. 2021, 331, 112934. [Google Scholar] [CrossRef]
  153. Van der Spek, D.; Van Arendonk, J.A.; Bovenhuis, H. Genome-wide association study for claw disorders and trimming status in dairy cattle. J. Dairy Sci. 2015, 98, 1286–1295. [Google Scholar] [CrossRef] [Green Version]
  154. Chen, X.; Wang, T.; Han, Y.; Lv, W.; Li, B.; Su, C.; Zeng, M.; Yang, J.; Hu, N.; Su, Y.; et al. Wearable NO2 sensing and wireless application based on zns nanoparticles/nitrogen-doped reduced graphene oxide. Sens. Actuators B Chem. 2021, 345, 130423. [Google Scholar] [CrossRef]
  155. Mamleyev, E.R.; Heissler, S.; Nefedov, A.; Weidler, P.G.; Nordin, N.; Kudryashov, V.V.; Länge, K.; MacKinnon, N.; Sharma, S. Laser-induced hierarchical carbon patterns on polyimide substrates for flexible urea sensors. NPJ Flex. Electron. 2019, 3, 2. [Google Scholar] [CrossRef]
  156. Qi, Y.; Hu, Y.; Wu, X.; Wu, W.; Bao, J.; Yang, H.; Zhao, J.; Hou, C.; Huo, D. Self-supporting flexible enzyme-free sensor based on CoS-PPy-CP for glucose detection. J. Electrochem. Soc. 2021, 168, 10507. [Google Scholar] [CrossRef]
  157. Patil, A.B.; Meng, Z.; Wu, R.; Ma, L.; Xu, Z.; Shi, C.; Qiu, W.; Liu, Q.; Zhang, Y.; Lin, Y.; et al. Tailoring the meso-structure of gold nanoparticles in keratin-based activated carbon toward high-performance flexible sensor. Nanomicro Lett. 2020, 12, 117. [Google Scholar] [CrossRef]
  158. Zhang, Q.; An, C.; Fan, S.; Shi, S.; Zhang, R.; Zhang, J.; Li, Q.; Zhang, D.; Hu, X.; Liu, J. Flexible gas sensor based on graphene/ethyl cellulose nanocomposite with ultra-low strain response for volatile organic compounds rapid detection. Nanotechnology 2018, 29, 285501. [Google Scholar] [CrossRef]
  159. Chuang, L.; Wei-Wei, L.; Li, C.; Dan, X.; Bao-Jun, L.; Lan, X.; Xiao-Kuo, Y.; Dan-Na, D.; Jia-Hao, L.; Ya-Bo, C. Flexible nitrogen dioxide gas sensor based on reduced graphene oxide sensing material using silver nanowire electrode. Acta Phys. Sin. 2020, 69, 20191390. [Google Scholar]
  160. Li, Y.; Jia, W.; Hou, X.; Zhang, L.; He, J.; Mu, J.; Wang, C.; Yu, J.; Bi, K.; Cui, M.; et al. Enhanced-sensitivity and highly flexible stress/strain sensor based on pzt nanowires-modified graphene with wide range carrier mobility. J. Mater. Sci. Mater. Electron. 2020, 31, 8436–8445. [Google Scholar] [CrossRef]
  161. Liang, B.; Chen, W.; He, Z.; Yang, R.; Lin, Z.; Du, H.; Shang, Y.; Cao, A.; Tang, Z.; Gui, X. Highly sensitive, flexible mems based pressure sensor with photoresist insulation layer. Small 2017, 13, 1702422. [Google Scholar] [CrossRef]
  162. Hao, B.; Mu, L.; Ma, Q.; Yang, S.; Ma, P.C. Stretchable and compressible strain sensor based on carbon nanotube foam/polymer nanocomposites with three-dimensional networks. Compos. Sci. Technol. 2018, 163, 162–170. [Google Scholar] [CrossRef]
  163. Huang, W.; Dai, K.; Zhai, Y.; Liu, H.; Zhan, P.; Gao, J.; Zheng, G.; Liu, C.; Shen, C. Flexible and lightweight pressure sensor based on carbon nanotube/thermoplastic polyurethane-aligned conductive foam with superior compressibility and stability. ACS Appl. Mater. Interfaces 2017, 9, 42266–42277. [Google Scholar] [CrossRef]
  164. Zhou, T.; Zhang, T. Recent progress of nanostructured sensing materials from 0D to 3D: Overview of structure-property-application relationship for gas sensors. Small Methods 2021, 5, e2100515. [Google Scholar] [CrossRef] [PubMed]
  165. 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]
  166. Yang, T.; Xiang, H.; Qin, C.; Liu, Y.; Zhao, X.; Liu, H.; Li, H.; Ouzounian, M.; Hong, G.; Chen, H.; et al. Highly sensitive 1T-MoS2 pressure sensor with wide linearity based on hierarchical microstructures of leaf vein as spacer. Adv. Electron. Mater. 2020, 6, 1900916. [Google Scholar] [CrossRef]
  167. Yue, Y.; Liu, N.; Liu, W.; Li, M.; Ma, Y.; Luo, C.; Wang, S.; Rao, J.; Hu, X.; Su, J.; et al. 3D hybrid porous mxene-sponge network and its application in piezoresistive sensor. Nano Energy 2018, 50, 79–87. [Google Scholar] [CrossRef]
  168. Chen, Z.; Hu, Y.; Zhuo, H.; Liu, L.; Jing, S.; Zhong, L.; Peng, X.; Sun, R.C. Compressible, elastic, and pressure-sensitive carbon aerogels derived from 2d titanium carbide nanosheets and bacterial cellulose for wearable sensors. Chem. Mater. 2019, 31, 3301–3312. [Google Scholar] [CrossRef]
  169. Yang, K.; Yin, F.; Xia, D.; Peng, H.; Yang, J.; Yuan, W. A highly flexible and multifunctional strain sensor based on a network-structured mxene/polyurethane mat with ultra-high sensitivity and a broad sensing range. Nanoscale 2019, 11, 9949–9957. [Google Scholar] [CrossRef] [PubMed]
  170. Dong, K.; Wu, Z.; Deng, J.; Wang, A.C.; Zou, H.; Chen, C.; Hu, D.; Gu, B.; Sun, B.; Wang, Z.L. A stretchable yarn embedded triboelectric nanogenerator as electronic skin for biomechanical energy harvesting and multifunctional pressure sensing. Adv. Mater. 2018, 30, e1804944. [Google Scholar] [CrossRef]
  171. Zeng, X.; Wang, Z.; Zhang, H.; Yang, W.; Xiang, L.; Zhao, Z.; Peng, L.M.; Hu, Y. Tunable, ultrasensitive, and flexible pressure sensors based on wrinkled microstructures for electronic skins. ACS Appl. Mater. Interfaces 2019, 11, 21218–21226. [Google Scholar] [CrossRef]
  172. Lee, Y.; Myoung, J.; Cho, S.; Park, J.; Kim, J.; Lee, H.; Lee, Y.; Lee, S.; Baig, C.; Ko, H. Bioinspired gradient conductivity and stiffness for ultrasensitive electronic skins. ACS Nano 2021, 15, 1795–1804. [Google Scholar] [CrossRef]
  173. Kano, S.; Kim, K.; Fujii, M. Fast-response and flexible nanocrystal-based humidity sensor for monitoring human respiration and water evaporation on skin. ACS Sens. 2017, 2, 828–833. [Google Scholar] [CrossRef]
  174. Sun, X.; Qin, Z.; Ye, L.; Zhang, H.; Yu, Q.; Wu, X.; Li, J.; Yao, F. Carbon nanotubes reinforced hydrogel as flexible strain sensor with high stretchability and mechanically toughness. Chem. Eng. J. 2020, 382, 12282. [Google Scholar] [CrossRef]
  175. Xu, H.; Xiang, J.X.; Lu, Y.F.; Zhang, M.K.; Li, J.J.; Gao, B.B.; Zhao, Y.J.; Gu, Z.Z. Multifunctional wearable sensing devices based on functionalized graphene films for simultaneous monitoring of physiological signals and volatile organic compound biomarkers. ACS Appl. Mater. Interfaces 2018, 10, 11785–11793. [Google Scholar] [CrossRef]
  176. Wang, Y.; Hao, J.; Huang, Z.; Zheng, G.; Dai, K.; Liu, C.; Shen, C. Flexible electrically resistive-type strain sensors based on reduced graphene oxide-decorated electrospun polymer fibrous mats for human motion monitoring. Carbon 2018, 126, 360–371. [Google Scholar] [CrossRef]
  177. Yang, H.; Gong, L.H.; Zheng, Z.; Yao, X.F. Highly stretchable and sensitive conductive rubber composites with tunable piezoresistivity for motion detection and flexible electrodes. Carbon 2020, 158, 893–903. [Google Scholar] [CrossRef]
  178. Ergen, O.; Celik, E.; Unal, A.H.; Erdolu, M.Y.; Sarac, F.E.; Unal, U. Real time chemical and mechanical human motion monitoring with aerogel-based wearable sensors. Lab Chip 2020, 20, 2689–2695. [Google Scholar] [CrossRef] [PubMed]
  179. Pu, J.H.; Zhao, X.; Zha, X.J.; Li, W.D.; Ke, K.; Bao, R.Y.; Liu, Z.Y.; Yang, M.B.; Yang, W. A strain localization directed crack control strategy for designing mxene-based customizable sensitivity and sensing range strain sensors for full-range human motion monitoring. Nano Energy 2020, 74, 104814. [Google Scholar] [CrossRef]
  180. Ma, J.; Wang, P.; Chen, H.; Bao, S.; Chen, W.; Lu, H. Highly sensitive and large-range strain sensor with a self-compensated two-order structure for human motion detection. ACS Appl. Mater. Interfaces 2019, 11, 8527–8536. [Google Scholar] [CrossRef]
  181. Anwar, S.; Prasad, R.; Chowdhary, B.S.; Anjum, M.R. A telemedicine platform for disaster management and emergency care. Wirel. Pers. Commun. 2019, 106, 191–204. [Google Scholar] [CrossRef]
  182. Xu, X.; Zhang, L.; Li, J.; Liu, R.; Lou, L.; Zhao, S.; Chen, H.; Wang, Q.; Wu, X.; Ji, R.; et al. Improved research of adjustable external fixation device for lower limbs based on semiconductor refrigeration sheet. Chin. J. Med. Instrum. 2020, 44, 499–502. [Google Scholar]
  183. Liu, F.; Chen, Z.; Wang, J. Intelligent medical iot system based on wsn with computer vision platforms. Concurr. Comput. Pract. Exp. 2018, 33, e5036. [Google Scholar] [CrossRef]
  184. Feng, R.; Mu, Y.; Zeng, X.; Jia, W.; Liu, Y.; Jiang, X.; Gong, Q.; Hu, Y. A flexible integrated bending strain and pressure sensor system for motion monitoring. Sensors 2021, 21, 3969. [Google Scholar] [CrossRef]
  185. Yu, D.; Zheng, Z.; Liu, J.; Xiao, H.; Huangfu, G.; Guo, Y. Superflexible and lead-free piezoelectric nanogenerator as a highly sensitive self-powered sensor for human motion monitoring. Nanomicro Lett. 2021, 13, 117. [Google Scholar] [CrossRef]
  186. Wan, H.; Zhao, J.; Lo, L.W.; Cao, Y.; Sepúlveda, N.; Wang, C. Multimodal artificial neurological sensory-memory system based on flexible carbon nanotube synaptic transistor. ACS Nano. 2021, 15, 14587–14597. [Google Scholar] [CrossRef]
  187. Zhao, J.; Guo, H.; Pang, Y.K.; Xi, F.; Yang, Z.W.; Liu, G.; Guo, T.; Dong, G.; Zhang, C.; Wang, Z.L. Flexible organic tribotronic transistor for pressure and magnetic sensing. ACS Nano 2017, 11, 11566–11573. [Google Scholar] [CrossRef] [PubMed]
  188. Liu, X.; Ren, Z.; Liu, F.; Zhao, L.; Ling, Q.; Gu, H. Multifunctional self-healing dual network hydrogels constructed via host-guest interaction and dynamic covalent bond as wearable strain sensors for monitoring human and organ motions. ACS Appl. Mater. Interfaces 2021, 13, 14612–14622. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Timeline of the major milestones in the development of flexible sensors in the smart wearable field.
Figure 1. Timeline of the major milestones in the development of flexible sensors in the smart wearable field.
Sensors 22 05089 g001
Figure 2. Flexible sensors based on different sensing mechanisms, detected objects, and contact forms: (a) a resistive strain sensor attached to the skin [31]; (b) a wearable triboelectric sensor for a dynamic prosthetic fit [32]; (c) a highly stretchable, tough, and sensitive strain sensor [33]; (d) organic chemical sensors for the detection of chemically corrosive vapors [34]; (e) contact sensors for bio-detection of flexographically printed fabrics made of fluorine-containing G-type nerve agents [35]; and (f) a non-contact sensor for a physiological monitoring system based on the wearer’s palm [36].
Figure 2. Flexible sensors based on different sensing mechanisms, detected objects, and contact forms: (a) a resistive strain sensor attached to the skin [31]; (b) a wearable triboelectric sensor for a dynamic prosthetic fit [32]; (c) a highly stretchable, tough, and sensitive strain sensor [33]; (d) organic chemical sensors for the detection of chemically corrosive vapors [34]; (e) contact sensors for bio-detection of flexographically printed fabrics made of fluorine-containing G-type nerve agents [35]; and (f) a non-contact sensor for a physiological monitoring system based on the wearer’s palm [36].
Sensors 22 05089 g002
Figure 3. Diagram of the signal conversion mechanism of a resistive flexible sensor.
Figure 3. Diagram of the signal conversion mechanism of a resistive flexible sensor.
Sensors 22 05089 g003
Figure 4. Application of the strain-based flexible sensor in smart wearables: (a) a schematic diagram of electron transport at the crack edge in a crack-flexible strain sensor [43]; (b) a wearable graphene fabric strain sensor for the detection of various human motions [44]; (c) injection of graphene nano-powders into transparent adhesive pads to fabricate flexible strain sensors with an excellent tensile–strain response [45]; and (d) the sensory performance in response to the bending of the finger and the degree thereof are both reflected in the relative rate of change in the resistance [46].
Figure 4. Application of the strain-based flexible sensor in smart wearables: (a) a schematic diagram of electron transport at the crack edge in a crack-flexible strain sensor [43]; (b) a wearable graphene fabric strain sensor for the detection of various human motions [44]; (c) injection of graphene nano-powders into transparent adhesive pads to fabricate flexible strain sensors with an excellent tensile–strain response [45]; and (d) the sensory performance in response to the bending of the finger and the degree thereof are both reflected in the relative rate of change in the resistance [46].
Sensors 22 05089 g004
Figure 5. Diagram of the signal conversion mechanism of a piezoelectric flexible sensor.
Figure 5. Diagram of the signal conversion mechanism of a piezoelectric flexible sensor.
Sensors 22 05089 g005
Figure 6. Applications of strain-based flexible sensors in smart wearable sensing: (a) a schematic diagram of flexible piezoresistive sensors adapted for human motion, including two layers of ITO electrodes, BNNS/PDMS piezoelectric active layer, and TFPS on the hand device [51]; (b) the structure of a high-performance strain-based sensor and a flexible transistor based on the piezoelectric effect, where PET is the flexible substrate and PMMA is the dielectric layer [52]; (c) the structure of a high-performance flexible piezoelectric sensor and transistor based on the piezoelectric effect, where PET is the flexible substrate and PMMA is the dielectric layer [53]; and (d) the mechanism of electrospinning BaTiO3/PVDF composite fibers and the open circuit voltage response to the different sounds of the device vibrating from the speaker [54].
Figure 6. Applications of strain-based flexible sensors in smart wearable sensing: (a) a schematic diagram of flexible piezoresistive sensors adapted for human motion, including two layers of ITO electrodes, BNNS/PDMS piezoelectric active layer, and TFPS on the hand device [51]; (b) the structure of a high-performance strain-based sensor and a flexible transistor based on the piezoelectric effect, where PET is the flexible substrate and PMMA is the dielectric layer [52]; (c) the structure of a high-performance flexible piezoelectric sensor and transistor based on the piezoelectric effect, where PET is the flexible substrate and PMMA is the dielectric layer [53]; and (d) the mechanism of electrospinning BaTiO3/PVDF composite fibers and the open circuit voltage response to the different sounds of the device vibrating from the speaker [54].
Sensors 22 05089 g006
Figure 7. Diagram of the capacitive flexible sensor signal conversion mechanism.
Figure 7. Diagram of the capacitive flexible sensor signal conversion mechanism.
Sensors 22 05089 g007
Figure 8. Diagram of the triboelectric flexible sensor signal conversion mechanism.
Figure 8. Diagram of the triboelectric flexible sensor signal conversion mechanism.
Sensors 22 05089 g008
Figure 9. Application of a flexible sensor based on a chemical formula: (a) schematic diagram of a self-powered H2 gas sensor [103]; (b) surface-active tuned metal oxide chemical resistor for direct analysis of human mouth odor [104].
Figure 9. Application of a flexible sensor based on a chemical formula: (a) schematic diagram of a self-powered H2 gas sensor [103]; (b) surface-active tuned metal oxide chemical resistor for direct analysis of human mouth odor [104].
Sensors 22 05089 g009
Figure 10. Application of flexible sensors on e-skin: (a) a schematic diagram of a flexible pressure sensor based on a wrinkled microstructure with relative capacitance change as a function of the pressure, mounted on a human throat for vocal recognition tests [171]; (b) a multilayer PEDOT:PSS/PUD located on the top of a micro-dome pattern PDMS layer with different layers of pressure-related relative current changes in a multilayered electronic skin [172].
Figure 10. Application of flexible sensors on e-skin: (a) a schematic diagram of a flexible pressure sensor based on a wrinkled microstructure with relative capacitance change as a function of the pressure, mounted on a human throat for vocal recognition tests [171]; (b) a multilayer PEDOT:PSS/PUD located on the top of a micro-dome pattern PDMS layer with different layers of pressure-related relative current changes in a multilayered electronic skin [172].
Sensors 22 05089 g010
Figure 11. Flexible sensors for health monitoring: (a) a schematic of bent forked finger electrodes and a humidity sensor on PI film [173]; (b) a schematic of PAAm-oxCNTs hydrogel and images under stretching as an application for monitoring subtle human motion [174]; (c) a graphene film flexible sensor with different functions for simultaneous monitoring of physiological signals and biomarkers [175]; and (d) a schematic of a low-cost nylon mesh pressure sensor with micro-sized square holes and pressure sensing sensitivity curves for three pressure sensors, each with different mesh sizes of nylon [73].
Figure 11. Flexible sensors for health monitoring: (a) a schematic of bent forked finger electrodes and a humidity sensor on PI film [173]; (b) a schematic of PAAm-oxCNTs hydrogel and images under stretching as an application for monitoring subtle human motion [174]; (c) a graphene film flexible sensor with different functions for simultaneous monitoring of physiological signals and biomarkers [175]; and (d) a schematic of a low-cost nylon mesh pressure sensor with micro-sized square holes and pressure sensing sensitivity curves for three pressure sensors, each with different mesh sizes of nylon [73].
Sensors 22 05089 g011
Figure 12. Application of flexible sensors for motion monitoring: (a) electron microscopy images of the surface and cross-section of the strain sensor and response curves on the wrist at different curvatures [176]; (b) self-compensating second-order structure high-sensitivity and large range strain sensor for human motion detection and its sensing position and corresponding overview [180].
Figure 12. Application of flexible sensors for motion monitoring: (a) electron microscopy images of the surface and cross-section of the strain sensor and response curves on the wrist at different curvatures [176]; (b) self-compensating second-order structure high-sensitivity and large range strain sensor for human motion detection and its sensing position and corresponding overview [180].
Sensors 22 05089 g012
Table 1. Comparative study of flexible sensors with different sensing mechanisms.
Table 1. Comparative study of flexible sensors with different sensing mechanisms.
ClassificationMaterialPreparation MethodLinear Range KPaSensitivity
KPa−1
Cyclic Stability (Times)References
No.
Flexible piezoresistive sensorMXene, PolyacrylateCoating Techniques0–16148.2613,000[67]
TPU, C-MWCNTSElectrospinning0–1021000[68]
PDMS, GrapheneSpraying0–251.21000[69]
PDMS, C-MWCNTSLaser processing0.2–3011.061000[70]
PDMS, IPALaser engraving20–8006.4171000[71]
SiO2, PANI, PLGAElectrospinning10–3800.0712000[72]
Flexible capacitive sensorGraphene, PDMSLaser processing0–6000.331000[73]
Graphene, PDMSChemical vapor deposition1–103.19500[74]
PDMS, CCP, NaClForming and dissolving0–101.111,000[75]
PS, GNP, MWCNTsDip coating0–50.0622000[76]
PDMS, AgNWs, PETSpin coating0–22.041000[77]
GNWs, PDMS, ZnOSpin coating0–2222.32000[78]
AgNWs, PDMSScreen printing0–102.13600[79]
Flexible triboelectric sensorPVA, PEITemplate assist5–500.0632000[80]
PU, ACCuring print0–100.941300[81]
TMAH, PDMS, P(VDF-TrFE)Direct deposition0–9002.9740,000[82]
CNT, FEP, PDMSSpin coating-0.2110,000[83]
PAN, C/BTOCarbon electrospinning0.15–251.1260,000[84]
Organic-based ferrofluids, PEG, PEGDA-0–446.7520,000[85]
Table 2. Comparative study of flexible sensors with different detection methods.
Table 2. Comparative study of flexible sensors with different detection methods.
ClassificationMaterialPreparation MethodLinearity RangeSensitivityLower Limit of DetectionReferences
Physical
Flexible Sensors
PVA, NaClWater Bath Heating24.8–587 KPa-24.8 KPa[106]
AAm, AC, SDS, KPS, TMEDAFine emulsion polymerization method54–935 KPa-54 KPa[107]
[EMIM]Cl, PHEA, PEGDAHybrid heating0–170 KPa-0[108]
PDMS, Basic ElastomerAnisotropic etching0.05–6 KPa317.19 KPa−10.05 KPa[109]
Silver Nano, TMAH, PDMSSynergy of microstructures1–10 KPa2 KPa−11 KPa[110]
Chemical
Flexible Sensors
CuO Nanowires, PETThermal Oxidation0–12 mM-0.05 mM[111]
PET, ITOLaser ScribingPH 4–10−55 mV/pHPH 4[112]
PVIM, MNC, Co5POM-1 fM–5 mM210 µAµM−1cm−21 fM[113]
PDMS, Basic ElastomerAnisotropic Etching2.00–30.00 mM10.72 µAµM−1cm−210[114]
PLLA–PTMC, AuNano film, polybutylenol filmElectrochemical deposition0.01–100 μM-3.97 nM[115]
PEDOT, PSSWet spinningPH 3–7−56 ± 7 mVpH−1PH 3[116]
Table 3. Comparison of flexible sensors using different methods of contact.
Table 3. Comparison of flexible sensors using different methods of contact.
ClassificationMaterialPreparation MethodKPa
Linearity Range KPa
KPa−1
Sensitivity KPa−1
Cycling
Stability (Times)
Reference
Contact
Flexible sensors
PEN, PVDF-TrFEChemical Vapor Deposition40–2000.15200[126]
AGNW, PVP, PETSandpaper forming0–40.0328-[127]
PDMS, CIP, PEGDASpin coating0–20.3015000[128]
Non-contact
Flexible sensors
PEDOT, PSS, EG, OTS, PDMSSpray coating1–50.12000[129]
PPG, TDI, copper powderUltrasonication1.3–38.2-500[130]
PVP, TEMPO, AgNWWet Spinning0–4605.491200[131]
SiO2, Si wafer, SigmaAldrichPhotolithography0.72–11.6-5000[132]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yin, Y.; Guo, C.; Li, H.; Yang, H.; Xiong, F.; Chen, D. The Progress of Research into Flexible Sensors in the Field of Smart Wearables. Sensors 2022, 22, 5089. https://doi.org/10.3390/s22145089

AMA Style

Yin Y, Guo C, Li H, Yang H, Xiong F, Chen D. The Progress of Research into Flexible Sensors in the Field of Smart Wearables. Sensors. 2022; 22(14):5089. https://doi.org/10.3390/s22145089

Chicago/Turabian Style

Yin, Yunlei, Cheng Guo, Hong Li, Hongying Yang, Fan Xiong, and Dongyi Chen. 2022. "The Progress of Research into Flexible Sensors in the Field of Smart Wearables" Sensors 22, no. 14: 5089. https://doi.org/10.3390/s22145089

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