Next Article in Journal / Special Issue
Application of the Artificial Neural Network (ANN) Approach for Prediction of the Kinetic Parameters of Lignocellulosic Fibers
Previous Article in Journal / Special Issue
Electrospun Fe3O4-PVDF Nanofiber Composite Mats for Cryogenic Magnetic Sensor Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Electrically Conductive Textile Materials—Application in Flexible Sensors and Antennas

School of Fashion, Advanced Materials & Liquid Crystals Institute, Kent State University, Kent, OH 44240, USA
Submission received: 26 June 2021 / Revised: 21 July 2021 / Accepted: 27 July 2021 / Published: 30 July 2021
(This article belongs to the Special Issue Fibrous Materials (Textiles) for Functional Applications)

Abstract

:
This paper reviews some prominent applications and approaches to developing smart fabrics for wearable technology. The importance of flexible and electrically conductive textiles in the emerging body-centric sensing and wireless communication systems is highlighted. Examples of applications are discussed with a focus on a range of textile-based sensors and antennas. Developments in alternative materials and structures for producing flexible and conductive textiles are reviewed, including inherently conductive polymers, carbon-based materials, and nano-enhanced composite fibers and fibrous structures.

1. Introduction

Smart fabrics and wearable electronics are set to succeed portable electronics as the next milestone in the modern information technology era. To be considered “smart”, a textile structure must integrate the capability to sense and respond to stimuli from the environment [1,2,3,4,5]. In the broad sense, the stimuli and response of smart textiles can be of varied origins, including chemical, thermal, magnetic, and electrical [1]. The latter is associated with the category of smart textiles that combines electronics with textile structures, referred to as electronic textiles or e-textiles [6], although other terminology such as “textronics” has been used [7,8,9]. One major challenge to the success of wearable e-textile technology resides in the development of lightweight and flexible components, and fibrous structures with high electrical conductivity able to withstand the stresses associated with wearing and caring for the textile [2,6]. Indeed, the lack of flexibility and the weight associated with metallic conductors are key obstacles to overcome in the bourgeoning field of electronic textiles or e-textiles [2,10]. Use of conventional metallic conductors often results in rigid and inflexible fabrics that cannot maintain their functionality when exposed to harsh environmental conditions or after undergoing fabric care processes such as washing [2,10]. Therefore, flexible, deformable, stretchable, and durable conductive threads are critically needed for durable smart fabrics that capture and convey information and enable computing while accommodating the drape and movement of the human body. In recent decades, multiple approaches have been pursued in research to address this challenge using the flexibility and versatility of textile structures, along with innovations in the field of particulate and fibrous materials. This paper presents a review of notable application examples, and some of the approaches reported in the research literature aimed at developing conductive textile structures that are flexible, deformable, and stretchable. In the following sections, we first discuss textile-based sensors and antennas, then review alternative materials adopted for processing conductive textiles, including inherently conductive polymers, carbon-based fibers and additives, and nano-enhanced composite structures.

2. Background—Examples of Applications

In recent years, the function of clothing has expanded beyond basic physical protection from the environment and conquered the role of an interface with the potential to sense, transduce, and communicate information about the body and its surroundings. Clothing is, today, envisioned as a primary node in the concept of Internet of Things (IoT), connecting humans to the broad range of smart objects with digital functionalities [11,12]. For this vision to be realized, there is a pressing need for flexible, deformable, stretchable, and durable conductive textiles that enable smart fabrics to capture and convey information while accommodating the drape and movement of the human body. As a result, new technology has been introduced to textile structures in order to integrate electronic devices, such as sensors [13,14,15,16], antennas [10,17], and energy storage devices [18,19,20].

2.1. Textile-Based Sensors

Examples of textile sensors range from simple responses to chemical or biological stimuli, to more sophisticated signal acquisition, computation, and communication. For instance, nanocomposite nanofiber membranes containing functional fillers that can sense and react to specific chemical stimuli have been envisaged in protective clothing for soldiers exposed to chemical and biological warfare agents, or agricultural workers against pesticides [21,22]. Such applications rely on the chemical reactions between the functional nano-fillers and the target substance to sense and possibly neutralize the potentially toxic chemical compound [21,22].
Another category of textile-based sensors relies on the electrochemical properties and responses of conductive textiles. The sensing capability of those textiles may be a function of the change in electric conductivity with exposure to the target stimulus. For example, Devaux et al. [15] melt-spun conductive multi-filament yarns with conductivity that varies depending on relative humidity, enabling their use as relative humidity textile sensors based on conductivity measurement. Textile sensors based on the change in electrical conductivity are referred to as conductometric sensors and have been suggested for use in detecting a variety of Volatile Organic Compound (VOC) vapors [23,24], and metabolic biological compounds such as urea and uric acid [25]. Other electrochemical sensors, referred to as amperometric or potentiometric, measure the electric current or potential, respectively, resulting from oxidation or reduction reactions in the presence of the targeted substance [14]. Both amperometric and potentiometric textile-based sensors are common in biomedical research [14,26,27,28]. Liu et al. [26] developed an amperometric glucose biosensor using polypyrrole (PPy) nanofibers. Ekanayake et al. [27] adopted a similar principle for an amperometric glucose biosensing, but used PPy nanotubes rather than nanofibers. An example of a textile-based potentiometric sensor was reported by Manjakkal et al. [29] and shown to be responsive to pH variation. Parrilla et al. [30] developed a textile-based stretchable potentiometric sensor array with the capability of simultaneously responding to multiple ions, thus allowing the measurement of electrolyte concentrations in human sweat.
In addition to chemical stimuli, textile-based sensors have also been developed with a response to physical stimuli including pressure, motion, and mechanical strain [31,32,33,34,35,36,37,38,39,40]. Those sensors are typically based on piezoresistive fibrous materials with electrical resistance that varies under mechanical loading [34,36,37,38,39,40,41], or on capacitive textile actuators with capacitance that varies with mechanical strain [31,42,43,44]. Piezoresistive properties are desirable in strain sensor applications because they allow measurement of the response to strain through the corresponding variation of resistivity. However, such variation becomes a challenge when consistent conductivity is needed in stretchable structures [45,46].
Capacitive strain sensors are typically based on the parallel-plate structure (Figure 1) in which one dielectric layer is sandwiched between two conductive electrodes and the capacitance is dependent on the area between the two electrodes [31,42,43,44]. In the structure developed by Nur et al. [43], the dielectric layer is pre-stretched, leading to a wrinkled capacitive sensor at a relaxed state (Figure 1). According to the authors, this design expands the gauge factor (GF or sensitivity) of the sensor beyond the theoretical limit of the material [43]. The strain sensor reported in this research exhibited a GF of 3.05 [43], whereas the maximum sensitivity achieved using the typical parallel plate model is GF = 1 (Figure 1).
A third type of textile-based sensors with response to mechanical stimuli relies on piezoelectric materials and structures, which generate an electric charge when subjected to mechanical loading [47,48,49,50]. Figure 2 shows an illustration of a textile-based piezoelectric pressure sensor developed by Tan et al. [50] using a layered structure consisting of a top and bottom conductive polyester fabric/reduced graphene oxide (rGO) substrates, a polyvinylidene fluoride (PVDF) membrane, and zinc oxide (ZnO) nanorods. The piezoelectric actuator transforms the energy generated by the mechanical stimulus, in this case pressure, into voltage (Figure 2e,f). The same principle based on piezoelectric textile structures has been suggested for use in energy harvesting [47,49,51].
Wearable applications utilizing textile sensors often combine arrays of sensors to simultaneously measure multiple parameters [52,53,54,55,56]. Liu et al. [52] presented a system that monitors health indicators including electrocardiogram (ECG), respiratory activity, and temperature using a combination of a dedicated ECG module, six motion sensors, and a temperature sensor. Fan et al. [55] developed a machine knitted sensor array made of a set of “all-textile” pressure sensors that simultaneously monitors arterial pulse and respiratory activity. Similarly, Keum et al. [56] combined capacitive pressure sensors on a large textile substrate, thus enabling multi-point monitoring.
In most monitoring applications, signal acquisition using sensors of different types is followed by wireless transmission of the signal for data processing. For instance, Mattmann et al. [32] developed a garment prototype with an array of strain sensors to collect data relevant to upper body postures during physical exercise, then wirelessly transmit the data to a computer using Bluetooth. The need for such data transmission has led to efforts seeking the development of textile-based antennas. Flexible planar—also referred to as microstrip—antennas represent the category with the greatest applicability to wearables and smart clothing with wireless communication capability [57]. The following section briefly reviews some of the advances made in developing such lightweight, flexible, and conformal textile-based antennas for wireless data transmission.

2.2. Textile-Based Antennas

With the development of wearable applications integrating both data acquisition and transmission capabilities, the scope of antenna devices has expanded into body-centric systems requiring lightweight and highly conformal alternatives to conventional solutions [58,59]. The use of wearable textile antenna systems was suggested for many fields, including space applications [58,59], medical diagnosis [60], and in support of first responders such as firefighters and paramedics [61,62]. Applications in competitive athletic garments that enable collection and transmission of biomechanical and physiological information were also explored [63].
As mentioned, microstrip patch antenna geometry was identified as the most adaptable to wearables and smart clothing because of its low profile and potential flexibility [57,64,65,66]. A typical microstrip patch antenna consists of a conductive radiating patch, a dielectric substrate, and a ground plane, as shown in Figure 3. The conducting patch is bonded to the dielectric substrate which is superposed on a conductive layer on the opposite side, acting as a ground plane. Figure 4 shows an example implementation of the microstrip patch antenna using a denim fabric as a substrate [67].
When it comes to building antennas, metals have long been the material of choice because of their outstanding electrical conductivities. Thus, to build conformal antennas, researchers [67,68,69,70,71] have developed textile microstrip patch structures by integrating copper tapes (e.g., Figure 4) or metallic woven fabrics (e.g., Figure 5) as patch and ground plates with dielectric fabric substrates. However, thin metallic wires cannot withstand the mechanical stresses in the weaving process; therefore, woven metallic cloths tend to be coarse and stiff [10]. Such metallic cloth is often rigid and inflexible and lacks the conformal properties required for wearable applications.
Flexible alternatives to pure metal conductors have been proposed to address the challenges above. Kohls et al. [72] used nickel/silver plated nylon fabric strips as conductive elements to develop a lightweight and conformal antenna for use in a “Body-Worn Antenna Vest” (BWAV) for military applications. More recently, Bayram et al. [17] developed a conductive fabric patch made of cotton dip-coated with single-walled carbon nanotubes (SWNTs), and further sputter-coated with gold or silver particles for enhanced conductivity. The researchers then embedded the patch in a polydimethylsiloxane (PDMS)-ceramic composite serving as a flexible dielectric substrate [17]. A similar flexible PDMS-ceramic composite substrate was reported by Wang et al. [73]. In Wang et al.’s research, the conductive material used for the radiating patch consisted of a silver-coated filament embroidered on a polyester fabric then affixed to the dielectric composite substrate [73]. In an all-textile implementation of the microstrip patch antenna, bicomponent threads based on traditional Zari silver yarn wrapped around a silk core were embroidered on a cotton substrate (Figure 6) and shown by Anbalagan et al. [74] to be an effective solution for a wearable antenna with robust properties and with industrial, scientific, and medical applicability.
Beyond the approaches discussed above, technology for alternative conductive materials has been sought to overcome the limits of metals in flexible sensors and antennas. The following sections present a review of research focusing on inherently conductive polymers, carbon-based materials, and nanocomposite structures with conductive performance.

3. Alternative Materials for Conductive Textiles

In the field of alternative non-metal-based conductive textiles, electrostatic discharge (ESD) structures are readily achievable and relatively common [75,76,77]. However, components for smart textiles require conductivity that significantly exceeds the ESD range, if they are to be used as substitutes for metallic conductors in fabrics that effectively convey information and enable computing. The ESD Association standard [78] defines dissipative materials as having a conductivity ranging from 1 × 10−11 S/cm to 1 × 10−4 S/cm. On the other hand, conductive materials are defined as materials with conductivity higher than 1 × 10−4 S/cm. Figure 7 illustrates the conductivity ranges of different materials based on the ESD Association standards.
Numerous references in the literature dealing with conductive textiles report conductivities in the dissipative range. Higher conductivity with good mechanical properties must be achieved in order to develop light-weight materials that could replace metallic fabrics. The reported attempts to achieve this goal can be classified in three major categories: (1) inherently conducting polymers, (2) carbon fibers and fibrous structures, and (3) nano-enhanced composite systems combining insulating polymer matrices with conductive additives. In the following section, we comment on those approaches.

3.1. Inherently Conductive Polymers

After Shirakawa, Heeger, and MacDiarmid were awarded the Nobel Prize in Chemistry in 2000 for the discovery of polyacetylene in the mid-1970s, and the subsequent development of conductive polymers into a new research field [80], conductive polymers gained renewed attention because of their potential utilization in antistatic and electromagnetic interference (EMI) shielding applications and in flexible conductive materials [81]. Unlike traditional polymers, conducting polymers have valence electrons in delocalized orbitals which will have high mobility when the materials are “doped” by oxidation [82].
Since the late 1970s, a number of inherently conductive polymers have been studied [83], including polyaniline (PANi) [84,85], PPy [84,86], and PEDOT:PSS [poly(3,4-ethylenedioxythiophene):poly(styrenesulfonate)] [87,88,89]. One of the early challenges with many inherently conductive polymers was their lack of processability and the difficulty of transforming them into usable structures [84]. A notable approach attempted to overcome this challenge was the use of textile substrates as structural reinforcement for the conductive polymers. For instance, Gregory et al. [84] polymerized PPy and PANi onto nylon, polyester, and quartz fabrics and formed flexible conductive textile structures without deteriorating the mechanical and tactile properties of the fabrics.
One of the polymers cited above, PANi, has attracted considerable attention, among the many different types of conductive polymers, because its monomer, aniline, is inexpensive. Oh et al. [90] prepared PANi-nylon conductive fabrics by immersing the nylon fabric in an aniline solution and initiating oxidative polymerization to synthetize the PANi component. The authors report achieving a fabric with a conductivity of 0.6 × 10−1 S/cm. Diaz-de Leon [91] electrospun doped PANi and obtained a fiber with an average diameter of 10 μm and a conductivity of 0.76 S/cm, similar to that of the bulk material. Zhou et al. [92] also fabricated ultra-thin electrospun nanofibers (<30 nm) using doped PANi and reported a bulk fiber conductivity of 10−2 S/cm. In situ polymerization of aniline on non-woven fabrics made of polypropylene was reported by Qi et al. [24] as a successful approach to producing a fabric with conductivity that varies in the presence of various VOCs (including ethanol, chloroform, toluene, acetone and ethyl acetate), enabling its application as conductometric gas sensor. More recently, Sharifi et al. [93] synthetized conductive PANi/cellulosic fiber nanocomposites using the in situ chemical oxidation polymerization of aniline. By optimizing the experimental conditions, the researchers achieved a conductivity of 1.486 × 10−2 S/cm.
Using another conductive polymer, Kaynak et al. [94] prepared conductive yarns of nylon, wool, and cotton. The method they used consisted of applying a layer of PPy on the substrate by continuous vapor polymerization. Vapor-phase polymerization is a process achieved by pre-treating a substrate with an oxidant, then exposing it to the monomer in vapor phase. After exposure of the substrate, a conducting polymer layer forms on the substrate surface. The researchers found that conductivity of the coated yarns could be controlled by modifying reactant concentrations and synthesis parameters. However, the authors state a maximum conductivity “approximately four orders of magnitude smaller than that of metals”.
Using similar methods, Xu et al. [95] reported higher conductivity levels in PPy/poly(p-phenylene terephthalamide) (PPTA) composite fibers with an electrical conductivity of 0.68 S/cm. Recently, Barani et al. [96] imparted conducive properties to cotton fabric through in-situ polymerization of PANi and PPy coatings onto the fabric. The authors compared the sheet resistance of the produced structures and found the PPy coating to produce the best conductivity [96].
According to the literature reviewed, it appears that a wide range of moderate conductivity levels could be achieved depending on the approach adopted (Figure 8). However, one major drawback of the conducting polymers discussed above is poor mechanical properties compared to metals and to other alternatives (see next section). In addition, conductive polymers do not match the electrical properties of metallic and semi-conductive materials and most of the literature shows conductivities in the ESD range or low EMI range at best (refer to Figure 7 and Figure 8).
Although the conductivity levels reported above are limited to the ESD and EMI ranges, the materials and structures based on inherently conductive polymers offer the potential for higher conductivity with optimized processing conditions. There have been reports of relatively high conductivity levels achieved with doped ultra-thin film structures of PEDOT:PSS. Secondary dopants or additives, such as alcohols or high boiling point solvents, are sometimes added to the dispersion to improve the electrical conductivity of PEDOT:PSS films [97,98]. Xia et al. [89] achieved a conductivity of 103 S/cm in a 130-nm thick PEDOT:PSS film through spin-coating and cosolvent doping [89]. Similarly, Yan et al. [88] used centrifugation, spin-coating, and solvent treatment to form a PEDOT:PSS thin film (93 nm in thickness) on a silicon wafer. The authors report a film conductivity in excess of 443 S/cm [88]. It has also been reported that under optimal processing conditions (doping solvent and pressure) PANi could reach a conductivity in the range of 5.35–7.56 S/cm in pellet form [99]. These results indicate the potential for higher conductivity levels in fibrous and laminate structures based on those materials. For instance, notable progress with PDOT:PSS-based fibers was recently made by Seyedin et al. [100,101,102]. The authors produced polyurethane (PU)/PDOT:PSS fibers with a range of conductivity levels depending on PDOT:PSS loading in the composite fiber [100,101,102]. Thus, a high conductivity of up to 25 S/cm was achieved with a 25 wt% loading of PDOT:PSS in PU [100].

3.2. Carbon-Based Fibrous Materials

Carbon fibers are lightweight, flexible, and high-strength materials produced from precursors such as polyacrylonitrile (PAN), pitch, or rayon by a series of carbonization processes. They are used in the construction of aircrafts, space shuttles, sporting goods, and industrial products where performance is sought in conjunction with weight saving. Composites reinforced with carbon fibrous and particulate fillers have been proposed as alternatives to metals in applications such as electrostatic dissipation and electromagnetic shielding [103,104,105].
Carbon fibers offer a range of electrical conductivity levels depending on the precursor and processing conditions. Compared to PAN carbon fibers, which are commonly used in the industry, pitch-based carbon exhibits better electrical conductivity [106]. According to Minus et al. [107], pitch-based carbon fibers have the potential for electrical conductivities as high as 104 S/cm. Using pitch precursors, Park et al. [106] prepared carbon fiber webs by electrospinning followed by stabilization and carbonization of the fiber mat. The electrical conductivity of the fiber web ranged from 63 to 83 S/cm depending on the carbonization temperature. This conductivity is about 10 times higher than the other polymer-based carbonized fibers, such as PAN, polyimide, and polyimidazole.
Another commonly used technique to produce conductive fibrous materials is to homogeneously disperse conductive nanofillers within a polymer matrix. Carbon-based nanomaterials have been widely used to achieve conductivity. Thus, conductive polymer nanocomposites were developed using insulating polymers matrices and conductive nanofiller reinforcements, including carbon nanofibers [108,109], carbon black [110], and carbon nanotubes [111,112,113,114]. Graphene and graphene oxide (GO) have also been cited in research aimed at imparting conductivity to textiles and to developing textile sensors and capacitors [19,39,115,116]. Using a method adapted from traditional yarn winding and twisting processes, Ko et al. [112] electrospun continuous polylactic acid (PLA)/PAN yarns filled with CNTs. In another study by Rosca et al. [114], the researchers synthesized bucky papers made of multiwalled CNT webs in epoxy resin and reported conductivities ranging from 13.68 to 36.29 S/cm.
Conductive carbon-filled composites have been spun into fibers using various synthetic and natural polymers. Regenerated cellulose fibers were solution-spun by Lee et al. [117] with multi-walled carbon nanotube (MWNTs) reinforcement. The authors reported an electrical conductivity of 2.7 S/cm achieved with a MWNT loading of 30 wt%. Also using MWNTs as reinforcement, Bilotti et al. [118] melt-spun thermoplastic polyurethane nanocomposite fibers with varied conductivity levels depending on processing condition.
Other polymers reinforced with carbon fillers to impart conductivity include polyamide [119,120] and polyamide/polypropylene composite [121]. Notably, Ra et al. [122] used MWNTs as a reinforcement in PAN prior to carbonization, thus creating a MWNT/PAN-based carbon nanofiber composite structure. Under optimal carbonization conditions and orientation of the MWNT fillers, the material exhibited a conductivity of up to 35 S/cm with 10 wt% loading of MWNT [122].
Using an ionic liquid solvent, 1-methyl-3-methylimidazolium acetate ([EMIM][Ac]), Miyauchi et al. [123] electrospun a core-sheath fiber structure with MWNT in the core, and cellulose constituting the sheath. The fibers had diameters ranging from several hundreds of nanometers to several micrometers. The electrospun MWNT-cellulose fiber mats exhibited a conductive pathway of bundled MWNTs. Fiber mat conductivity increased with an increasing ratio of MWNT in the fibers; however, the maximum conductivity obtained at 45 wt% MWNT loading was only 0.107 S/cm, which is in the EMI shielding range [124]. Sekitani et al. [125] developed a highly conductive elastic film by dispersing single-walled carbon nanotubes (SWNTs) in a vinylidene fluoride-hexafluoropropylene (PVDF-HFP) copolymer matrix, followed by coating with dimethylsiloxane–based rubber. The conductivity level achieved in this research was 57 S/cm [125]. Dalton et al. [126] used a coagulation-based SWNT/polyvinyl alcohol (PVA) spinning method to produce composite fibers with high SWNT content (60% by weight), and exceptional mechanical and electrical properties [126].
In addition to nanocomposite structures, another approach to imparting electrical conductivity using carbon-based additives consists of coating the target textile structures [116,127]. For instance, Shim et al. [127] coated cotton yarns with both SWNTs and MWNTs using a polyelectrolyte base for the binding. Not unlike the results obtained using the nanocomposite approach, SWNT-coated cotton yarns exhibited a higher conductivity (approximately 4 × 10−2 S/cm) compared with MWNT samples [127].
Xiang et al. [116] coated Kevlar fibers with three types of carbon materials: graphene nanoribbons (GNRs), MWNTs, and SWNTs. The carbon structures were affixed to the fiber surface through layer-by-layer spray-coating using polyurethane as an interlayer binder, which provided strong adherence on the fiber [116] (Figure 9). The authors found that among the three types of carbon materials, SWNTs coatings produced the best results with a durable conductivity of 65 S/cm. Using GO in a succession of dip-coating and infra-red baking on a polyester fabric substrate, Yang et al. [39] achieved a conductive structure with piezoresistive properties useable as a strain sensor. With an analogous dipping and drying process of cotton fabric in a SWNT ink, Hu et al. [18] produced a highly flexible and stretchable conductive fabric with a conductivity of 125 S/cm.
Based on the review discussed above, carbon-based fibrous materials appear to offer the potential for higher conductivity than the previous approach using conductive polymers. Representative conductivity results reported with carbon coatings, nanofibers, and nanofillers are summarized in Figure 10.

3.3. Metal Nanocomposite and Nano-Enhanced Conductive Materials

Other than the widely used carbon fillers discussed above, textile structures with metal nano-additives have also been extensively documented with a broad range of applications depending on the conductivity performance. For instance, conductivity levels in the electrostatic charge dissipation range were achieved by Zhou et al. [128] by incorporating zinc oxide whiskers in polyurethane, PVA, and natural rubber. Xue et al. [129] coated woven cotton fabric with silver nanoparticles and generated a multifunctional fabric combining biocidal and conductive performance.
Nickel is another metal that has been reported in research to impart electrical conductivity to textile structures using both coating and nano-filling approaches [130,131,132,133,134,135]. As a nano-filler, nickel has been commercialized in the form of nanostrands (NiNs) with a branched 3D structure (Figure 11) generated by chemical vapor deposition [132,136,137,138]. The branched 3D structure has been cited as an advantage of NiNs over other nanofillers (including carbon and metals) in terms of electrical conductivity because it enables more interconnects and thus more conductive pathways within the polymer matrix [136,139,140,141]. NiNs have been reported as nanofillers with electrical conductivity in polymers such as nylon [132], as well as epoxy, thermoplastic polyurethane, and polyimide [137,138,139,140].
Attempts were also made to incorporate NiNS in inherently conductive polymers such as PEDOT:PSS and yielded a significant two-orders-of-magnitude improvement in conductivity [132]. The investigation of the electrical properties of nanocomposites made from intrinsically conductive polymers has been of interest. Gangopadhyay and De [142] discussed the use of various inorganic nanoparticles, such as SiO2, SnO2, TiO2, and Fe2O3, in reinforcing several intrinsically conductive polymers including PPy, PANi, and PEDOT.
In terms of coating approaches, Guo et al. [130] deposited a nickel layer on polyester fabric using electroless plating and succeeded in reaching conductivity levels with effective EMI shielding. More recently, Moazzenchi and Montazer [131] used a similar electroless plating method to synthetize nickel nanoparticles onto polyester fabric, yielding superior EMI shielding performance.
Recently, there has been growing interest in bimetallic layered double hydroxide (LDH) nanostructures, with metals such as nickel, cobalt, and aluminum as important components [133,134,135]. For instance, Nagaraju et al. [133] grew nickel-cobalt layered double hydroxide (Ni-Co LDH) nanosheets on conductive fabrics useable for energy storage. Similarly, Lu et al. [135] electrodeposited nickel-aluminum layered double hydroxide (Ni-Al LDH) onto cotton fabrics and reported a superior conductivity of 143 S/cm.
Another recent innovation in developing stretchable and highly conductive textile structures was reported by Zheng et al. [46]. The researchers aimed to circumvent the potential for deterioration of conductivity in e-textiles due to mechanical strain and used a coaxial wet spinning process to generate liquid metal sheath-core microfibers with a conductivity of 435 S/cm that remains stable within a 4% range at strains of up to 200% [46]. The liquid metal core used in this research consists of eutectic gallium indium (EGaIn) alloy, which is liquid at room temperature, and offers high electrical conductivity (3.4 × 104 S/cm) that is maintained under mechanical deformation [46,143]. Poly(vinylidene fluoride-hexafluoropropylene-tetrafluoroethylene)/poly(ethylene glycol) diacrylate (PVDF-HFP-TFE/PEGDA) was used as the solution in the sheath channel of the coaxial spinneret [46].
Overall, the different approaches discussed above led to the development of a broad range of flexible, stretchable, and conformal textile structures with conductivity levels that range from antistatic to fully conductive performance. Current challenges remain in ensuring consistent performance while withstanding the mechanical stresses endured during wear and fabric care. For instance, resistance to washing is considered one of the primary challenges to the marketability of e-textiles [144,145,146,147]. One of the problems highlighted by Rotzler et al. [144] is the absence of standardized methods for testing the washability of e-textiles, leading to inconsistent washability data. Another review by the same authors [145] presents a comprehensive summary of challenges related to the washability of e-textiles, and the various failure modes caused by washing and impacting reliability. Because of the growing awareness of those challenges, washability and robust performance after laundering have come to the forefront of concerns in recent research aimed at developing e-textiles [146,147,148,149]. Recent developments such as metal-based nanostructures and liquid metal core-sheath filaments offer promising solutions toward durable and stable conductivity levels that can withstand wear and wash.

4. Summary

Flexible smart fabrics and conformal textile-based electronic devices are a critical component in the development of body-centric sensing and information communication systems. In the concept of the Internet of Things (IoT), smart clothing is viewed as the interface connecting humans to the numerous and more and more ubiquitous objects with digital functionalities. This paper reviews some representative textile applications for information acquisition and transmission, i.e., textile-based sensors and antennas. In both categories, research has evolved from reliance on metallic structures with limitations in flexibility and drape, to more and more flexible and conformal all-textile structures with effective performance. To achieve this goal, textiles with a range of electrical conductivity have been developed using various approaches, including inherently conductive polymers, carbon-based conductors, and nano-enhanced fibers and fibrous structures. Conductivities achieved in these structures ranged from the antistatic and dissipative bands using inherently conductive polymers, to the semi-conductor and conductor bands achieved with carbon fibrous structures and nanofillers, and other innovations such as Ni-Al LDH coating and liquid-metal core-sheath fibers and filaments.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Van Langenhove, L.; Puers, R.; Matthys, D. Intelligent textiles for protection. In Textiles for Protection; Scott, R.A., Ed.; Woodhead Publishing: Cambridge, UK, 2005; pp. 176–195. [Google Scholar]
  2. Cherenack, K.; Van Pieterson, L. Smart textiles: Challenges and opportunities. J. Appl. Phys. 2012, 112, 091301. [Google Scholar] [CrossRef] [Green Version]
  3. van Langenhove, L. Smart Textiles for Protection: An Overview, in Smart Textiles for Protection; Chapman, R.A., Ed.; Woodhead Publishing: Cambridge, UK, 2013; pp. 3–33. [Google Scholar]
  4. Koncar, V. Introduction to smart textiles and their applications, in Smart Textiles and their Applications; Koncar, V., Ed.; Woodhead Publishing: Cambridge, UK, 2016; pp. 1–8. [Google Scholar]
  5. Kongahage, D.; Foroughi, J. Actuator Materials: Review on Recent Advances and Future Outlook for Smart Textiles. Fibers 2019, 7, 21. [Google Scholar] [CrossRef] [Green Version]
  6. Köhler, A.R. Challenges for eco-design of emerging technologies: The case of electronic textiles. Mater. Des. 2013, 51, 51–60. [Google Scholar] [CrossRef]
  7. Bielska, S.; Sibiński, M.; Lukasik, A. Polymer temperature sensor for textronic applications. Mater. Sci. Eng. B 2009, 165, 50–52. [Google Scholar] [CrossRef]
  8. Gniotek, K.; Krucińska, I. The basic problems of textronics. Fibers Text. East. Eur. 2004, 12, 13–16. [Google Scholar]
  9. Ho, D.H.; Cheon, S.; Hong, P.; Park, J.H.; Suk, J.W.; Kim, D.H.; Han, J.T.; Cho, J.H. Multifunctional Smart Textronics with Blow-Spun Nonwoven Fabrics. Adv. Funct. Mater. 2019, 29, 1900025. [Google Scholar] [CrossRef]
  10. Shawl, R.K.; Long, B.R.; Werner, D.H.; Gavrin, A. The Characterization of Conductive Textile Materials Intended for Radio Frequency Applications. IEEE Antennas Propag. Mag. 2007, 49, 28–40. [Google Scholar] [CrossRef]
  11. Kranz, M.; Holleis, P.; Schmidt, A. Embedded Interaction: Interacting with the Internet of Things. IEEE Internet Comput. 2009, 14, 46–53. [Google Scholar] [CrossRef] [Green Version]
  12. Metcalf, D.; Milliard, S.T.; Gomez, M.; Schwartz, M. Wearables and the Internet of Things for Health: Wearable, Interconnected Devices Promise More Efficient and Comprehensive Health Care. IEEE Pulse 2016, 7, 35–39. [Google Scholar] [CrossRef]
  13. Huang, J.; Virji, S.; Weiller, B.H.; Kaner, R.B. Polyaniline Nanofibers: Facile Synthesis and Chemical Sensors. J. Am. Chem. Soc. 2003, 125, 314–315. [Google Scholar] [CrossRef] [PubMed]
  14. Nambiar, S.; Yeow, J.T. Conductive polymer-based sensors for biomedical applications. Biosens. Bioelectron. 2011, 26, 1825–1832. [Google Scholar] [CrossRef] [PubMed]
  15. Devaux, E.; Aubry, C.; Campagne, C.; Rochery, M. PLA/Carbon Nanotubes Multifilament Yarns for Relative Humidity Textile Sensor. J. Eng. Fibers Fabr. 2011, 6, 13–24. [Google Scholar] [CrossRef] [Green Version]
  16. Wang, J.; Lu, C.; Zhang, K. Textile-Based Strain Sensor for Human Motion Detection. Energy Environ. Mater. 2020, 3, 80–100. [Google Scholar] [CrossRef] [Green Version]
  17. Bayram, Y.; Zhou, Y.; Shim, B.S.; Xu, S.; Zhu, J.; Kotov, N.; Volakis, J.L. E-Textile Conductors and Polymer Composites for Conformal Lightweight Antennas. IEEE Trans. Antennas Propag. 2010, 58, 2732–2736. [Google Scholar] [CrossRef]
  18. Hu, L.; Pasta, M.; La Mantia, F.; Cui, L.; Jeong, S.; Deshazer, H.D.; Choi, J.W.; Han, S.M.; Cui, Y. Stretchable, Porous, and Conductive Energy Textiles. Nano Lett. 2010, 10, 708–714. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Lu, C.; Meng, J.; Zhang, J.; Chen, X.; Du, M.; Chen, Y.; Hou, C.; Wang, J.; Ju, A.; Wang, X.; et al. Three-Dimensional Hierarchically Porous Graphene Fiber-Shaped Supercapacitors with High Specific Capacitance and Rate Capability. ACS Appl. Mater. Interfaces 2019, 11, 25205–25217. [Google Scholar] [CrossRef] [PubMed]
  20. Xu, Q.; Lu, C.; Sun, S.; Zhang, K. Electrochemical properties of PEDOT:PSS/V2O5 hybrid fiber based supercapacitors. J. Phys. Chem. Solids 2019, 129, 234–241. [Google Scholar] [CrossRef]
  21. Schreuder-Gibson, H.L.; Truong, Q.; Walker, J.E.; Owens, J.R.; Wander, J.D.; Jones, W.E. Chemical and Biological Protection and Detection in Fabrics for Protective Clothing. MRS Bull. 2003, 28, 574–578. [Google Scholar] [CrossRef] [Green Version]
  22. Bhuiyan, M.A.R.; Wang, L.; Shaid, A.; Shanks, R.A.; Ding, J. Advances and applications of chemical protective clothing system. J. Ind. Text. 2019, 49, 97–138. [Google Scholar] [CrossRef]
  23. Sharma, S.; Nirkhe, C.; Pethkar, S.; Athawale, A.A. Chloroform vapour sensor based on copper/polyaniline nanocomposite. Sens. Actuators B Chem. 2002, 85, 131–136. [Google Scholar] [CrossRef]
  24. Qi, J.; Xu, X.; Liu, X.; Lau, K.T. Fabrication of textile based conductometric polyaniline gas sensor. Sens. Actuators B Chem. 2014, 202, 732–740. [Google Scholar] [CrossRef]
  25. Castillo-Ortega, M.; Rodriguez, D.; Encinas, J.C.; Plascencia-Jatomea, M.; Méndez-Velarde, F.; Olayo, R. Conductometric uric acid and urea biosensor prepared from electroconductive polyaniline–poly(n-butyl methacrylate) composites. Sens. Actuators B Chem. 2002, 85, 19–25. [Google Scholar] [CrossRef]
  26. Liu, L.; Jia, N.-Q.; Zhou, Q.; Yan, M.-M.; Jiang, Z.-Y. Electrochemically fabricated nanoelectrode ensembles for glucose biosensors. Mater. Sci. Eng. C 2007, 27, 57–60. [Google Scholar] [CrossRef]
  27. Mala Ekanayake, E.; Preethichandra, D.; Kaneto, K. Polypyrrole nanotube array sensor for enhanced adsorption of glucose oxidase in glucose biosensors. Biosens. Bioelectron. 2007, 23, 107–113. [Google Scholar] [CrossRef] [PubMed]
  28. Gualandi, I.; Tessarolo, M.; Mariani, F.; Possanzini, L.; Scavetta, E.; Fraboni, B. Textile Chemical Sensors Based on Conductive Polymers for the Analysis of Sweat. Polymers 2021, 13, 894. [Google Scholar] [CrossRef] [PubMed]
  29. Manjakkal, L.; Dang, W.; Yogeswaran, N.; Dahiya, R. Textile-Based Potentiometric Electrochemical pH Sensor for Wearable Applications. Biosensors 2019, 9, 14. [Google Scholar] [CrossRef] [Green Version]
  30. Parrilla, M.; Cánovas, R.; Jeerapan, I.; Andrade, F.J.; Wang, J. A Textile-Based Stretchable Multi-Ion Potentiometric Sensor. Adv. Health Mater. 2016, 5, 996–1001. [Google Scholar] [CrossRef]
  31. Meyer, J.; Lukowicz, P.; Troster, G. Textile Pressure Sensor for Muscle Activity and Motion Detection. In Proceedings of the 2006 10th IEEE International Symposium on Wearable Computers, Montreux, Switzerland, 11–14 October 2006; pp. 69–72. [Google Scholar]
  32. Mattmann, C.; Amft, O.; Harms, H.; Troster, G.; Clemens, F. Recognizing Upper Body Postures using Textile Strain Sensors. In Proceedings of the 2007 11th IEEE International Symposium on Wearable Computers, Boston, MA, USA, 11–13 October 2007; pp. 29–36. [Google Scholar]
  33. Mattmann, C.; Clemens, F.; Tröster, G. Sensor for Measuring Strain in Textile. Sensors 2008, 8, 3719–3732. [Google Scholar] [CrossRef]
  34. 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]
  35. Munro, B.J.; Campbell, T.E.; Wallace, G.; Steele, J. The intelligent knee sleeve: A wearable biofeedback device. Sens. Actuators B Chem. 2008, 131, 541–547. [Google Scholar] [CrossRef]
  36. Melnykowycz, M.; Koll, B.; Scharf, D.; Clemens, F. Comparison of Piezoresistive Monofilament Polymer Sensors. Sensors 2014, 14, 1278–1294. [Google Scholar] [CrossRef]
  37. Lee, T.; Lee, W.; Kim, S.-W.; Kim, J.J.; Kim, B.-S. Flexible Textile Strain Wireless Sensor Functionalized with Hybrid Carbon Nanomaterials Supported ZnO Nanowires with Controlled Aspect Ratio. Adv. Funct. Mater. 2016, 26, 6206–6214. [Google Scholar] [CrossRef]
  38. Oliveri, A.; Maselli, M.; Lodi, M.; Storace, M.; Cianchetti, M. Model-Based Compensation of Rate-Dependent Hysteresis in a Piezoresistive Strain Sensor. IEEE Trans. Ind. Electron. 2018, 66, 8205–8213. [Google Scholar] [CrossRef]
  39. Yang, Z.; Pang, Y.; Han, X.-L.; Yang, Y.; Ling, J.; Jian, M.; Zhang, Y.; Yang, Y.; Ren, T.-L. Graphene Textile Strain Sensor with Negative Resistance Variation for Human Motion Detection. ACS Nano 2018, 12, 9134–9141. [Google Scholar] [CrossRef] [PubMed]
  40. Raji, R.K.; Miao, X.; Zhang, S.; Li, Y.; Wan, A.; Boakye, A. Knitted piezoresistive strain sensor performance, impact of conductive area and profile design. J. Ind. Text. 2020, 50, 616–634. [Google Scholar] [CrossRef]
  41. Wang, C.; Tianling, R.; Gao, E.; Jian, M.; Xia, K.; Wang, Q.; Xu, Z.; Ren, T.; Zhang, Y. Carbonized Silk Fabric for Ultrastretchable, Highly Sensitive, and Wearable Strain Sensors. Adv. Mater. 2016, 28, 6640–6648. [Google Scholar] [CrossRef] [PubMed]
  42. Atalay, O. Textile-Based, Interdigital, Capacitive, Soft-Strain Sensor for Wearable Applications. Materials 2018, 11, 768. [Google Scholar] [CrossRef] [Green Version]
  43. Nur, R.; Matsuhisa, N.; Jiang, Z.; Nayeem, M.O.G.; Yokota, T.; Someya, T. A Highly Sensitive Capacitive-type Strain Sensor Using Wrinkled Ultrathin Gold Films. Nano Lett. 2018, 18, 5610–5617. [Google Scholar] [CrossRef]
  44. Zhang, Q.; Wang, Y.L.; Xia, Y.; Zhang, P.F.; Kirk, T.V.; Chen, X.D. Textile-Only Capacitive Sensors for Facile Fabric Integration without Compromise of Wearability. Adv. Mater. Technol. 2019, 4, 1900485. [Google Scholar] [CrossRef]
  45. Wang, C.; Zhang, M.; Xia, K.; Gong, X.; Wang, H.; Yin, Z.; Guan, B.; Zhang, Y. Intrinsically Stretchable and Conductive Textile by a Scalable Process for Elastic Wearable Electronics. ACS Appl. Mater. Interfaces 2017, 9, 13331–13338. [Google Scholar] [CrossRef]
  46. Zheng, L.; Zhu, M.; Wu, B.; Li, Z.; Sun, S.; Wu, P. Conductance-stable liquid metal sheath-core microfibers for stretchy smart fabrics and self-powered sensing. Sci. Adv. 2021, 7, eabg4041. [Google Scholar] [CrossRef]
  47. Nilsson, E.; Lund, A.; Jonasson, C.; Johansson, C.; Hagström, B. Poling and characterization of piezoelectric polymer fibers for use in textile sensors. Sens. Actuators A Phys. 2013, 201, 477–486. [Google Scholar] [CrossRef]
  48. Åkerfeldt, M.; Nilsson, E.; Gillgard, P.; Walkenström, P. Textile piezoelectric sensors–melt spun bi-component poly(vinylidene fluoride) fibres with conductive cores and poly(3,4-ethylene dioxythiophene)-poly(styrene sulfonate) coating as the outer electrode. Fash. Text. 2014, 1, 13. [Google Scholar] [CrossRef] [Green Version]
  49. Nilsson, E.; Mateu, L.; Spies, P.; Hagström, B. Energy Harvesting from Piezoelectric Textile Fibers. Procedia Eng. 2014, 87, 1569–1572. [Google Scholar] [CrossRef]
  50. Tan, Y.; Yang, K.; Wang, B.; Li, H.; Wang, L.; Wang, C. High-performance textile piezoelectric pressure sensor with novel structural hierarchy based on ZnO nanorods array for wearable application. Nano Res. 2021, 1–8. [Google Scholar] [CrossRef]
  51. Granstrom, J.; Feenstra, J.; Sodano, H.A.; Farinholt, K.M. Energy harvesting from a backpack instrumented with piezoelectric shoulder straps. Smart Mater. Struct. 2007, 16, 1810–1820. [Google Scholar] [CrossRef]
  52. Liu, J.; Xie, F.; Zhou, Y.; Zou, Q.; Wu, J. A wearable health monitoring system with multi-parameters. In Proceedings of the 2013 6th International Conference on Biomedical Engineering and Informatics, Hangzhou, China, 16–18 December 2013. [Google Scholar]
  53. Ahn, Y.; Song, S.; Yun, K.-S. Woven flexible textile structure for wearable power-generating tactile sensor array. Smart Mater. Struct. 2015, 24, 075002. [Google Scholar] [CrossRef]
  54. Zhu, Z.; Liu, T.; Li, G.; Li, T.; Inoue, Y. Wearable Sensor Systems for Infants. Sensors 2015, 15, 3721–3749. [Google Scholar] [CrossRef] [PubMed]
  55. Fan, W.; He, Q.; Meng, K.; Tan, X.; Zhou, Z.; Zhang, G.; Yang, J.; Wang, Z.L. Machine-knitted washable sensor array textile for precise epidermal physiological signal monitoring. Sci. Adv. 2020, 6, eaay2840. [Google Scholar] [CrossRef] [Green Version]
  56. 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]
  57. Salonen, P.; Keskilammi, M.; Rantanen, J.; Sydänheimo, L. A novel Bluetooth antenna on flexible substrate for smart clothing. In Proceedings of the 2001 IEEE International Conference on Systems, Man and Cybernetics. e-Systems and e-Man for Cybernetics in Cyberspace (Cat.No.01CH37236), Tucson, AZ, USA, 7–10 October 2001. [Google Scholar]
  58. Kennedy, T.; Fink, P.; Chu, A.; Studor, G. Potential space applications for body-centric wireless and E-textile antennas. In Proceedings of the 2007 IET Seminar on Antennas and Propagation for Body-Centric Wireless Communications, London, UK, 24–24 April 2007; Institution of Engineering and Technology (IET): London, UK, 2007; pp. 77–83. [Google Scholar] [CrossRef] [Green Version]
  59. Kennedy, T.F.; Fink, P.W.; Chu, A.W.; Champagne, N.J.; Lin, G.Y.; Khayat, M.A. Body-Worn E-Textile Antennas: The Good, the Low-Mass, and the Conformal. IEEE Trans. Antennas Propag. 2009, 57, 910–918. [Google Scholar] [CrossRef]
  60. SankarGanesh, S.; Sreelatha, P.; Rekha, V.B.; Puttumraju, A.K.; Murugesan, B.; Sakthisudhan, K. Textile antennas for breast carcinoma diagnosis application. Mater. Today Proc. 2021, 45, 3147–3152. [Google Scholar] [CrossRef]
  61. Hertleer, C.; Van Langenhove, L.; Rogier, H.; Vallozzi, L. A Textile Antenna For Fire Fighter Garments. In Proceedings of the Autex 2007 Conference, Tampere, Finland, 26–28 June 2007. [Google Scholar]
  62. Vallozzi, L.; Van Torre, P.; Hertleer, C.; Rogier, H.; Moeneclaey, M.; Verhaevert, J. Wireless Communication for Firefighters Using Dual-Polarized Textile Antennas Integrated in Their Garment. IEEE Trans. Antennas Propag. 2010, 58, 1357–1368. [Google Scholar] [CrossRef] [Green Version]
  63. Galehdar, A.; Thiel, D.V. Flexible, light-weight antenna at 2.4 GHz for athlete clothing. In Proceedings of the 2007 IEEE Antennas and Propagation Society International Symposium, Honolulu, HI, USA, 9–15 June 2007. [Google Scholar]
  64. Locher, I.; Klemm, M.; Kirstein, T.; Troster, G. Design and Characterization of Purely Textile Patch Antennas. IEEE Trans. Adv. Packag. 2006, 29, 777–788. [Google Scholar] [CrossRef] [Green Version]
  65. Salonen, P.; Rahmat-Samii, Y. Textile antennas: Effects of antenna bending on input matching and impedance bandwidth. Aerosp. Electron. Syst. Mag. 2007, 22, 18–22. [Google Scholar] [CrossRef]
  66. Hertleer, C.; Tronquo, A.; Rogier, H.; van Langenhove, L. The Use of Textile Materials to Design Wearable Microstrip Patch Antennas. Text. Res. J. 2008, 78, 651–658. [Google Scholar] [CrossRef]
  67. Purohit, S.; Rava, F. Wearable—Textile Patch Antenna using Jeans as Substrate at 2.45 GHz. Int. J. Eng. Res. Technol. 2014, 3, 2456–2460. Available online: https://www.ijert.org/wearable-textile-patch-antenna-using-jeans-as-substrate-at-2.45-ghz (accessed on 6 May 2021).
  68. Ouyang, Y.; Karayianni, E.; Chappell, W.J. Effect of fabric patterns on electrotextile patch antennas. In Proceedings of the Antennas and Propagation Society International Symposium, 2005 IEEE, Washington, DC, USA, 3–8 July 2005. [Google Scholar]
  69. Ouyang, Y.; Chappell, W.J. High Frequency Properties of Electro-Textiles for Wearable Antenna Applications. IEEE Trans. Antennas Propag. 2008, 56, 381–389. [Google Scholar] [CrossRef]
  70. Yao, L.; Qiu, Y. Design and fabrication of microstrip antennas integrated in three dimensional orthogonal woven composites. Compos. Sci. Technol. 2009, 69, 1004–1008. [Google Scholar] [CrossRef]
  71. Xu, F.; Yao, L.; Zhao, D.; Jiang, M.; Qiu, Y. Effect of Weaving Direction of Conductive Yarns on Electromagnetic Performance of 3D Integrated Microstrip Antenna. Appl. Compos. Mater. 2013, 20, 827–838. [Google Scholar] [CrossRef] [Green Version]
  72. Kohls, E.; Abler, A.; Siemsen, P.; Hughes, J.L.A.; Perez, R.V.; Widdoes, D. A multi-band body-worn antenna vest. In Proceedings of the IEEE Antennas and Propagation Society Symposium, Monterey, CA, USA, 20–25 June 2004; Institute of Electrical and Electronics Engineers (IEEE): Piscataway, NJ, USA, 2004; Volume 1, pp. 447–450. [Google Scholar]
  73. Wang, Z.; Zhang, L.; Bayram, Y.; Volakis, J.L. Embroidered Conductive Fibers on Polymer Composite for Conformal Antennas. IEEE Trans. Antennas Propag. 2012, 60, 4141–4147. [Google Scholar] [CrossRef]
  74. Anbalagan, A.; Sundarsingh, E.F.; Ramalingam, V.S. Design and experimental evaluation of a novel on-body textile antenna for unicast applications. Microw. Opt. Technol. Lett. 2020, 62, 789–799. [Google Scholar] [CrossRef]
  75. Cheng, K.B.; Ueng, T.H.; Dixon, G. Electrostatic Discharge Properties of Stainless Steel/Polyester Woven Fabrics. Text. Res. J. 2001, 71, 732–738. [Google Scholar] [CrossRef]
  76. Hebeish, A.A.; El-Gamal, M.A.; Said, T.S.; El-Hady, R.A.A. Major factors affecting the performance of ESD-protective fabrics. J. Text. Inst. 2010, 101, 389–398. [Google Scholar] [CrossRef]
  77. Zhang, X. 2—Antistatic and conductive textiles. In Functional Textiles for Improved Performance, Protection and Health; Pan, N., Sun, G., Eds.; Woodhead Publishing: Southston, UK, 2011; pp. 27–44. [Google Scholar]
  78. Electrostatic Discharge Association ESD Association Advisory for Electrostatic Discharge Terminology-Glossary. ESD ADV1.0–2009. Available online: www.esda.org/assets/Documents/c23d92d4ab/Fundamentals-of-ESD-Part-1-An-Introduction-to-ESD.pdf (accessed on 25 March 2018).
  79. Hyperion Catalysis International. Electrical Resistivity in Semi-Crystalline Polymers. 2002. Available online: www.hyperioncatalysis.com (accessed on 24 August 2012).
  80. Nordén, B.; Krutmeijer, E. The Nobel Prize in Chemistry, 2000: Conductive Polymers; The Royal Swedish Academy of Sciences: Stockholm, Sweden, 2000. [Google Scholar]
  81. Maity, S.; Singha, K.; Pandit, P. 11—Advanced applications of green materials in electromagnetic shielding. In Applications of Advanced Green Materials; Ahmed, S., Ed.; Woodhead Publishing: Duxford, UK, 2021; pp. 265–292. [Google Scholar]
  82. Heinze, J. Self-Doped Conducting Polymers. By Michael S. Freund and Bhavana Deore. Angew. Chem. Int. Ed. 2007, 46, 7922. [Google Scholar] [CrossRef]
  83. Aldissi, M. Advances in inherently conducting polymers. Makromol. Chem. Macromol. Symp. 1989, 24, 1–20. [Google Scholar] [CrossRef] [Green Version]
  84. Gregory, R.; Kimbrell, W.; Kuhn, H. Conductive textiles. Synth. Met. 1989, 28, 823–835. [Google Scholar] [CrossRef]
  85. Kang, E.T.; Neoh, K.G.; Tan, K.L. Polyaniline: A polymer with many interesting intrinsic redox states. Prog. Polym. Sci. 1998, 23, 277–324. [Google Scholar] [CrossRef]
  86. Ramanavičius, A.; Ramanavičienė, A.; Malinauskas, A. Electrochemical sensors based on conducting polymer—Polypyrrole. Electrochim. Acta 2006, 51, 6025–6037. [Google Scholar] [CrossRef]
  87. Jang, J.; Chang, M.; Yoon, H. Chemical Sensors Based on Highly Conductive Poly(3,4-ethylenedioxythiophene) Nanorods. Adv. Mater. 2005, 17, 1616–1620. [Google Scholar] [CrossRef]
  88. Yan, H.; Jo, T.; Okuzaki, H. Highly Conductive and Transparent Poly(3,4-ethylenedioxythiophene)/Poly(4-styrenesulfonate) (PEDOT/PSS) Thin Films. Polym. J. 2009, 41, 1028–1029. [Google Scholar] [CrossRef]
  89. Xia, Y.; Ouyang, J. PEDOT:PSS films with significantly enhanced conductivities induced by preferential solvation with cosolvents and their application in polymer photovoltaic cells. J. Mater. Chem. 2011, 21, 4927–4936. [Google Scholar] [CrossRef]
  90. Oh, K.W.; Hong, K.H.; Kim, S.H. Electrically conductive textiles byin situ polymerization of aniline. J. Appl. Polym. Sci. 1999, 74, 2094–2101. [Google Scholar] [CrossRef]
  91. Diaz-de Leon, M.J. Electrospinning Nanofibers of Polyaniline and Polyaniline/(Polystyrene and Polyehtylene Oxide) Blends. In Proceedings of the National Conference on Undergraduate Research (NCUR), Lexington, KY, USA, 15–17 March 2001. [Google Scholar]
  92. Zhou, Y.; Freitag, M.; Hone, J.; Staii, C.; Johnson, A.T.; Pinto, N.J.; MacDiarmid, A.G. Fabrication and electrical characterization of polyaniline-based nanofibers with diameter below 30 nm. Appl. Phys. Lett. 2003, 83, 3800–3802. [Google Scholar] [CrossRef] [Green Version]
  93. Sharifi, H.; Zabihzadeh, S.M.; Ghorbani, M. The application of response surface methodology on the synthesis of conductive polyaniline/cellulosic fiber nanocomposites. Carbohydr. Polym. 2018, 194, 384–394. [Google Scholar] [CrossRef] [PubMed]
  94. Kaynak, A.; Najar, S.S.; Foitzik, R.C. Conducting nylon, cotton and wool yarns by continuous vapor polymerization of pyrrole. Synth. Met. 2008, 158, 1–5. [Google Scholar] [CrossRef] [Green Version]
  95. Xu, C.; Wang, P.; Bi, X. Continuous vapor phase polymerization of pyrrole. I. Electrically conductive composite fiber of polypyrrole with poly(p-phenylene terephthalamide). J. Appl. Polym. Sci. 1995, 58, 2155–2159. [Google Scholar] [CrossRef]
  96. Barani, H.; Miri, A.; Sheibani, H. Comparative study of electrically conductive cotton fabric prepared through the in situ synthesis of different conductive materials. Cellulose 2021, 28, 6629–6649. [Google Scholar] [CrossRef]
  97. Crispin, X.; Jakobsson, F.L.E.; Crispin, A.; Grim, P.C.M.; Andersson, P.; Volodin, A.; Van Haesendonck, C.; Van der Auweraer, M.; Salaneck, W.R.; Berggren, M. The Origin of the High Conductivity of Poly(3,4-ethylenedioxythiophene)−Poly(styrenesulfonate) (PEDOT−PSS) Plastic Electrodes. Chem. Mater. 2006, 18, 4354–4360. [Google Scholar] [CrossRef]
  98. Crispin, X.; Marciniak, S.; Osikowicz, W.; Zotti, G.; Van Der Gon, A.A.D.; Louwet, F.; Fahlman, M.; Groenendaal, L.B.; De Schryver, F.; Salaneck, W.R. Conductivity, morphology, interfacial chemistry, and stability of poly(3,4-ethylene dioxythiophene)-poly(styrene sulfonate): A photoelectron spectroscopy study. J. Polym. Sci. Part B Polym. Phys. 2003, 41, 2561–2583. [Google Scholar] [CrossRef]
  99. Zhang, D. On the conductivity measurement of polyaniline pellets. Polym. Test. 2007, 26, 9–13. [Google Scholar] [CrossRef]
  100. Seyedin, S.; Razal, J.M.; Innis, P.; Wallace, G.G. Strain-Responsive Polyurethane/PEDOT:PSS Elastomeric Composite Fibers with High Electrical Conductivity. Adv. Funct. Mater. 2014, 24, 2957–2966. [Google Scholar] [CrossRef] [Green Version]
  101. Seyedin, S.; Razal, J.M.; Innis, P.C.; Jeiranikhameneh, A.; Beirne, S.; Wallace, G.G. Knitted Strain Sensor Textiles of Highly Conductive All-Polymeric Fibers. ACS Appl. Mater. Interfaces 2015, 7, 21150–21158. [Google Scholar] [CrossRef] [PubMed]
  102. Seyedin, S.; Moradi, S.; Singh, C.; Razal, J.M. Continuous production of stretchable conductive multifilaments in kilometer scale enables facile knitting of wearable strain sensing textiles. Appl. Mater. Today 2018, 11, 255–263. [Google Scholar] [CrossRef]
  103. Clingerman, M.L.; Weber, E.H.; King, J.A.; Schulz, K.H. Development of an additive equation for predicting the electrical conductivity of carbon-filled composites. J. Appl. Polym. Sci. 2003, 88, 2280–2299. [Google Scholar] [CrossRef]
  104. Mahmoodi, M.; Arjmand, M.; Sundararaj, U.; Park, S. The electrical conductivity and electromagnetic interference shielding of injection molded multi-walled carbon nanotube/polystyrene composites. Carbon 2012, 50, 1455–1464. [Google Scholar] [CrossRef]
  105. Munalli, D.; Dimitrakis, G.; Chronopoulos, D.; Greedy, S.; Long, A. Electromagnetic shielding effectiveness of carbon fibre reinforced composites. Compos. Part B Eng. 2019, 173, 106906. [Google Scholar] [CrossRef]
  106. Park, S.H.; Kim, C.; Yang, K.S. Preparation of carbonized fiber web from electrospinning of isotropic pitch. Synth. Met. 2004, 143, 175–179. [Google Scholar] [CrossRef]
  107. Minus, M.L.; Kumar, S. The processing, properties, and structure of carbon fibers. JOM 2005, 57, 52–58. [Google Scholar] [CrossRef]
  108. Morales-Asencio, J.M.; Gonzalo-Jiménez, E.; Martin-Santos, F.; Morilla-Herrera, J.; Celdráan-Mañas, M.; Carrasco, A.M.; García-Arrabal, J.; Toral-López, I. Effectiveness of a nurse-led case management home care model in Primary Health Care. A quasi-experimental, controlled, multi-centre study. BMC Health Serv. Res. 2008, 8, 193. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Tibbetts, G.; Lake, M.; Strong, K.; Rice, B. A review of the fabrication and properties of vapor-grown carbon nanofiber/polymer composites. Compos. Sci. Technol. 2007, 67, 1709–1718. [Google Scholar] [CrossRef]
  110. Ali Farshidfar, V.H.A.; Nazokdast, H. Electrical and Mechanical Properties Of Conducive Carbon Black/Polyolefin Composites Mixed With Carbon Fiber. In COMPOSITES 2006; Convention and Trade Show American Composites Manufacturers Association: St. Louis, MO, USA, 2006. [Google Scholar]
  111. Bryning, M.B.; Islam, M.; Kikkawa, J.M.; Yodh, A.G. Very Low Conductivity Threshold in Bulk Isotropic Single-Walled Carbon Nanotube-Epoxy Composites. Adv. Mater. 2005, 17, 1186–1191. [Google Scholar] [CrossRef]
  112. Ko, F.; Gogotsi, Y.; Ali, A.; Naguib, N.; Ye, H.; Yang, G.; Li, C.; Willis, P. Electrospinning of Continuous Carbon Nanotube-Filled Nanofiber Yarns. Adv. Mater. 2003, 15, 1161–1165. [Google Scholar] [CrossRef]
  113. Lima, M.D.; Fang, S.; Lepró, X.; Lewis, C.; Ovalle-Robles, R.; Carretero-González, J.; Castillo-Martínez, E.; Kozlov, M.E.; Oh, J.; Rawat, N.; et al. Biscrolling Nanotube Sheets and Functional Guests into Yarns. Science 2011, 331, 51–55. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Rosca, I.D.; Hoa, S.V. Highly conductive multiwall carbon nanotube and epoxy composites produced by three-roll milling. Carbon 2009, 47, 1958–1968. [Google Scholar] [CrossRef]
  115. Wang, X.; Zhi, L.; Müllen, K. Transparent, Conductive Graphene Electrodes for Dye-Sensitized Solar Cells. Nano Lett. 2008, 8, 323–327. [Google Scholar] [CrossRef]
  116. Xiang, C.; Lu, W.; Zhu, Y.; Sun, Z.; Yan, Z.; Hwang, C.-C.; Tour, J.M. Carbon Nanotube and Graphene Nanoribbon-Coated Conductive Kevlar Fibers. ACS Appl. Mater. Interfaces 2011, 4, 131–136. [Google Scholar] [CrossRef]
  117. Lee, T.-W.; Han, M.; Lee, S.-E.; Jeong, Y.G. Electrically conductive and strong cellulose-based composite fibers reinforced with multiwalled carbon nanotube containing multiple hydrogen bonding moiety. Compos. Sci. Technol. 2016, 123, 57–64. [Google Scholar] [CrossRef]
  118. Bilotti, E.; Zhang, R.; Deng, H.; Baxendale, M.; Peijs, T. Fabrication and property prediction of conductive and strain sensing TPU/CNT nanocomposite fibres. J. Mater. Chem. 2010, 20, 9449–9455. [Google Scholar] [CrossRef]
  119. Hooshmand, S.; Soroudi, A.; Skrifvars, M. Electro-conductive composite fibers by melt spinning of polypropylene/polyamide/carbon nanotubes. Synth. Met. 2011, 161, 1731–1737. [Google Scholar] [CrossRef]
  120. Zhu, X.-D.; Zang, C.-G.; Jiao, Q.-J. High electrical conductivity of nylon 6 composites obtained with hybrid multiwalled carbon nanotube/carbon fiber fillers. J. Appl. Polym. Sci. 2014, 131, 131. [Google Scholar] [CrossRef]
  121. Zhang, X.; Yan, X.; He, Q.; Wei, H.; Long, J.; Guo, J.; Gu, H.; Yu, J.; Liu, J.; Ding, D.; et al. Electrically Conductive Polypropylene Nanocomposites with Negative Permittivity at Low Carbon Nanotube Loading Levels. ACS Appl. Mater. Interfaces 2015, 7, 6125–6138. [Google Scholar] [CrossRef] [PubMed]
  122. Ra, E.J.; An, K.H.; Kim, K.K.; Jeong, S.Y.; Lee, Y.H. Anisotropic electrical conductivity of MWCNT/PAN nanofiber paper. Chem. Phys. Lett. 2005, 413, 188–193. [Google Scholar] [CrossRef]
  123. Miyauchi, M.; Miao, J.; Simmons, T.J.; Lee, J.-W.; Doherty, T.V.; Dordick, J.S.; Linhardt, R.J. Conductive Cable Fibers with Insulating Surface Prepared by Coaxial Electrospinning of Multiwalled Nanotubes and Cellulose. Biomacromolecules 2010, 11, 2440–2445. [Google Scholar] [CrossRef] [PubMed]
  124. Geetha, S.; Kumar, K.K.S.; Rao, C.R.; Vijayan, M.; Trivedi, D.C.K. EMI shielding: Methods and materials-A review. J. Appl. Polym. Sci. 2009, 112, 2073–2086. [Google Scholar] [CrossRef]
  125. Sekitani, T.; Noguchi, Y.; Hata, K.; Fukushima, T.; Aida, T.; Someya, T. A Rubberlike Stretchable Active Matrix Using Elastic Conductors. Science 2008, 321, 1468–1472. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Dalton, A.; Collins, S.; Muñoz, E.; Razal, J.; Ebron, V.H.; Ferraris, J.P.; Coleman, J.; Kim, B.G.; Baughman, R.H. Super-tough carbon-nanotube fibres. Nat. Cell Biol. 2003, 423, 703. [Google Scholar] [CrossRef]
  127. Shim, B.S.; Chen, W.; Doty, C.; Xu, C.; Kotov, N. Smart Electronic Yarns and Wearable Fabrics for Human Biomonitoring made by Carbon Nanotube Coating with Polyelectrolytes. Nano Lett. 2008, 8, 4151–4157. [Google Scholar] [CrossRef]
  128. Zhou, Z.; Chu, L.; Tang, W.; Gu, L. Studies on the antistatic mechanism of tetrapod-shaped zinc oxide whisker. J. Electrost. 2003, 57, 347–354. [Google Scholar] [CrossRef]
  129. Xue, C.-H.; Chen, J.; Yin, W.; Jia, S.-T.; Ma, J.-Z. Superhydrophobic conductive textiles with antibacterial property by coating fibers with silver nanoparticles. Appl. Surf. Sci. 2012, 258, 2468–2472. [Google Scholar] [CrossRef]
  130. Guo, R.H.; Jiang, S.-X.K.; Yuen, C.W.M.; Ng, M.C.F.; Lan, J.W. Optimization of electroless nickel plating on polyester fabric. Fibers Polym. 2013, 14, 459–464. [Google Scholar] [CrossRef]
  131. Moazzenchi, B.; Montazer, M. Click electroless plating of nickel nanoparticles on polyester fabric: Electrical conductivity, magnetic and EMI shielding properties. Colloids Surf. A Physicochem. Eng. Asp. 2019, 571, 110–124. [Google Scholar] [CrossRef]
  132. Lu, C.; Krifa, M.; Koo, J.H. Conductive Poly(3,4-Ethylenedioxythio-phene):Poly(4-styrene sulfonate) (PEDOT:PSS)/Nickel Nanostrands Nanocomposites. In SAMPE 2013; Society for the Advancement of Material and Process Engineering: Long Beach, CA, USA, 2013. [Google Scholar]
  133. Nagaraju, G.; Raju, G.S.R.; Ko, Y.H.; Yu, J.S. Hierarchical Ni–Co layered double hydroxide nanosheets entrapped on conductive textile fibers: A cost-effective and flexible electrode for high-performance pseudocapacitors. Nanoscale 2016, 8, 812–825. [Google Scholar] [CrossRef] [PubMed]
  134. Elmoubarki, R.; Mahjoubi, F.Z.; Elhalil, A.; Tounsadi, H.; Abdennouri, M.; Sadiq, M.; Qourzal, S.; Zouhri, A.; Barka, N. Ni/Fe and Mg/Fe layered double hydroxides and their calcined derivatives: Preparation, characterization and application on textile dyes removal. J. Mater. Res. Technol. 2017, 6, 271–283. [Google Scholar] [CrossRef]
  135. Lu, H.; Chen, J.; Tian, Q. Wearable high-performance supercapacitors based on Ni-coated cotton textile with low-crystalline Ni-Al layered double hydroxide nanoparticles. J. Colloid Interface Sci. 2018, 513, 342–348. [Google Scholar] [CrossRef] [PubMed]
  136. Conductive Composites. Nanostrands. 2011. Available online: http://www.conductivecomposites.com/nanostrands.html (accessed on 3 September 2012).
  137. Koecher, M.; Yeager, J.D.; Park, T.; Fullwood, D.; Colton, J.S.; Mara, N.; Hansen, N. Characterization of nickel nanostrand nanocomposites through dielectric spectroscopy and nanoindentation. Polym. Eng. Sci. 2013, 53, 2666–2673. [Google Scholar] [CrossRef]
  138. Krifa, M.; Prichard, C. Nanotechnology in textile and apparel research—An overview of technologies and processes. J. Text. Inst. 2020, 111, 1778–1793. [Google Scholar] [CrossRef]
  139. Hansen, N.; Hansen, G. From Inception to Insertion: Successful Products and Applications using Nickel Nanostrands. In SAMPE International Symposium; Society for the Advancement of Material and Process Engineering: Long Beach, CA, USA, 2011. [Google Scholar]
  140. Hansen, N.; Adams, D.O.; Fullwood, D.T. Quantitative methods for correlating dispersion and electrical conductivity in conductor–polymer nanostrand composites. Compos. Part A Appl. Sci. Manuf. 2012, 43, 1939–1946. [Google Scholar] [CrossRef]
  141. Hansen, N.; Adams, D.O.; Devries, K.L.; Goff, A.; Hansen, G. Investigation of Electrically Conductive Structural Adhesives using Nickel Nanostrands. J. Adhes. Sci. Technol. 2011, 25, 2659–2670. [Google Scholar] [CrossRef]
  142. Gangopadhyay, R.; De, A. Conducting Polymer Nanocomposites: A Brief Overview. Chem. Mater. 2000, 12, 608–622. [Google Scholar] [CrossRef]
  143. Morales, D.; Stoute, N.A.; Yu, Z.; Aspnes, D.E.; Dickey, M.D. Liquid gallium and the eutectic gallium indium (EGaIn) alloy: Dielectric functions from 1.24 to 3.1 eV by electrochemical reduction of surface oxides. Appl. Phys. Lett. 2016, 109, 091905. [Google Scholar] [CrossRef]
  144. Rotzler, S.; von Krshiwoblozki, M.; Schneider-Ramelow, M. Washability of e-textiles: Current testing practices and the need for standardization. Text. Res. J. 2021, 0040517521996727. [Google Scholar] [CrossRef]
  145. Rotzler, S.; Schneider-Ramelow, M. Washability of E-Textiles: Failure Modes and Influences on Washing Reliability. Text. Res. J. 2021, 1, 4. [Google Scholar] [CrossRef]
  146. Niu, B.; Yang, S.; Hua, T.; Tian, X.; Koo, M. Facile fabrication of highly conductive, waterproof, and washable e-textiles for wearable applications. Nano Res. 2021, 14, 1043–1052. [Google Scholar] [CrossRef]
  147. de Medeiros, M.S.; Goswami, D.; Chanci, D.; Moreno, C.; Martinez, R.V. Washable, breathable, and stretchable e-textiles wirelessly powered by omniphobic silk-based coils. Nano Energy 2021, 87, 106155. [Google Scholar] [CrossRef]
  148. Afroj, S.; Tan, S.; Abdelkader, A.M.; Novoselov, K.S.; Karim, N. Highly Conductive, Scalable, and Machine Washable Graphene-Based E-Textiles for Multifunctional Wearable Electronic Applications. Adv. Funct. Mater. 2020, 30, 2000293. [Google Scholar] [CrossRef] [Green Version]
  149. Molla, T.I.; Compton, C.; Dunne, L.E. Launderability of surface-insulated cut and sew E-textiles. In Proceedings of the 2018 ACM International Symposium on Wearable Computers, Singapore, 8–12 October 2018; pp. 104–111. [Google Scholar]
Figure 1. Illustration of a capacitive strain sensor (left), and capacitive response curves during loading and unloading (right). Reprinted with permission from [43]. Copyright 2018 American Chemical Society.
Figure 1. Illustration of a capacitive strain sensor (left), and capacitive response curves during loading and unloading (right). Reprinted with permission from [43]. Copyright 2018 American Chemical Society.
Textiles 01 00012 g001
Figure 2. Illustration of a piezoelectric pressure sensor at different load states (ad) and output voltage response for each state (e) and under varied pressure (f). Reprinted with permission from Springer Nature [50]. Copyright 2021 Springer Nature.
Figure 2. Illustration of a piezoelectric pressure sensor at different load states (ad) and output voltage response for each state (e) and under varied pressure (f). Reprinted with permission from Springer Nature [50]. Copyright 2021 Springer Nature.
Textiles 01 00012 g002
Figure 3. Schematic illustration of a planar microstrip patch antenna.
Figure 3. Schematic illustration of a planar microstrip patch antenna.
Textiles 01 00012 g003
Figure 4. Example of a textile-based implementation of the microstrip patch antenna design using copper tape as radiating patch and denim fabric as the substrate. Reprinted with permission from [67] under the Creative Commons Attribution License.
Figure 4. Example of a textile-based implementation of the microstrip patch antenna design using copper tape as radiating patch and denim fabric as the substrate. Reprinted with permission from [67] under the Creative Commons Attribution License.
Textiles 01 00012 g004
Figure 5. Example of a textile-based implementation of the microstrip patch antenna design using woven copper yarn as a radiating patch and woven glass fabric as the substrate. Reprinted with permission from [71] under the Creative Commons Attribution License.
Figure 5. Example of a textile-based implementation of the microstrip patch antenna design using woven copper yarn as a radiating patch and woven glass fabric as the substrate. Reprinted with permission from [71] under the Creative Commons Attribution License.
Textiles 01 00012 g005
Figure 6. Example of a textile-based implementation of the microstrip patch antenna design using embroidered Zari (silver metallic yarn wrapped on a silk core) as radiating patch. Reprinted with permission from [74]. Copyright 2019, Wiley Periodicals, Inc.
Figure 6. Example of a textile-based implementation of the microstrip patch antenna design using embroidered Zari (silver metallic yarn wrapped on a silk core) as radiating patch. Reprinted with permission from [74]. Copyright 2019, Wiley Periodicals, Inc.
Textiles 01 00012 g006
Figure 7. Conductivity ranges for different applications [78,79].
Figure 7. Conductivity ranges for different applications [78,79].
Textiles 01 00012 g007
Figure 8. Example conductivity levels of PANi and PPy based materials according to the literature.
Figure 8. Example conductivity levels of PANi and PPy based materials according to the literature.
Textiles 01 00012 g008
Figure 9. SEM images of (A) GNR-coated Kevlar fiber; the scale bar is 20 μm, (B) A MWCNT-coated Kevlar fiber; the scale bar is 40 μm; (C) SWCNT-coated Kevlar fiber, scale bar is 20 μm. Adapted with permission from [116]. Copyright 2011 American Chemical Society.
Figure 9. SEM images of (A) GNR-coated Kevlar fiber; the scale bar is 20 μm, (B) A MWCNT-coated Kevlar fiber; the scale bar is 40 μm; (C) SWCNT-coated Kevlar fiber, scale bar is 20 μm. Adapted with permission from [116]. Copyright 2011 American Chemical Society.
Textiles 01 00012 g009
Figure 10. Example conductivity levels of carbon-based materials reported in the literature.
Figure 10. Example conductivity levels of carbon-based materials reported in the literature.
Textiles 01 00012 g010
Figure 11. Three-dimensional branched structure of nickel nanostrands [132].
Figure 11. Three-dimensional branched structure of nickel nanostrands [132].
Textiles 01 00012 g011
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Krifa, M. Electrically Conductive Textile Materials—Application in Flexible Sensors and Antennas. Textiles 2021, 1, 239-257. https://0-doi-org.brum.beds.ac.uk/10.3390/textiles1020012

AMA Style

Krifa M. Electrically Conductive Textile Materials—Application in Flexible Sensors and Antennas. Textiles. 2021; 1(2):239-257. https://0-doi-org.brum.beds.ac.uk/10.3390/textiles1020012

Chicago/Turabian Style

Krifa, Mourad. 2021. "Electrically Conductive Textile Materials—Application in Flexible Sensors and Antennas" Textiles 1, no. 2: 239-257. https://0-doi-org.brum.beds.ac.uk/10.3390/textiles1020012

Article Metrics

Back to TopTop