Next Article in Journal
Numerical Simulation and Design of a High-Temperature, High-Pressure Fluid Transport Pipe
Next Article in Special Issue
Peroxidase-Like Metal-Based Nanozymes: Synthesis, Catalytic Properties, and Analytical Application
Previous Article in Journal
Thermal Performance Visualization Using Object−Oriented Physical and Building Information Modeling
Previous Article in Special Issue
Theranostic Nanoplatforms of Thiolated Reduced Graphene Oxide Nanosheets and Gold Nanoparticles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Development of Morphology Controlled Conducting Polymer Nanomaterial-Based Biosensor

1
School of Chemical Engineering, Yeungnam University, Gyeongsan 38541, Korea
2
Department of Materials Science and Engineering, Gachon University, 1342 Seongnam-Daero, Sujeong-Gu, Seongnam-Si, Gyeonggi-Do 13120, Korea
*
Author to whom correspondence should be addressed.
Appl. Sci. 2020, 10(17), 5889; https://doi.org/10.3390/app10175889
Submission received: 28 July 2020 / Revised: 22 August 2020 / Accepted: 24 August 2020 / Published: 25 August 2020
(This article belongs to the Special Issue World of Biosensing)

Abstract

:
Biosensors are of particular importance for the detection of biological analytes at low concentrations. Conducting polymer nanomaterials, which often serve as sensing transducers, are renowned for their small dimensions, high surface-to-volume ratio, and amplified sensitivity. Despite these traits, the widespread implementation of conventional conducting polymer nanomaterials is hampered by their scarcity and lack of structural uniformity. Herein, a brief overview of the latest developments in the synthesis of morphologically tunable conducting polymer-based biosensors is discussed. Research related to the dimensional (0, 1, 2, and 3D) hetero-nanostructures of conducting polymers are highlighted in this paper, and how these structures affect traits such as the speed of charge transfer processes, low-working temperature, high sensitivity and cycle stability are discussed.

1. Introduction

Biosensors are crucial for numerous applications in healthcare, medical science, agriculture, environmental monitoring, food, and biosecurity because of their extremely precise diagnostic and monitoring capabilities [1,2,3,4,5]. Analytes are detected with the aid of biological-sensing elements such as antibodies, cells, receptors, and aptamers, which are bound to transducers made from conductive nanomaterials [6,7,8]. Quantitative analysis of the target analyte is achieved by selectively converting the molecular recognition processes between the analyte and the biological element bound to the transducers from a nonelectrical domain to an electrical signal [9,10,11,12]. The binding of the analyte induces changes in the electrical properties of the transducers. An effective biosensor exhibits several crucial characteristics, namely, high sensitivity, good selectivity, fast response/recovery time, cycle stability, and the capability to operate at low temperatures. In addition to satisfying all of these prerequisites, conductive nanomaterial-based electrodes are renowned for their small size, high surface-to-volume ratio, and unique optical/electrical properties, thereby making them excellent candidates for biosensor-assisted applications [13,14,15,16,17].
Conducting polymers are often employed as transducer materials for biosensor applications because of the unique traits inferred by their unusual π-conjugated systems [18,19,20,21,22]. Conducting polymers contain polyconjugated chains consisting of alternating single (σ) and double (π) bonds that influence the electrical and optical properties of the sensing transducer [23,24,25]. Several parameters govern the physical properties of conducting polymers, including the length of the conjugation, the existing intra-/interchain interactions, and the extent of the disorder in the system [26,27]. Furthermore, research has shown that the partial or total inclusion of conducting polymer nanostructures in sensor transducers leads to significantly improved sensing performance [28,29,30,31,32]. The ability to selectively tune defects, electronic states, and the surface chemistry has been the driving force behind diverse methods for fabricating morphologically tunable conducting polymer nanomaterials [33,34,35,36]. Additionally, assembling conducting polymer nanostructures into an ordered array is often necessary to render them functional [37,38]. Novel nanostructure-based sensing devices exhibit enhanced abilities that are generated through the development of new strategies for combining and assembling nanoscale units [39,40,41,42].
Various synthetic methods have been reported, of which the template method is the most promising and powerful tool for fabricating conducting polymer nanomaterials [43,44,45,46,47,48,49,50,51]. The template method involves the inclusion of inorganic or organic guest molecules in the voids of the host material. These voids act as the “template”, thereby defining the shape, size, and orientation of the compound being produced. Generally, the template method is classified into three different types, namely, hard, soft, and template free methods. Briefly, the “hard” template process involves the fabrication of 1D nanostructures such as nanotubes and nanorods using anodic aluminum oxide membranes, track-etched polycarbonates, and zeolites [52,53,54,55]. The “soft” template method involves the fabrication of various morphologies of conducting polymer nanomaterials. Here, the templates used include surfactants, liquid–crystalline polymers, cyclodextrins, and functionalized polymers [56,57,58,59]. Surfactants that include cationic, anionic, and non-ionic amphiphiles are commonly used to form micelles as nanoreactors. Relative to the other two methods, which provide facile, practical routes for the production of pure, uniformly sized, high-quality nanofibers, the “template-free” method encompasses various techniques such as dispersion polymerization and electrochemical, chemical, aqueous/organic interfacial syntheses [60,61,62,63,64,65,66].
In this review, recent trends in the development of biosensor electrodes based on conducting polymer nanomaterials are presented. Additionally, the synthetic routes that facilitate the production of morphologically tunable nanostructures are closely examined.

2. Polypyrrole-Based Electrodes

Polypyrrole (PPy), a class of conducting polymers possessing five-membered heterocyclic rings, is renowned for their high electrical conductivity, reversible redox activity, and environmental stability. Ease of synthesis and tunable conductivity make PPy particularly useful for applications involving nanoscale architectures [67,68] as their nanostructures can be readily prepared via electrochemical and chemical routes in both aqueous and non-aqueous solutions [69,70]. Electrochemical polymerization is commonly used to fabricate deposited films on substrates, whereas chemical polymerization is used to produce powdered products. Structurally, the PPy chain is intrinsically planar and linear because of α–α coupling. However, conformational defects (e.g., α–α bonds with non-regular rotation) and structural defects (i.e., the presence of α–β bonds, hydroxyl moieties, and carbonyl groups) induce structural changes that facilitate their application for the production of biosensors [71,72].

2.1. 0D Nanostructures

As stated earlier, the popularity of PPy nanoparticles as biosensor sensing materials is due to the ability to incorporate various diverse receptors for analyte detection thanks to the large surface to volume ratio of these polymers. In particular, spherical PPy, which are most commonly used for producing biosensors, can be prepared via numerous methods, namely, microemulsion polymerization, self-assembly, and dispersion polymerization. Of the three methods listed, microemulsion polymerization is the most extensively utilized means of fabricating diverse nanometer-sized conducting polymer particles at low temperatures [73,74]. In contrast, the self-assembly approach takes advantage of the evaporation processes occurring in the liquid reaction medium [75]. Dispersion polymerization features the synthesis of sterically stabilized PPy colloids using a tailor-made reactive polymeric stabilizer [76,77]. Hong et al. described the preparation of PPy nanoparticles with a diameter range of 20–100 nm using FeCl3 in aqueous solutions containing PVA as a stabilizer [78]. Moreover, the generation of uniformly sized PPy nanoparticles for biosensor transducers was best conducted when the carboxylated polypyrrole (CPPy) chain was synthesized via the co-polymerization of pyrrole and 3-carboxyl pyrrole monomers [79].
Oh et al. reported introducing uniformly sized nanoparticles to the electrode’s surface using the spin coating technique to solve the issues of limited surface area caused by nanoparticle aggregation. Additionally, amide coupling between the particle of interest and the electrode’s surface induces greater electrode stability (Figure 1a) [79]. In Oh’s study, the human parathyroid hormone receptor (PTH), which is vital for maintaining calcium homeostasis in mammals, was incorporated into the close-packed CPPy nanoparticle-based electrode. Irregularities associated with internal PTH levels result in potentially fatal diseases such as adenoma, hyperplasia, osteoporosis, and cancer. Here, the sensor electrode was extremely sensitive to human PTH and exhibited high stability because of amide formation reactions between the nanoparticle and the electrode. Lee et al. incorporated platinum particles onto the surface of CPPy to fabricate nonenzymatic biosensors capable of detecting the neurotransmitter, dopamine (Figure 1b) [80]. In conjunction with platinum catalysts, this biosensor consistently detected dopamine at concentrations as low as 100 fM without the involvement of other biomaterials. Oh et al. recently manufactured a biosensor electrode that could detect the odor molecule, cadaverine, in the gas phase using nickel-attached PPy nanoparticles (Figure 1c) [81]. Here, the strong binding interaction between the olfactory receptor embedded nanoparticles, namely, TAAR13c-embedded nanodisk, and nickel was vital for cadaverine detection. In this study, the TAAR13c-embedded nanodisk was introduced stably onto the electrode coated with the nanoparticles of interest. This artificial olfactory biosensor was capable of detecting low concentrations of cadaverine in both liquid (100 aM) and gaseous (10 ppb) phases.

2.2. 1D Nanostructures

One-dimensional nanomaterials are popular sensing elements in analytical systems as they enhance the efficiency and performance of the associated biosensors via the rapid transfer of electrical signals and their large surface-to-volume ratio. These nanostructures significantly decrease the detection limit of biosensors by providing numerous immobilization sites, which, in turn, increases the number of biomolecular interactions [82]. Moreover, their high porosity facilitates fast penetration of the biomolecules, thereby reducing the detection time [83].
1D PPy-based nanomaterials are also synthesized using various manufacturing methods. Oh et al. synthesized human taste receptor-functionalized CPPy nanotubes to detect bitter compounds that produced phenylthiocarbamide and propylthiouracil [84]. CPPy nanotubes were fabricated as taste receptor templates using reverse microemulsion polymerization. Recently, Na et al. reported synthesizing CPPy nanotubes via a facile self-degradation method utilizing fibrillary templates composed of methyl orange and iron chloride (Figure 2a) [85]. Here, the carboxyl pyrrole monomer was incorporated into the fibrillar template via electrostatic interactions between the pyrrole monomer and the sulfite group of methyl orange. Simultaneously, there was a reaction between carboxyl pyrrole and Fe3+ cations on the surface of the fibrillar template. The synthesized CPPy nanotubes were then applied as sensing transducers to detect 17β-estradiol, a renowned inhibitor that affects hormone immunity and other physiological processes.
Kim et al. fabricated platinum-coated CPPy nanofibers using an oxidative template and sonochemical methods (Figure 2b) [86]. Here, CPPy nanofibers were formed when the electrode underwent charge transfer processes via an oxidative template composed of cetrimonium bromide and ammonium persulfate. The oxidative templates exhibited lamellar micelle structures and were composed of folded, twin-tailed complexes of the cetylammonium cations and peroxydisulfate anions. The unique micelle structure generated by the cetylammonium cation–peroxydisulfate anion interactions resulted in the polymerized nanofibers’ fibrous shape. Next, composite nanomaterials were formed using a simple chemical reduction of the platinum precursors. The platinum-coated CPPy nanofibers were useful for nonenzymatic biosensors to detect oxalic acid, a notable indicator of renal tubular acidosis, hyperparathyroidism, medullary sponge kidney, hyperoxaluria, and Dent’s diseases.

2.3. 2D Nanostructures

Two-dimensional nanomaterials are widely used as sensor electrodes as they exhibit uniformed structures in a wide surface area, making them ideal for the production of thin films. Most PPy-based films are easily synthesized using electrochemical deposition techniques and can be used to generate biosensor electrodes via the introduction of biological receptors [87,88]. Electrochemical deposition is a simple method used to obtain conducting polymer thin films containing electrolytes incorporated as dopants. Another class of 2D nanostructure films, namely, composite-based conducting polymers, can be employed for various functions. Arora et al. synthesized double-stranded calf thymus DNA-functionalized PPy-polyvinyl sulphonate thin-films using electrochemical deposition for detecting 2-aminoanthracene [87]. Ozer et al. fabricated polymer composite films consisting of PPy-dodecylbenzene sulphonate for the synthesis of cholesterol-sensing electrodes [88]. Weng et al. manufactured a composite ink made from PPy and polyvinyl alcohol mixtures that were then used to generate nanofilms for screen printing [89]. This conductive polymeric ink-based electrode was an effective biosensor for detecting glucose complexed glucose oxidase.

2.4. 3D Nanostructures

The detection capabilities of biosensor electrodes can be improved by maximizing the number of attached receptors per unit volume. This is achieved using 3D nanostructures possessing maximized surface areas as the biological receptor templates. Researchers have been actively involved in manufacturing 3D nanostructures using PPy for subsequent applications as biosensors.
As seen in one-dimensional structures, multidimensional PPy nanotubes with vertically coated bumps were manufactured using polyvinyl alcohol fibers as the mold via vapor deposition polymerization of the pyrrole monomer (Figure 3a) [68,90]. The 3D PPy nanotubes are ideal for monitoring cortisol hormone levels, thereby ensuring accurate regulation of blood pressure, glucose levels, and carbohydrate metabolism in response to environmental factors and the circadian rhythm. Relative to 2D nanofilms, 3D PPy films are manufactured via vapor deposition polymerization on the surface of the Co3O4 nanoplate attached perpendicular to the graphene electrode. The resulting thin film can be used to detect platelet-derived growth factor BB (PDGF BB), a bioindicator of blood cancer [91]. Cho et al. modified the electrochemical deposition technique to synthesize 3D PPy nanofilms with vertically coated nanorods on the electrode [92], thereby enabling the rapid transfer of electrical signals for sensing purposes. These nanofilms were ideal sensor electrodes for detecting HBsAg, a hepatitis B factor (Figure 3b). Relative to the 0D nanostructures discussed previously, urchin-like 3D PPy nanoparticles were fabricated using a combination of electrospray and vapor deposition polymerization of pyrrole (Figure 3c) [93]. Next, binding aptamers were introduced onto the particle’s surface to detect the environmental hormone, bisphenol A, as these aptamer-functionalized 3D nanoparticles were extremely sensitive environmental hormone sensors. The sensing performance of PPy-based biosensors are suggested in Table 1.

3. Polyaniline-Based Electrodes

The conductive states of polyaniline (PANI), which were first highlighted by Macdiarmid et al. are of research importance because of their redox behavior, simple doping/dedoping conditions, good environmental stability, facile synthesis, and lower synthetic costs compared to their counterparts [94,95]. PANI is particularly useful for analyte detection in a matrix of immobilized biomolecules, as it facilitates electron transfer processes in redox or enzymatic reactions. Therefore, PANI serves as a mediator for enzyme electrodes, as it undergoes redox cycling reactions and facilitates the direct coupling of enzyme active site electrons to the electrode’s surface. The size- and shape-dependent properties of PANI-based nanomaterials are crucial for fabricating high-performance sensors because of their unique charge-transport properties such as their doping and oxidation levels and the length of conjugation. These traits are strongly influenced by the shapes of the materials, whether they are nanofibers, nanowires, nanorods, nanotubes, or nanoparticles [96]. Extensive research is still needed to establish sustainable synthetic routes that enable researchers to adjust the morphology of the resulting PANI nanomaterials, thereby making them effective biosensor electrodes.

3.1. 0D Nanostructures

PANI-based nanoparticles, which are often available as pristine nanoparticles rather than functional coating materials, can be employed to improve binding attractions with the respective biomolecules. He et al. synthesized hollow carbon spheres coated with needle-like polyaniline to produce biosensor electrodes for the detection of malathion, an organophosphate insecticide (Figure 4a) [97]. Briefly, hollow carbon spheres were synthesized using the SiO2 template method, after which PANI was coated on the surface via in situ oxidative polymerization of the aniline within the hollow carbon sphere. Acetylcholine esterase was then immobilized on the PANI’s surface by the amide group in the chain. As reported by Salimian et al. Au nanoparticles were attached to PANI’s surface to enhance binding attractions with the HBV binding aptamer biomolecules [98]. Conversely, Al-Sagur et al. fabricated lutetium phthalocyanine containing silica–PANI conducting nanobeads without the need to etch the core particle (Figure 4b) [99]. In this case, the SiO2 core was vital for attaching glucose oxidase bioreceptors by introducing lutetium phthalocyanine. Thus, PANI served to interact with glucose oxidase and form thin films, thereby enabling glucose sensing.

3.2. 1D Nanostructures

One-dimensional PANI nanostructures are easily generated using various methods and exhibit excellent electrical signal transmission capacity. Lou et al. fabricated PANI nanowires using electrochemical polymerization, followed by the incorporation of Au nanoparticles onto the electrode’s surface (Figure 5a) [100]. This composite nanostructure was then bound to glucose oxidase for use as electrochemiluminescent glucose sensors. Additionally, Hui et al. coated polyethylene glycol (PEG) layers on the surface of PANI to enhance the antifouling properties of the nanofiber. The PEGlated PANI nanofibers were applied as breast cancer susceptibility gene aptamer sensors (Figure 5b) [101]. Additionally, Asmatuli et al. synthesized PANI-based nanofibers using a facile electrospinning process to obtain uniformly sized polymer nanofiber webs (Figure 5c) [102]. The nanofibers were then applied to detect cyclooxygenase-2 related to tumor cell proliferation, differentiation, pain, and adhesion. By contrast, PANI nanotubes were generated through a different technique. Wang et al. synthesized PANI nanotube-based glucose enzyme sensors using self-assembly and molecular imprinting methods [103]. Conversely, Soni et al. fabricated PANI nanotubes using the MnO2 nanofiber template to produce chronic myelogenous leukemia-sensing aptamer sensors [104].

3.3. 2D Nanostructures

Various methods were used to make PANI nanofilm-based biosensor electrodes with a large surface area. Oh et al. utilized the inkjet printing technique to create ammonium persulfate initiator solutions that were used to produce patterned PANI nanofilms via vapor deposition polymerization of aniline (Figure 6a) [105]. Inkjet printing is an efficient alternative to conventional photolithographic methods for the production of versatile CP-based micro- or nanoelectronics because of its cost-effectiveness, high-speed patterning capabilities, flexibility, morphological adjustability, and widespread applicability. The patterned PANI layer was applied for dopamine detection by monitoring the observed changes in the flowing current. Kaliasa et al. synthesized PANI nanosheet-based thin-films using the screen printing method to generate glucose electrochemical sensors [106]. Additionally, they synthesized NiO-coated PANI nanosheets using in situ polymerization processes to facilitate glucose detection (Figure 6b) [107]. Conversely, Liu et al. attached 2D PANI layers directly onto silicon substrates using dilute oxidative polymerization reactions to fabricate B-type natriuretic peptide (BNP) biomarker FET sensors [108]. Note also that electrochemical deposition was effective for manufacturing polyaniline nanofilms. Yun et al. generated electrodeposited 2D PANI layers on a gold electrode pattern for use as FET BNP antibody FET sensors (Figure 6c) [109].

3.4. 3D Nanostructures

Three-dimensional PANI nanostructures are most commonly obtained via the formation of hydrogels using crosslinking agents; this facilitates their application as diverse biosensor systems that are grounded in the working mechanisms governing the introduction of nanomaterials. Using amino trimethylene phosphonic acid as the crosslinking agent, Xu et al. added silver nanoparticles functionalized by a luminol derivative, N-(aminobutyl)-N-(ethylisoluminol) (ABEI), onto the surface of 3D PANI nanostructures to generate electrochemiluminescent sensors for xanthine detection (Figure 7a) [110]. Qing et al. employed phytic acid as the crosslinking agent to facilitate the introduction of photoactive materials such as poly {4,8-bis[5-(2-ethylhexyl)thiophen-2-yl]-benzo[1–2-b:4,5-b’]dithiophene-2,6-diyl-alt-3-fluoro-2-[(2-ethylhexyl)carbonyl]thieno[3–4-b]thi-ophene-4,6-diyl} (PTB7–Th) to the electrode’s surface for subsequent use as photoelectrochemical guanine sensors (Figure 7b) [111]. Li et al. also used 3D PANI derived from phytic acid as uric acid electrochemical biosensors [112].
Additionally, research has been conducted on the production of unique 3D composite materials via a simultaneous synthetic route using PANI and other materials. Soni et al. fabricated PANI–MoS2 nanoflower composites by taking advantage of the hydrothermal reaction between the PANI nanoneedle and the MoS2 precursor [113]. Here, the resulting nanocomposites were functionalized for chronic myelogenous leukemia (CML)-binding aptamers as biosensors for impedance spectroscopy. Lakshmi et al. made PANI-graphene composite microflowers via the in situ polymerization of aniline and graphene layers for use as cholesterol detecting biosensors [114]. The sensing performance of PANI-based biosensors are suggested in Table 2.

4. Poly(3,4-ethylenedioxythiophene)-Based Electrodes

Poly(3,4-ethylenedioxythiophene) (PEDOT), which is a polythiophene (PT) derivative, is an outstanding conducting polymer for practical applications because of its small bandgap, good optical properties, and high electrical conductivity compared to other conducting polymers [115]. PEDOT exhibits excellent electroactivity in phosphate buffers, making it ideal for biosensor applications, particularly in enzyme matrices for immobilization processes. Most importantly, its mode of action features the direct transfer of electrons during detection [116]. Given these findings, numerous studies have attempted to synthesize PEDOT nanostructures using different electrochemical and chemical oxidative polymerization methods.

4.1. 0D Nanostructures

The functional groups of PEDOT-based 0D materials facilitate the formation of composites containing various components and moieties in the polymer chain’s structure. Jiang et al. synthesized Pt-coated PEDOT nanospheres using a Pt precursor (H2PtCl4) as an oxidant during polymerization [117] for subsequent application as H2O2 non-enzyme sensors. Liu et al. fabricated Pt-coated PEDOT nanospheres using CaCO3 sphere template-assisted chemical polymerization to generate electrochemical glucose sensors (Figure 8a) [118]. Kim et al. synthesized MnO2-coated PEDOT nanoellipsoids by conducting reverse microemulsion reactions following reduction processes. The composite nanoellipsoids were then used to detect catechol amine in PC-12 cells (Figure 8b) [119]. Additionally, commercial PEDOT:PSS has been used to generate the associated composite nanoparticle biosensors. Wang et al. synthesized nanobeads composed of polyethylenimine binding with ferrocene (BPEI–FC) and PEDOT:PSS via chemical coupling reactions to create glucose biosensors (Figure 8c) [120]. Zhang et al. produced PEDOT:PSS-based micelle structures via electrostatic interactions with chitosan for subsequent application as H2O2 biosensors (Figure 8d) [121].

4.2. 1D Nanostructures

Park et al. synthesized carboxylated PEDOT nanotubes using microemulsion polymerization as templates for immobilizing human dopamine receptor nanovesicles [122] that were used to coat extremely sensitive PEDOT-based dopamine sensors. These sensors had a low limit of detection of 10 fM. Commercial PEDOT:PSS offers the advantage of easy processing, and is integral to the formation of 1D structures via numerous manufacturing methods. Jiang et al. generated PEDOT:PSS/silk fibroin core/shell wires for detecting ascorbic acid using wet-spinning and photo-crosslinking methods (Figure 9a) [123]. Wang et al. fabricated PEDOT:PSS nanowires via soft-lithography before introducing biotin-derivatized poly(L-lysine)-grafted oligo-ethylene glycol (PLL-g-OEG-Biotin) into the doped polymer chain [124]. The PEDOT:PSS-based nanowires were ideal electrochemical biosensors for detecting streptavidin. Moreover, PEDOT offers the advantage of expanded functionality since the surface of the 1D materials can be covered using PEDOT-coated PAN nanofibers via the VDP of EDOT. Next, the PAN/PEDOT core/shell nanofibers were employed as electrodes for electrochemical glucose sensors [125]. 1D PEDOT nanostructures were also fabricated using a simple electrochemical polymerization method; in this case, Chen et al. synthesized Au–hemoglobin–PEDOT nanowhisker composites for H2O2 detection (Figure 9b) [126].

4.3. 2D Nanostructures

Even though electrochemical polymerization is most commonly employed to manufacture conducting polymer films, the use of PEDOT tends to be more widespread since it results in uniformity in the large surface area and high stability. David et al. fabricated PEDOT:PSS-based “layer-by-layer” films consisting of PEDOT:PSS/chitosan/glucose oxidase + N-graphene molecules for glucose detection [127]. Briefly, PEDOT:PSS was obtained via electrodeposition method before the positively charged chitosan layer was adsorbed onto the negative PSS chain. A third layer consisting of glucose oxidase + N-graphene was introduced via physical adsorption onto the positively charged chitosan layer. Kim et al. generated Au nanoparticles coated with PEDOT films using a one-pot modified electrodeposition method [128]. Next, anti-vascular endothelial growth factor (VEGF) antibodies were added for detecting VEGF, which is associated with severe vision loss due to exudative age-related macular degeneration, advanced diabetic retinopathy, and retinal vascular occlusive disease. Moreover, Scotto et al. fabricated layer-by-layer PEDOT/graphite nanosheet-based film using simple spin-coating method [129]. In detail, EDOT and graphite composite solution was deposited by spin-coating process and then exposed oxidant (Fe (III) tosylate) with high temperature. Then, glucose oxidase (GOx) and osmium-modified polyallylamine (OsPA) were coated layer-by-layer on the films surface to detect glucose. In the sensing performance, incorporated graphite nanosheet in the PEDOT structure not only improves the voltammetric response of the resulting all-polymer electrodes but also produces a better integration of the electrochemically active supramolecular assembly allowing the effective glucose sensing. In addition, Wang et al. suggested low fouling electrochemical biosensor using zwitterionic polypeptide doted PEDOT film by simple electrodeposition process to detect breast cancer biomarker (BRCA1) detection [130]. Especially, the polypeptide in the film structure acted as a bifunctional material that not only inhibits the nonspecific adsorption of proteins but also allows for the usage of its carboxylic to fix DNA probes (terminated with amino group).

4.4. 3D Nanostructures

PEDOT-based 3D nanostructures can be manufactured by modifying conventional synthetic routes. Here, Jia et al. fabricated 3D Au–PEDOT–graphene aerogels using the in situ polymerization of EDOT on 3D graphene aerogels and the electrodeposition of Au on the surface [131]. This composite nanostructure is a highly sought after material with a 3D PEDOT-based surface via PEDOT coating of the 3D graphene skeleton. Antibodies for detecting prostate cancer factors were introduced on top of this material to generate highly sensitive biosensor electrodes. Meng et al. synthesized PEDOT nanofiber-based 3D networks using tetrabutylammonium perchlorate (TBAP) soft templates (Figure 10a) [132]. The nanofibers’ network structure possessed a more active surface area to facilitate the binding of lactate dehydrogenase. The group noted improved sensitivity of the electrode to the target analyte (i.e., lactate). Park et al. modified VDP of 3,4-ethylenedioxythiophene (EDOT) on the PMMA surface (Figure 10b) by controlling the temperature and pressure conditions of the reaction, thereby generating nanorods on the fibers’ surface [133]. The resulting 3D PEDOT nanofiber-based membranes were applied for dopamine detection through the immobilization of the human dopamine receptor (hDRD1). The sensing performance of PEDOT-based biosensors are suggested in Table 3.

5. Others

As noted throughout this review, researchers have taken advantage of the unique electrical and chemical properties offered by the alternative structures of conductive polymer chains to synthesize more effective biosensor electrodes. Soylemez et al. synthesized poly (2,5-di(furan-2-yl) thiazolo [5,4-d]thiazole) (PTTzFr)-based films using electrodeposition before introducing glucose oxidase for electrochemical glucose detection (Figure 11a) [134]. The excellent electrochromic properties of this polymer chain facilitate easy detection, as noted by the obvious color changes of the film material. One clear example of this can be seen in the study conducted by Liu et al. in which colorimetric thrombin aptasensors were generated using poly [3-(3′-N,N,N-triethylamino-1′-propyloxy)-4-methyl-2,5-thiophene hydrochloride] (PMNT) films via intrinsic peroxidase-like activities that catalyzed the reaction of the peroxidase substrate 3,3′,5,5′-tetramethylbenzidine (TMB) in the presence of hydrogen peroxide (H2O2). As a result, a notable blue color was observed (Figure 11b) [135]. Voccia et al. synthesized “layer-by-layer” films consisting of polythiophene/streptavidin/biotinylated captured probes to detect small RNA (e.g., miR-221 microRNA) biomarkers of lung and breast cancer cell lines [136]. The sensing performance of other conducting polymer-based biosensors are suggested in Table 4.

6. Conclusions and Outlook

Various conducting polymers are of interest for biosensor applications because of their excellent electrical, chemical, and physical properties. Researchers have utilized these properties to actively optimize the performance of these biosensors by controlling the structure of the conducting polymer in nanoscale such as 0D, 1D, 2D, and 3D. In the case of the 0D structure, most researches were conducted to generate uniformly sized conductive polymer nanoparticles using template-based methods (microemulsion and hard template) and then applied to the biosensor electrodes by attaching inorganic components or bio-receptors to the surface. In the case of 1D structure, the tube or fiber structure-based textiles were manufactured using methods such as electrospinning, electrodeposition, and hard template and applied them as a flexible biosensor electrode by attaching various receptors. Using simple electrodeposition, spin coating, and screen printing methods, the 2D structure manufactured a flexible, large-area film and used it as a conductive polymer film-based sensor electrode. Finally, in order to expand the active surface area of the nanomaterials to the analyte, various manufacturing methods were modified to generate the 3D-conducting polymer nanostructure, and applied as a highly reactive biosensor electrode material. Insights gained from this field will promote the development of flexible, wearable sensors that could have far-reaching consequences in the healthcare industry.

Author Contributions

The manuscript was written through contributions of all authors. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

This work was supported by the National Research Foundation of Korea (NRF) grant funded by Korea government (NRF-2018R1C1B5041562).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Nakajima, A. Application of Single-Electron Transistor to Biomolecule and Ion Sensors. Appl. Sci. 2016, 6, 94. [Google Scholar] [CrossRef]
  2. Wang, Y.; Nie, J.; Fang, W.; Yang, L.; Hu, Q.; Wang, Z.; Sun, J.Z.; Tang, B.Z. Sugar-Based Aggregation-Induced Emission Luminogens: Design, Structures, and Applications. Chem. Rev. 2020, 120, 4534–4577. [Google Scholar] [CrossRef] [PubMed]
  3. Jung, D.; Park, H.W.; Ma, G.; Lee, C.Y.; Kwon, T.; Han, J.H. Optimal Synthesis of Horizontally Aligned Single-Walled Carbon Nanotubes and Their Biofunctionalization for Biosensing Applications. J. Nanomater. 2016, 2016, 5140241. [Google Scholar] [CrossRef] [Green Version]
  4. Kim, K.; Seo, S. Flexible Pressure and Touch Sensor with Liquid Metal Droplet Based on Gallium Alloys. Mol. Cryst. Liq. Cryst. 2019, 685, 40–46. [Google Scholar] [CrossRef]
  5. Ye, H.; Kim, J.H.; Seo, S. Fabrication and Applications of Flexible Stacked Devices via Open-Hole Integration. Sci. Adv. Mater. 2017, 9, 1464–1467. [Google Scholar] [CrossRef]
  6. Yoon, H.; Jang, J. Conducting-Polymer Nanomaterials for High-Performance Sensor Applications: Issues and Challenges. Adv. Funct. Mater. 2009, 19, 1567–1576. [Google Scholar] [CrossRef]
  7. Islam, A.E.; Martineau, R.; Crasto, C.M.; Kim, H.; Rao, R.S.; Maruyama, B.; Kim, S.S.; Drummy, L.F. Graphene-Based Electrolyte-Gated Field-Effect Transistors for Potentiometrically Sensing Neuropeptide Y in Physiologically Relevant Environments. ACS Appl. Nano Mater. 2020, 3, 5088–5097. [Google Scholar] [CrossRef]
  8. Ban, D.K.; Liu, Y.; Wang, Z.; Ramachandran, S.; Sarkar, N.; Shi, Z.; Liu, W.; Karkisaval, A.G.; Martinez-Loran, E.; Zhang, F.; et al. Direct DNA Methylation Profiling with an Electric Biosensor. ACS Nano 2020, 14, 6743–6751. [Google Scholar] [CrossRef]
  9. Chung, Y.J.; Kim, J.; Park, C.B. Photonic Carbon Dots as an Emerging Nanoagent for Biomedical and Healthcare Applications. ACS Nano 2020, 14, 6470–6497. [Google Scholar] [CrossRef]
  10. Kim, S.; Lee, K.H.; Lee, J.Y.; Kim, K.K.; Choa, Y.H.; Lim, J.H. Single-Walled Carbon Nanotube-Based Chemi-Capacitive Sensor for Hexane and Ammonia. Electron. Mater. Lett. 2019, 15, 712–719. [Google Scholar] [CrossRef]
  11. Chang, W.-S.; Li, P.; Kakade, S.; Xiong, Y.; Shang, H.; Zhang, Y.; Lee, G.U. Rapid and Sensitive Detection of Cardiac Troponin I Using a Force Enhanced Immunoassay with Nanoporous Membrane. Nanoscale 2020, 12, 12568–12577. [Google Scholar] [CrossRef] [PubMed]
  12. Yoo, H.; Lee, S.B.; Lee, D.K.; Smits, E.C.P.; Gelinck, G.H.; Cho, K.; Kim, J.J. Top-Split-Gate Ambipolar Organic Thin-Film Transistors. Adv. Electron. Mater. 2018, 4, 1700536. [Google Scholar] [CrossRef]
  13. Xue, Q.; Li, Z.; Wang, Q.; Pan, W.; Chang, Y.; Duan, X. Nanostrip Flexible Microwave Enzymatic Biosensor for Noninvasive Epidermal Glucose Sensing. Nanoscale Horiz. 2020, 5, 934–943. [Google Scholar] [CrossRef] [PubMed]
  14. Zhang, Y.; Jiang, J.; Liu, Y.; Li, P.; Liu, Y.; Chen, L.; Zhao, J. Multi-Praseodymium-and-Tungsten Bridging Octameric Tellurotungstate and Its 2D Honeycomb Composite Film for Detecting Estrogen. Nanoscale 2020, 12, 10842–10853. [Google Scholar] [CrossRef] [PubMed]
  15. Song, S.; Lee, J.U.; Kang, J.; Park, K.H.; Sim, S.J. Real-Time Monitoring of Distinct Binding Kinetics of Hot-Spot Mutant p53 Protein in Human Cancer Cells Using an Individual Nanorod-Based Plasmonic Biosensor. Sens. Actuator B Chem. 2020, 322, 128584. [Google Scholar] [CrossRef]
  16. Yoo, H.; Hong, S.; Moon, H.; On, S.; Ahn, H.; Lee, H.K.; Kim, S.; Hong, Y.K.; Kim, J.J. Chemical Doping Effects on CVD-Grown Multilayer MoSe2 Transistor. Adv. Electron. Mater. 2018, 4, 1700639. [Google Scholar] [CrossRef]
  17. Sun, H.; Zhang, J.; Bai, W.; Li, Y. Electrogenerated Chemiluminescence Biosensing Method for DNA Hydroxymethylation Detection via Glycosylation and a New Multi-Functional ECL Signal Compound. Sens. Actuator B Chem. 2020, 322, 128582. [Google Scholar] [CrossRef]
  18. Jang, J. Conducting Polymer Nanomaterials and Their Applications. Adv. Polym. Sci. 2006, 199, 189–259. [Google Scholar]
  19. Kwon, O.S.; Song, H.S.; Park, T.H.; Jang, J. Conducting Nanomaterial Sensor Using Natural Receptors. Chem. Rev. 2019, 119, 36–93. [Google Scholar] [CrossRef]
  20. Yoo, H.; Ghittorelli, M.; Smits, E.C.P.; Gelinck, G.H.; Lee, H.K.; Torricelli, F.; Kim, J.J. Reconfigurable Complementary Logic Circuits with Ambipolar Organic Transistors. Sci. Rep. 2016, 6, 35585. [Google Scholar] [CrossRef]
  21. Roh, S.H.; Kim, J.; Park, W.I.; Kim, Y.D.; Lim, J.H. Fabrication and Characterization of Ge2Sb2Te5 Nanowire Arrays and PEDOT: PSS Hybrid Thermoelectric Composites. Korean J. Met. Mater. 2017, 55, 432–439. [Google Scholar]
  22. Yoo, H.; On, S.; Lee, S.B.; Cho, K.; Kim, J.J. Negative Transconductance Heterojunction Organic Transistors and their Application to Full-Swing Ternary Circuits. Adv. Mater. 2019, 31, 1808265. [Google Scholar] [CrossRef] [PubMed]
  23. Freudenberg, J.; Jänsch, D.; Hinkel, F.; Bunz, U.H.F. Immobilization Strategies for Organic Semiconducting Conjugated Polymers. Chem. Rev. 2018, 118, 5598–5689. [Google Scholar] [CrossRef] [PubMed]
  24. Das, G.S.; Shim, J.P.; Bhatnagar, A.; Tripathi, K.M.; Kim, T.Y. Biomass-Derived Carbon Quantum Dots for Visible-Light-Induced Photocatalysis and Label-Free Detection of Fe (iii) and Ascorbic Acid. Sci. Rep. 2019, 9, 15084. [Google Scholar] [CrossRef] [PubMed]
  25. Yoo, H.; Park, H.; Yoo, S.; On, S.; Seong, H.; Im, S.G.; Kim, J.J. Highly Stacked 3D Organic Integrated Circuits with via-Hole-Less Multilevel Metal Interconnects. Nat. Commun. 2019, 10, 2424. [Google Scholar] [CrossRef] [PubMed]
  26. Ibanez, J.G.; Rincon, M.E.; Gutierrez-Granados, S.; Chahma, M.; Jaramillo-Quintero, O.A.; Frontana-Uribe, B.A. Conducting Polymers in the Fields of Energy, Environmental Remediation, and Chemical−Chiral Sensors. Chem. Rev. 2018, 118, 4731–4816. [Google Scholar] [CrossRef]
  27. Chia, P.-J.; Chua, L.-L.; Sivaramakrishnan, S.; Zhuo, J.-M.; Zhao, L.-H.; Sim, W.-S.; Eo, Y.-C.; Ho, P.K.-H. Injection-induced De-doping in a Conducting Polymer During Device Operation: Asymmetry in the Hole Injection and Extraction Rates. Adv. Mater. 2007, 19, 4202–4207. [Google Scholar] [CrossRef]
  28. Babeli, I.; Ruano, G.; Casanovas, J.; Ginebra, M.-P.; Garcıa-Torres, J.; Aleman, C. Conductive, Self-Healable and Reusable Poly(3,4-ethylenedioxythiophene)-Based Hydrogels for Highly Sensitive Pressure Arrays. J. Mater. Chem. C 2020, 8, 8654–8667. [Google Scholar] [CrossRef]
  29. Vaidya, S.; Veselska, O.; Zhadan, A.; Daniel, M.; Ledoux, G.; Fateeva, A.; Tsuruoka, T.; Demessence, A. Flexible and Luminescent Fibers of a 1D Au(I)–Thiophenolate Coordination Polymer and Formation of Gold Nanoparticle-Based Composite Materials for SERS. J. Mater. Chem. C 2020, 8, 8018–8026. [Google Scholar] [CrossRef]
  30. Teng, Y.; Wei, J.; Du, H.; Mojtaba, M.; Li, D. A Solar and Thermal Multi-Sensing Microfiber Supercapacitor with Intelligent Self-Conditioned Capacitance and Body Temperature Monitoring. J. Mater. Chem. A 2020, 8, 11695–11711. [Google Scholar] [CrossRef]
  31. Kim, D.; Jung, D.; Noh, J.K.; Han, J.H. Graphene Chemocapacitive Sensors for Dimethyl Methylphosphonate Detection. Sci. Adv. Mater. 2018, 10, 1268–1273. [Google Scholar] [CrossRef]
  32. Fernandez, F.D.M.; Khadka, R.; Yim, J.-H. Highly Porous, Soft, and Flexible Vapor-Phase Polymerized Polypyrrole–Styrene–Ethylene–Butylene–Styrene Hybrid Scaffold as Ammonia and Strain Sensor. RSC Adv. 2020, 10, 22533–22541. [Google Scholar] [CrossRef]
  33. Puiggalí-Jou, A.; Valle, L.J.D.; Alemań, C. Encapsulation and Storage of Therapeutic Fibrin-Homing Peptides using Conducting Polymer Nanoparticles for Programmed Release by Electrical Stimulation. ACS Biomater. Sci. Eng. 2020, 6, 2135–2145. [Google Scholar] [CrossRef]
  34. Chen, S.; Kang, E.S.H.; Chaharsoughi, M.S.; Stanishev, V.; Kühne, P.; Sun, H.; Wang, C.; Fahlman, M.; Fabiano, S.; Darakchieva, V.; et al. Conductive Polymer Nanoantennas for Dynamic Organic Plasmonics. Nat. Nanotechnol. 2020, 15, 35–40. [Google Scholar] [CrossRef] [PubMed]
  35. Kim, S.G.; Jun, J.; Lee, J.S.; Jang, J. A Highly Sensitive Wireless Nitrogen Dioxide Gas Sensor Based on an Organic Conductive Nanocomposite Paste. J. Mater. Chem. A 2019, 7, 8451–8459. [Google Scholar] [CrossRef]
  36. Rajesh, M.; Manikandan, R.; Kim, B.C.; Becuwe, M.; Yu, K.H.; Raj, J. Electrochemical Polymerization of Chloride Doped PEDOT Hierarchical Porous Nanostructure on Graphite as a Potential Electrode for High Performance Supercapacitor. Electrochim. Acta 2020, 354, 136669. [Google Scholar] [CrossRef]
  37. Yang, S.; Liu, D.; Meng, Q.B.; Wu, S.; Song, X.-M. Reduced Graphene Oxide-Supported Methylene Blue Nanocomposite as a Glucose Oxidase-Mimetic for Electrochemical Glucose Sensing. RSC Adv. 2018, 8, 32565–32573. [Google Scholar] [CrossRef] [Green Version]
  38. Gao, Z.; Siow, K.S.; Chan, H.S.O. Self-Assembled Conducting Polymer Monolayers of Poly(3-octylthiophene) on Gold Electrodes. Synth. Met. 1995, 75, 5–10. [Google Scholar] [CrossRef]
  39. Taheri, N.; Alizadeh, N. Vertically Grown Nanosheets Conductive Polypyrrole as a Sorbent for Nanomolar Detection of Salicylic Acid. J. Pharm. Biomed. Anal. 2020, 188, 113365. [Google Scholar] [CrossRef]
  40. Kim, W.; Lee, J.S.; Jang, J. Aptamer-Functionalized Three-Dimensional Carbon Nanowebs for Ultrasensitive and Free-Standing PDGF Biosensor. ACS Appl. Mater. Interfaces 2020, 12, 20882–20890. [Google Scholar] [CrossRef]
  41. Wan, A.M.-D.; Inal, S.; Williams, T.; Wang, K.; Leleux, P.; Estevez, L.; Giannelis, E.P.; Fischbach, C.; Malliaras, G.G.; Gourdon, D. 3D Conducting Polymer Platforms for Electrical Control of Protein Conformation and Cellular Functions. J. Mater. Chem. B 2015, 3, 5040–5048. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Yuk, H.; Lu, B.; Lin, S.; Qu, K.; Xu, J.; Luo, J.; Zhao, X. 3D Printing of Conducting Polymers. Nat. Commun. 2020, 11, 1604. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Jang, J.; Oh, J.H.; Stucky, G.D. Fabrication of Ultrafine Conducting Polymer and Graphite Nanoparticles. Angew. Chem. Int. Ed. 2002, 41, 4016–4019. [Google Scholar] [CrossRef]
  44. Jang, J.; Yoon, H. Fabrication of Magnetic Carbon Nanotubes using a Metal-Impregnated Polymer Precursor. Adv. Mater. 2003, 15, 2088–2091. [Google Scholar] [CrossRef]
  45. Jang, J.; Lim, B. Facile Fabrication of Inorganic-Polymer Core–Shell Nanostructures by a One-Step Vapor Deposition Polymerization. Angew. Chem. Int. Ed. 2003, 42, 5600–5603. [Google Scholar] [CrossRef] [PubMed]
  46. Li, C.; Bai, H.; Shi, G. Conducting Polymer Nanomaterials: Electrosynthesis and Applications. Chem. Soc. Rev. 2009, 38, 2397–2409. [Google Scholar] [CrossRef] [PubMed]
  47. Wang, J.; Bunimovich, Y.L.; Sui, G.; Savvas, S.; Wang, J.; Guo, Y.; Heath, J.R.; Tseng, H.-R. Electrochemical Fabrication of Conducting Polymer Nanowires in an Integrated Microfluidic System. Chem. Commun. 2006, 19, 3075–3077. [Google Scholar] [CrossRef] [Green Version]
  48. Yang, T.; Liu, Y.; Wang, H.; Duo, Y.; Zhang, B.; Ge, Y.; Zhang, H.; Chen, W. Recent Advances in 0D Nanostructure-Functionalized Low-Dimensional Nanomaterials for Chemiresistive Gas Sensors. J. Mater. Chem. C 2020, 8, 7272–7299. [Google Scholar] [CrossRef]
  49. Jang, J.; Oh, J.H. Facile Fabrication of Photochromic Dye-Conducting Polymer Core-Shell Nanomaterials and Their Photoluminescence. Adv. Mater. 2003, 15, 977–980. [Google Scholar] [CrossRef]
  50. Lu, X.; Zhang, W.; Wang, C.; Wen, T.-C.; Wei, Y. One-Dimensional Conducting Polymer Nanocomposites: Synthesis, Properties and Applications. Prog. Polym. Sci. 2011, 36, 671–712. [Google Scholar] [CrossRef]
  51. He, H.-C.; Yan, A.-L.; Karapala, V.K.; Wang, S.-F.; Shen, M.-H.; Lin, Y.-L.; Chen, Y.-F.; Sugiyama, T.; Chen, J.-T. Laser-Assisted Nanowetting: Selective Fabrication of Polymer/Gold Nanorod Arrays Using Anodic Aluminum Oxide Templates. Macromol. Rapid Commun. 2020, 41, 2000035. [Google Scholar] [CrossRef] [PubMed]
  52. Pan, L.; Qiu, H.; Dou, C.; Li, Y.; Pu, L.; Xu, J.; Shi, Y. Conducting Polymer Nanostructures: Template Synthesis and Applications in Energy Storage. Int. J. Mol. Sci. 2010, 11, 2636–2657. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Dai, T.; Lu, Y. Water-Soluble Methyl Orange Fibrils as Versatile Templates for the Fabrication of Conducting Polymer Microtubules. Macromol. Rapid Commun. 2007, 28, 629–633. [Google Scholar] [CrossRef]
  54. Joo, S.; Kim, J.H.; Seo, S. Direct Fabrication of Electrochromic Devices with Complex Patterns on Three-Dimensional Substrates Using Polymeric Stencil Films. RSC Adv. 2017, 7, 43283–43288. [Google Scholar] [CrossRef] [Green Version]
  55. Nam, V.B.; Shin, J.; Yoon, Y.; Giang, T.T.; Kwon, J.; Suh, Y.D.; Yeo, J.; Hong, S.; Ko, S.H.; Lee, D. Highly Stable Ni-Based Flexible Transparent Conducting Panels Fabricated by Laser Digital Patterning. Adv. Funct. Mater. 2019, 29, 1806895. [Google Scholar] [CrossRef]
  56. Xia, L.; Wei, Z.; Wan, M. Conducting Polymer Nanostructures and Their Application in Biosensors. J. Colloid Interface Sci. 2010, 341, 1–11. [Google Scholar] [CrossRef]
  57. Long, Y.-Z.; Li, M.-M.; Gu, C.; Wan, M.; Duvail, J.-L.; Liu, Z.; Fan, Z. Recent Advances in Synthesis, Physical Properties and Applications of Conducting Polymer Nanotubes and Nanofibers. Prog. Polym. Sci. 2011, 36, 1415–1442. [Google Scholar] [CrossRef]
  58. Mulchandani, A.; Myung, N.V. Conducting Polymer Nanowires-Based Label-Free Biosensors. Curr. Opin. Biotechnol. 2011, 22, 502–508. [Google Scholar] [CrossRef]
  59. Jang, J.; Ha, J.; Kim, S. Fabrication of Polyaniline Nanoparticles Using Microemulsion Polymerization. Macromol. Res. 2007, 15, 154–159. [Google Scholar] [CrossRef]
  60. Zhou, F.; Chen, M.; Liu, W.; Liu, J.; Liu, Z.; Mu, Z. Fabrication of Positively Patterned Conducting Polymer Microstructures via One-Step Electrodeposition. Adv. Mater. 2003, 15, 1367–1370. [Google Scholar] [CrossRef]
  61. Jang, J.; Yoon, H. Multigram-Scale Fabrication of Monodisperse Conducting Polymer and Magnetic Carbon Nanoparticles. Small 2005, 1, 1195–1199. [Google Scholar] [CrossRef] [PubMed]
  62. Yang, J.; Martin, D.C. Microporous Conducting Polymers on Neural Microelectrode Arrays: II. Physical Characterization. Sens. Actuator A Phys. 2004, 113, 204–211. [Google Scholar] [CrossRef]
  63. Wan, M. A Template-Free Method towards Conducting Polymer Nanostructures. Adv. Mater. 2008, 20, 2926–2932. [Google Scholar] [CrossRef]
  64. Rahman, A.; Sanyal, M.K. Novel Switching Transition of Resistance Observed in Conducting Polymer Nanowires. Adv. Mater. 2007, 19, 3956–3960. [Google Scholar] [CrossRef]
  65. Seo, Y.D.; Jung, J.; Lee, H.; Yeo, J.; Hong, S.; Lee, P.; Lee, D. Nanowire Reinforced Nanoparticle Nanocomposite for Highly Flexible Transparent Electrodes: Borrowing Idea from Macrocomposites in Steel-Wire Reinforced Concrete. J. Mater. Chem. C 2017, 5, 791–798. [Google Scholar]
  66. Geethalakshmi, D.; Muthukumarasamy, N.; Balasundaraprabhu, R. CSA-Doped PANI Semiconductor Nanofilms: Synthesis and Characterization. J. Mater. Sci. Mater. Electron. 2015, 26, 7797–7803. [Google Scholar] [CrossRef]
  67. Cao, L.; Zhang, B.; Ou, X.; Wang, C.; Peng, C.; Zhang, J. Synergistical Coupling Interconnected ZnS/SnS2 Nanoboxes with Polypyrrole-Derived N/S Dual-Doped Carbon for Boosting High-Performance Sodium Storage. Small 2019, 15, 1804861. [Google Scholar] [CrossRef]
  68. Kwon, O.S.; Park, S.J.; Yoon, H.; Jang, J. Highly Sensitive and Selective Chemiresistive Sensors Based on Multidimensional Polypyrrole Nanotubes. Chem. Commun. 2012, 48, 10526–10528. [Google Scholar] [CrossRef]
  69. Jang, J.; Ha, H. Fabrication of Hollow Polystyrene Nanospheres in Microemulsion Polymerization Using Triblock Copolymers. Langmuir 2002, 18, 5613–5618. [Google Scholar] [CrossRef]
  70. Jang, J.; Yoon, H. Facile Fabrication of Polypyrrole Nanotubes Using Reverse Microemulsion Polymerization. Chem. Commun. 2003, 6, 720–721. [Google Scholar] [CrossRef]
  71. Plausinaitis, D.; Sinkevicius, L.; Samukaite-Bubniene, U.; Ratautaite, V.; Ramanavicius, A. Evaluation of Electrochemical Quartz Crystal Microbalance Based Sensor Modified by Uric Acid-Imprinted Polypyrrole. Talanta 2020, 220, 121414. [Google Scholar] [CrossRef]
  72. Colak, O.; Yasar, A.; Cete, S.; Arslan, F. Glucose Biosensor Based on the Immobilization of Glucose Oxidase on Electrochemically Synthesized Polypyrrole-Poly(vinyl sulphonate) Composite Film by Cross-Linking with Glutaraldehyde. Artif. Cells Blood Substit. Biotechnol. 2012, 40, 354–361. [Google Scholar] [CrossRef] [PubMed]
  73. Liu, Y.; Chu, Y.; Yang, L. Adjusting the Inner-Structure of Polypyrrole Nanoparticles through Microemulsion. Polymerization 2006, 98, 304–308. [Google Scholar] [CrossRef]
  74. Ovando-Medina, V.M.; Peralta, R.D.; Mendizábal, E.; Martínez-Gutiérrez, H.; Lara-Ceniceros, T.E.; Ledezma-Rodríguez, R. Synthesis of Polypyrrole Nanoparticles by Oil-in-Water Microemulsion Polymerization with Narrow Size Distribution. Colloid Polym. Sci. 2011, 289, 759–765. [Google Scholar] [CrossRef]
  75. Jang, J.; Oh, J.H. Novel Crystalline Supramolecular Assemblies of Amorphous Polypyrrole Nanoparticles through Surfactant Templating. Chem. Commun. 2002, 19, 2200–2201. [Google Scholar] [CrossRef]
  76. Zha, Z.; Yue, X.; Ren, Q.; Dai, Z. Uniform Polypyrrole Nanoparticles with High Photothermal Conversion Efficiency for Photothermal Ablation of Cancer Cells. Adv. Mater. 2013, 25, 777–782. [Google Scholar] [CrossRef]
  77. Men’shikova, A.Y.; Shabsel’s, B.M.; Evseeva, T.G. Synthesis of Polypyrrole Nanoparticles by Dispersion Polymerization. Russ. J. Appl. Chem. 2003, 76, 822–826. [Google Scholar]
  78. Hong, J.-Y.; Yoon, H.; Jang, J. Kinetic Study of the Formation of Polypyrrole Nanoparticles in Water-Soluble Polymer/Metal Cation Systems: A Light-Scattering Analysis. Small 2010, 6, 679–686. [Google Scholar] [CrossRef]
  79. Kwon, O.S.; Ahn, S.R.; Park, S.J.; Song, H.S.; Lee, S.H.; Lee, J.S.; Hong, J.-Y.; Lee, J.S.; You, S.A.; Yoon, H.; et al. Ultrasensitive and Selective Recognition of Peptide Hormone Using Close-Packed Arrays of hPTHR-Conjugated Polymer Nanoparticles. ACS Nano 2012, 6, 5549–5558. [Google Scholar] [CrossRef]
  80. Lee, J.S.; Oh, J.; Kim, S.G.; Jang, J. Highly Sensitive and Selective Field-Effect-Transistor NonEnzyme Dopamine Sensors Based on Pt/Conducting Polymer Hybrid Nanoparticles. Small 2015, 11, 2399–2406. [Google Scholar] [CrossRef]
  81. Oh, J.; Yang, H.; Jeong, G.E.; Moon, D.; Kwon, O.S.; Phyo, S.; Lee, J.; Song, H.S.; Park, T.H.; Jang, J. Ultrasensitive, Selective, and Highly Stable Bioelectronic Nose That Detects the Liquid and Gaseous Cadaverine. Anal. Chem. 2019, 91, 12181–12190. [Google Scholar] [CrossRef] [PubMed]
  82. Park, J.W.; Na, W.; Jang, J. One-Pot Synthesis of Multidimensional Conducting Polymer Nanotubes for Superior Performance Field Effect Transistor-Type Carcinoembryonic Antigen Biosensors. RSC Adv. 2016, 6, 14335–14343. [Google Scholar] [CrossRef]
  83. Ghanbari, K.; Bathaie, S.Z.; Mousavi, M.F. Electrochemically Fabricated Polypyrrole Nanofiber-Modified Electrode as a New Electrochemical DNA Biosensor. Biosens. Bioelectron. 2008, 23, 1825–1831. [Google Scholar] [CrossRef] [PubMed]
  84. Song, H.S.; Kwon, O.S.; Lee, S.H.; Park, S.J.; Kim, U.-K.; Jang, J.; Park, T.H. Human Taste Receptor-Functionalized Field Effect Transistor as a Human-Like Nanobioelectronic Tongue. Nano Lett. 2013, 13, 172–178. [Google Scholar] [CrossRef]
  85. Na, W.; Park, J.W.; An, J.H.; Jang, J. Size-Controllable Ultrathin Carboxylated Polypyrrole Nanotube Transducer for Extremely Sensitive 17β-Estradiol FET-Type Biosensors. J. Mater. Chem. B 2016, 4, 5025–5034. [Google Scholar] [CrossRef]
  86. Kim, W.; Lee, J.S.; Shin, D.H.; Jang, J. Platinum Nanoparticles Immobilized on Polypyrrole Nanofibers for Non-Enzyme Oxalic Acid Sensor. J. Mater. Chem. B 2018, 6, 1272–1278. [Google Scholar] [CrossRef]
  87. Arora, K.; Chaubey, A.; Singhal, R.; Singh, R.P.; Pandey, M.K.; Samanta, S.B.; Malhotra, B.D.; Chand, S. Application of Electrochemically Prepared Polypyrrole–Polyvinyl Sulphonate Films to DNA Biosensor. Biosens. Bioelectron. 2006, 21, 1777–1783. [Google Scholar] [CrossRef]
  88. Ozer, B.O.; Cete, S. Development of a Novel Biosensor Based on a Polypyrrole–Dodecyl benzene Sulphonate (PPy–DBS) Film for the Determination of Amperometric Cholesterol. Artif. Cell Nanomed. Biotechnol. 2017, 45, 824–832. [Google Scholar] [CrossRef] [Green Version]
  89. Weng, B.; Morrin, A.; Shepherd, R.; Crowley, K.; Killard, A.J.; Innis, P.C.; Wallace, G.G. Wholly Printed Polypyrrole Nanoparticle-Based Biosensors on Flexible Substrate. J. Mater. Chem. B 2014, 2, 793–799. [Google Scholar] [CrossRef]
  90. Jeong, G.; Oh, J.; Jang, J. Fabrication of N-Doped Multidimensional Carbon Nanofibers for High Performance Cortisol Biosensors. Biosens. Bioelectron. 2019, 131, 30–36. [Google Scholar] [CrossRef]
  91. Lee, J.S.; Kim, W.; Cho, S.; Jun, J.; Cho, K.H.; Jang, J. Multidimensional Hybrid Conductive Nanoplate-Based Aptasensor for Platelet-Derived Growth Factor Detection. J. Mater. Chem. B 2016, 4, 4447–4454. [Google Scholar] [CrossRef] [PubMed]
  92. Cho, K.H.; Shin, D.H.; Oh, J.; An, J.H.; Lee, J.S.; Jang, J. Multidimensional Conductive Nanofilm-Based Flexible Aptasensor for Ultrasensitive and Selective HBsAg Detection. ACS Appl. Mater. Interfaces 2018, 10, 28412–28419. [Google Scholar] [CrossRef] [PubMed]
  93. Lee, J.S.; Kim, S.G.; Jun, J.; Shin, D.H.; Jang, J. Aptamer-Functionalized Multidimensional Conducting-Polymer Nanoparticles for an Ultrasensitive and Selective Field-Effect-Transistor Endocrine-Disruptor Sensors. Adv. Funct. Mater. 2014, 24, 6145–6153. [Google Scholar] [CrossRef]
  94. Cho, S.; Lee, J.S.; Jun, J.; Jang, J. High-Sensitivity Hydrogen Gas Sensors Based on Pd-Decorated Nanoporous Poly(aniline-co-aniline-2-sulfonic acid):Poly(4-styrenesulfonic acid). J. Mater. Chem. A 2014, 2, 1955–1966. [Google Scholar] [CrossRef]
  95. Yang, L.; Yang, L.; Wu, S.; Wei, F.; Hu, Y.; Xu, X.; Zhang, L.; Sun, D. Three-Dimensional Conductive Organic Sulfonic Acid Co-Doped Bacterial Cellulose/Polyaniline Nanocomposite Films for Detection of Ammonia at Room Temperature. Sens. Actuator B Chem. 2020, 323, 128689. [Google Scholar] [CrossRef]
  96. Cochet, M.; Maser, W.K.; Benito, A.M.; Callejas, M.A.; Martínez, M.T.; Benoit, J.-M.; Schreiber, J.; Chauvet, O. Synthesis of a New Polyaniline/Nanotube Composite: “In-situ” Polymerisation and Charge Transfer Through Site-Selective Interaction. Chem. Commun. 2001, 16, 1450–1451. [Google Scholar] [CrossRef]
  97. He, L.; Cui, B.; Liu, J.; Song, Y.; Wang, M.; Peng, D.; Zhang, Z. Novel Electrochemical Biosensor Based on Core-Shell Nanostructured Composite of Hollow Carbon Spheres and Polyaniline for Sensitively Detecting Malathion. Sens. Actuator B Chem. 2018, 258, 813–821. [Google Scholar] [CrossRef]
  98. Salimian, R.; Shahrokhian, S.; Panahi, S. Enhanced Electrochemical Activity of a Hollow Carbon Sphere/Polyaniline-Based Electrochemical Biosensor for HBV DNA Marker Detection. ACS Biomater. Sci. Eng. 2019, 5, 2587–2594. [Google Scholar] [CrossRef]
  99. Al-Sagur, H.; Komathi, S.; Karakaş, H.; Atilla, D.; Gürek, A.G.; Basova, T.; Farmilo, N.; Hassan, A.K. A Glucose Biosensor Based on Novel Lutetium Bis-Phthalocyanine Incorporated Silica-Polyaniline Conducting Nanobeads. Biosens. Bioelectron. 2018, 102, 637–645. [Google Scholar] [CrossRef] [Green Version]
  100. Lou, F.; Lu, Z.; Hu, F.; Li, C.M. A 3D Bio-Platform Constructed by Glucose Oxidase Adsorbed on Au Nanoparticles Assembled Polyaniline Nanowires to Sensitively Detect Glucose by Electrochemiluminescence. J. Electroanal. Chem. 2017, 787, 125–131. [Google Scholar] [CrossRef] [Green Version]
  101. Hui, N.; Sun, X.; Niu, S.; Luo, X. PEGylated Polyaniline Nanofibers: Antifouling and Conducting Biomaterial for Electrochemical DNA Sensing. ACS Appl. Mater. Interfaces 2017, 9, 2914–2923. [Google Scholar] [CrossRef] [PubMed]
  102. Asmatulu, R.; Veisi, Z.; Uddin, M.N.; Mahapatro, A. Highly Sensitive and Reliable Electrospun Polyaniline Nanofiber Based Biosensor as a Robust Platform for COX-2 Enzyme Detections. Fibers Polym. 2019, 20, 966–974. [Google Scholar] [CrossRef]
  103. Wang, Q.; Xue, R.; Guo, H.; Wei, Y.; Yang, W. A Facile Horseradish Peroxidase Electrochemical Biosensor with Surface Molecular Imprinting Based on Polyaniline Nanotubes. J. Electroanal. Chem. 2018, 817, 184–194. [Google Scholar] [CrossRef]
  104. Soni, A.; Pandey, C.M.; Solanki, S.; Kotnala, R.K.; Sumana, G. Electrochemical Genosensor Based on Template Assisted Synthesized Polyaniline Nanotubes for Chronic Myelogenous Leukemia Detection. Talanta 2018, 187, 379–389. [Google Scholar] [CrossRef]
  105. Oh, W.-K.; Kim, S.; Shin, K.-H.; Jang, Y.; Choi, M.; Jang, J. Inkjet-Printed Polyaniline Patterns for Exocytosed Molecule Detection from Live Cells. Talanta 2013, 105, 333–339. [Google Scholar] [CrossRef]
  106. Kailasa, S.; Geeta, B.; Jayarambabu, N.; Reddy, R.K.K.; Sharma, S.; Rao, K.V. Conductive Polyaniline Nanosheets (CPANINS) for a Non-Enzymatic Glucose Sensor. Mater. Lett. 2019, 245, 118–121. [Google Scholar] [CrossRef]
  107. Kailasa, S.; Rani, B.G.; Reddy, M.S.B.; Jayarambabu, N.; Munindra, P.; Sharma, S.; Rao, K.V. NiO Nanoparticles-Decorated Conductive Polyaniline Nanosheets for Amperometric Glucose Biosensor. Mater. Chem. Phys. 2020, 242, 122524. [Google Scholar] [CrossRef]
  108. Liu, P.; Huang, J.; Sanchez, D.V.P.; Schwartzman, D.; Lee, S.H.; Yun, M. High Yield Two-Dimensional (2-D) Polyaniline Layer and Its Application in Detection of B-type Natriuretic Peptide in Human Serum. Sens. Actuator B Chem. 2016, 230, 184–190. [Google Scholar] [CrossRef] [Green Version]
  109. Liu, P.; Zhu, Y.; Lee, S.H.; Yun, M. Two-Dimensional Polyaniline Nanostructure to the Development of Microfluidic Integrated Flexible Biosensors for Biomarker Detection. Biomed. Microdevices 2016, 18, 113. [Google Scholar] [CrossRef]
  110. Xu, L.-H.; Li, J.-J.; Zeng, H.-B.; Zhang, X.-J.; Cosnier, S.; Marks, R.S.; Shan, D. ATMP-Induced Three-Dimensional Conductive Polymer Hydrogel Scaffold for a Novel Enhanced Solid-State Electrochemiluminescence Biosensor. Biosens. Bioelectron. 2019, 143, 111601. [Google Scholar] [CrossRef]
  111. Qing, M.; Chen, S.L.; Han, L.; Yang, Y.Z.; Luo, H.Q.; Li, N.B. Three–Dimensional Donor–Acceptor–Type Photoactive Material/Conducting Polyaniline Hydrogel Complex for Sensitive Photocathodic Enzymatic Bioanalysis. Biosens. Bioelectron. 2020, 158, 112179. [Google Scholar] [CrossRef] [PubMed]
  112. Li, L.; Wang, Y.; Pan, L.; Shi, Y.; Cheng, W.; Shi, Y.; Yu, G. A Nanostructured Conductive Hydrogels-Based Biosensor Platform for Human Metabolite Detection. Nano Lett. 2015, 15, 1146–1151. [Google Scholar] [CrossRef] [PubMed]
  113. Soni, A.; Pandey, C.M.; Pandey, M.K.; Sumana, G. Highly Efficient Polyaniline-MoS2 Hybrid Nanostructures Based Biosensor for Cancer Biomarker Detection. Anal. Chim. Acta 2019, 1055, 26–35. [Google Scholar] [CrossRef] [PubMed]
  114. Lakshmi, G.B.V.S.; Sharma, A.; Solanki, P.R.; Avasthi, D.K. Mesoporous Polyaniline Nanofiber Decorated Graphene Micro-Flowers for Enzyme-Less Cholesterol Biosensors. Nanotechnology 2016, 27, 345101. [Google Scholar] [CrossRef] [PubMed]
  115. Lee, S.; Gleason, K.K. Enhanced Optical Property with Tunable Band Gap of Cross-linked PEDOT Copolymers via Oxidative Chemical Vapor Deposition. Adv. Funct. Mater. 2015, 25, 85–93. [Google Scholar] [CrossRef]
  116. Park, J.; Kim, H.K.; Son, Y. Glucose Biosensor Constructed from Capped Conducting Microtubules of PEDOT. Sens. Actuator B Chem. 2008, 133, 244–250. [Google Scholar] [CrossRef]
  117. Jiang, F.; Yue, R.; Du, Y.; Xu, J.; Yang, P. A One-Pot ‘Green’ Synthesis of Pd-Decorated PEDOT Nanospheres for Nonenzymatic Hydrogen Peroxide Sensing. Biosens. Bioelectron. 2013, 44, 127–131. [Google Scholar] [CrossRef]
  118. Liu, Y.; Turner, A.P.F.; Zhao, M.; Mak, W.C. Processable Enzyme-Hybrid Conductive Polymer Composites for Electrochemical Biosensing. Biosens. Bioelectron. 2018, 100, 374–381. [Google Scholar] [CrossRef]
  119. Kim, S.; Oh, W.-K.; Jeong, Y.S.; Jang, J. Dual-Functional Poly(3,4-ethylenedioxy thiophene)/MnO2 Nanoellipsoids for Enhancement of Neurite Outgrowth and Exocytosed Biomolecule Sensing in PC12 Cells. Adv. Funct. Mater. 2013, 23, 1947–1956. [Google Scholar] [CrossRef]
  120. Wang, J.-Y.; Chen, L.-C.; Ho, K.-C. Synthesis of Redox Polymer Nanobeads and Nanocomposites for Glucose Biosensors. ACS Appl. Mater. Interfaces 2013, 5, 7852–7861. [Google Scholar] [CrossRef]
  121. Zhang, R.; Xu, X.; Fan, X.; Yang, R.; Wu, T.; Zhang, C. Application of Conducting Micelles Self-Assembled from Commercial Poly(3,4-ethylenedioxythiophene):Poly(styrene sulfonate) and Chitosan for Electrochemical Biosensor. Colloid Polym. Sci. 2018, 296, 495–502. [Google Scholar] [CrossRef]
  122. Park, S.J.; Song, H.S.; Kwon, O.S.; Chung, J.H.; Lee, S.H.; An, J.H.; Ahn, S.R.; Lee, J.E.; Yoon, H.; Park, T.H.; et al. Human Dopamine Receptor Nanovesicles for Gate-Potential Modulators in High-Performance Field-Effect Transistor Biosensors. Sci. Rep. 2015, 4, 4342. [Google Scholar] [CrossRef] [PubMed]
  123. Jiang, Y.; Xu, M.; Yadavalli, V.K. Silk Fibroin-Sheathed Conducting Polymer Wires as Organic Connectors for Biosensors. Biosensors 2019, 9, 103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Wang, Q.; Tang, N.; Zhang, M.; Xue, Q.; Duan, X. Highly Sensitive Protein Detection Using Conductive Polymer Nanowires Fabricated by Nanoscale Soft Lithography. In Proceedings of the 2018 IEEE Sensors, New Delhi, India, 28–31 October 2018; pp. 1–3. [Google Scholar]
  125. Çetin, M.Z.; Camurlu, P. An Amperometric Glucose Biosensor Based on PEDOT Nanofibers. RSC Adv. 2018, 8, 19724–19731. [Google Scholar] [CrossRef] [Green Version]
  126. Chen, Y.; Gai, P.; Jin, L.; Zhu, D.; Tian, D.; Abdel-Halim, E.S.; Zhang, J.; Zhu, J.-J. Fabrication of PEDOT Nanowhiskers for Electrical Connection of the Hemoglobin Active Center for H2O2 Electrochemical Biosensing. J. Mater. Chem. B 2013, 1, 3451–3457. [Google Scholar] [CrossRef]
  127. David, M.; Barsana, M.M.; Bretta, C.M.A.; Florescu, M. Improved Glucose Label-Free Biosensor with Layer-by-Layer Architecture and Conducting Polymer Poly(3,4-ethylenedioxythiophene). Sens. Actuator B Chem. 2018, 255, 3227–3234. [Google Scholar] [CrossRef]
  128. Kim, M.; Iezzi, R., Jr.; Shim, B.S.; Martin, D.C. Impedimetric Biosensors for Detecting Vascular Endothelial Growth Factor (VEGF) Based on Poly(3,4-ethylene dioxythiophene) (PEDOT)/Gold Nanoparticle (Au NP) Composites. Front. Chem. 2019, 7, 234. [Google Scholar] [CrossRef]
  129. Scotto, J.; Piccinini, E.; Bilderling, C.V.; Coria-Oriundoc, L.L.; Battaglini, F.; Knoll, W.; Marmisolle, W.A.; Azzaroni, O. Flexible Conducting Platforms Based on PEDOT and Graphite Nanosheets for Electrochemical Biosensing Applications. Appl. Surf. Sci. 2020, 525, 146440. [Google Scholar] [CrossRef]
  130. Wang, J.; Wang, D.; Hui, N. A Low Fouling Electrochemical Biosensor Based on the Zwitterionic Polypeptide Doped Conducting Polymer PEDOT for Breast Cancer Marker BRCA1 Detection. Bioelectrochemistry 2020, 136, 107595. [Google Scholar] [CrossRef]
  131. Jia, H.; Xu, J.; Lu, L.; Yu, Y.; Zuo, Y.; Tian, Q.; Li, P. Three-Dimensional Au Nanoparticles/ Nano-Poly(3,4-ethylene dioxythiophene)-Graphene Aerogel Nanocomposite: A High-Performance Electrochemical Immunosensing Platform for Prostate Specific Antigen Detection. Sens. Actuator B Chem. 2018, 260, 990–997. [Google Scholar] [CrossRef]
  132. Meng, L.; Turner, A.P.F.; Mak, W.C. Tunable 3D Nanofibrous and Bio-Functionalised PEDOT Network Explored as a Conducting Polymer-Based Biosensor. Biosens. Bioelectron. 2020, 159, 112181. [Google Scholar] [CrossRef] [PubMed]
  133. Park, S.J.; Lee, S.H.; Yang, H.; Park, C.S.; Lee, C.-S.; Kwon, O.S.; Park, T.H.; Jang, J. Human Dopamine Receptor-Conjugated Multidimensional Conducting Polymer Nanofiber Membrane for Dopamine Detection. ACS Appl. Mater. Interfaces 2016, 8, 28897–28903. [Google Scholar] [CrossRef] [PubMed]
  134. Soylemez, S.; Kaya, H.Z.; Udum, Y.A.; Toppare, L. A Multipurpose Conjugated Polymer: Electrochromic Device and Biosensor Construction for Glucose Detection. Org. Electron. 2019, 65, 327–333. [Google Scholar] [CrossRef]
  135. Liu, M.; Li, J.; Li, B. A Colorimetric Aptamer Biosensor Based on Cationic Polythiophene Derivative as Peroxidase Mimetics for the Ultrasensitive Detection of Thrombin. Talanta 2017, 175, 224–228. [Google Scholar] [CrossRef]
  136. Voccia, D.; Sosnowska, M.; Bettazzi, F.; Roscigno, G.; Fratini, E.; Franciscis, V.D.; Condorelli, G.; Chitta, R.; D’Souza, F.; Kutner, W.; et al. Direct Determination of Small RNAs Using a Biotinylated Polythiophene Impedimetric Genosensor. Biosens. Bioelectron. 2017, 87, 1012–1019. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. (a) Schematic of CPPy nanoparticle-based biosensor electrodes for the detection of peptide hormones [79]. Copyright 2012, American Chemical Society. (b) Platinum-coated CPPy nanoparticle-based non-enzyme biosensors for dopamine detection [80]. Copyright 2015, Wiley. (c) A nanodisk-functionalized bioelectronic nose based on nickel-coated CPPy nanoparticles [81]. Copyright 2019, American Chemical Society.
Figure 1. (a) Schematic of CPPy nanoparticle-based biosensor electrodes for the detection of peptide hormones [79]. Copyright 2012, American Chemical Society. (b) Platinum-coated CPPy nanoparticle-based non-enzyme biosensors for dopamine detection [80]. Copyright 2015, Wiley. (c) A nanodisk-functionalized bioelectronic nose based on nickel-coated CPPy nanoparticles [81]. Copyright 2019, American Chemical Society.
Applsci 10 05889 g001
Figure 2. (a) Synthesis of methyl orange–iron chloride template CPPy nanotubes for 17β-estradiol-based biosensors [85]. Copyright 2016, Royal Society of Chemistry. (b) Platinum-coated CPPy nanofibers for the production of nonenzymatic biosensors capable of detecting oxalic acid [86]. Copyright 2018, Royal Society of Chemistry.
Figure 2. (a) Synthesis of methyl orange–iron chloride template CPPy nanotubes for 17β-estradiol-based biosensors [85]. Copyright 2016, Royal Society of Chemistry. (b) Platinum-coated CPPy nanofibers for the production of nonenzymatic biosensors capable of detecting oxalic acid [86]. Copyright 2018, Royal Society of Chemistry.
Applsci 10 05889 g002
Figure 3. (a) Multidimensional PPy nanotubes based on electrospinning and vapor deposition polymerization techniques [68]. Copyright 2012, Royal Society of Chemistry. (b) 3D PPy-based conductive nanofilms for use as serum hepatitis B surface antigen biosensors [92]. Copyright 2018, American Chemical Society. (c) Urchin-like CPPy nanoparticle aptasensors for bisphenol A detection [93]. Copyright 2014, Wiley.
Figure 3. (a) Multidimensional PPy nanotubes based on electrospinning and vapor deposition polymerization techniques [68]. Copyright 2012, Royal Society of Chemistry. (b) 3D PPy-based conductive nanofilms for use as serum hepatitis B surface antigen biosensors [92]. Copyright 2018, American Chemical Society. (c) Urchin-like CPPy nanoparticle aptasensors for bisphenol A detection [93]. Copyright 2014, Wiley.
Applsci 10 05889 g003
Figure 4. (a) Hollow carbon spheres coated with needle-like PANI-based pesticide biosensors [97]. Copyright 2018, Elsevier. (b) Lutetium phthalocyanine incorporated silica–PANI nanobeads for use as glucose electrochemical sensors [99]. Copyright 2018, Elsevier.
Figure 4. (a) Hollow carbon spheres coated with needle-like PANI-based pesticide biosensors [97]. Copyright 2018, Elsevier. (b) Lutetium phthalocyanine incorporated silica–PANI nanobeads for use as glucose electrochemical sensors [99]. Copyright 2018, Elsevier.
Applsci 10 05889 g004
Figure 5. (a) Au nanoparticles incorporated into PANI nanowire-based glucose electrochemiluminescent biosensors [100]. Copyright 2017, Elsevier. (b) The electrochemical biosensors generated from PEGlated PANI nanofiber-based breast cancer susceptibility gene (BRCA1) [101]. Copyright 2017, American Chemical Society. (c) The synthesis of PANI nanofibers via electrospinning; these nanofibers were subsequently employed to manufacture cyclooxygenase-2 (COX-2) sensor transducers [102]. Copyright 2019, Springer.
Figure 5. (a) Au nanoparticles incorporated into PANI nanowire-based glucose electrochemiluminescent biosensors [100]. Copyright 2017, Elsevier. (b) The electrochemical biosensors generated from PEGlated PANI nanofiber-based breast cancer susceptibility gene (BRCA1) [101]. Copyright 2017, American Chemical Society. (c) The synthesis of PANI nanofibers via electrospinning; these nanofibers were subsequently employed to manufacture cyclooxygenase-2 (COX-2) sensor transducers [102]. Copyright 2019, Springer.
Applsci 10 05889 g005
Figure 6. (a) Patterned 2D PANI structures via inkjet printing and vapor deposition polymerization methods; these nanomolecules were useful for detecting neurotransmitters [105]. Copyright 2013, Elsevier. (b) NiO-coated PANI nanosheets via in situ polymerization to generate glucose biosensors [107]. Copyright 2020, Elsevier. (c) Flexible B-type natriuretic peptide biosensors based on electrodeposited 2D PANI nanostructures [109]. Copyright 2016, Springer.
Figure 6. (a) Patterned 2D PANI structures via inkjet printing and vapor deposition polymerization methods; these nanomolecules were useful for detecting neurotransmitters [105]. Copyright 2013, Elsevier. (b) NiO-coated PANI nanosheets via in situ polymerization to generate glucose biosensors [107]. Copyright 2020, Elsevier. (c) Flexible B-type natriuretic peptide biosensors based on electrodeposited 2D PANI nanostructures [109]. Copyright 2016, Springer.
Applsci 10 05889 g006
Figure 7. (a) 3D PANI hydrogel-based electrochemiluminescent sensors for xanthine detection [110]. Copyright 2019, Elsevier. (b) 3D PANI hydrogel-based guanine sensors that feature photoelectrochemical signals [111]. Copyright 2020, Elsevier.
Figure 7. (a) 3D PANI hydrogel-based electrochemiluminescent sensors for xanthine detection [110]. Copyright 2019, Elsevier. (b) 3D PANI hydrogel-based guanine sensors that feature photoelectrochemical signals [111]. Copyright 2020, Elsevier.
Applsci 10 05889 g007
Figure 8. (a) Pt nanoparticle-decorated–PEDOT microspheres using CaCo3 sphere template-assisted chemical polymerization reactions for glucose detection [118]. Copyright 2018, Elsevier. (b) PEDOT–MnO2 nanoellipsoids via reverse microemulsion reactions following reduction processes to detect catechol amine in PC-12 cells [119]. Copyright 2013, Wiley. (c) Polyethylenimine binding with ferrocene (BPEI–FC)/PEDOT:PSS nanobeads for electrochemical glucose biosensors [120]. Copyright 2013, American Chemical Society. (d) PEDOT:PSS/chitosan micelle-based H2O2 electrochemical sensors [121]. Copyright 2018, Springer.
Figure 8. (a) Pt nanoparticle-decorated–PEDOT microspheres using CaCo3 sphere template-assisted chemical polymerization reactions for glucose detection [118]. Copyright 2018, Elsevier. (b) PEDOT–MnO2 nanoellipsoids via reverse microemulsion reactions following reduction processes to detect catechol amine in PC-12 cells [119]. Copyright 2013, Wiley. (c) Polyethylenimine binding with ferrocene (BPEI–FC)/PEDOT:PSS nanobeads for electrochemical glucose biosensors [120]. Copyright 2013, American Chemical Society. (d) PEDOT:PSS/chitosan micelle-based H2O2 electrochemical sensors [121]. Copyright 2018, Springer.
Applsci 10 05889 g008
Figure 9. (a) PANI:PSS/silk fibroin core/shell wires derived via wet spinning and photo crosslinking reactions for use as ascorbic acid biosensors [123]. Copyright 2019, MDPI. (b) PEDOT nanowhiskers/Au nanoparticles–hemoglobin composite electrodes for H2O2 detection via electrodeposition [126]. Copyright 2013, Royal Society of Chemistry.
Figure 9. (a) PANI:PSS/silk fibroin core/shell wires derived via wet spinning and photo crosslinking reactions for use as ascorbic acid biosensors [123]. Copyright 2019, MDPI. (b) PEDOT nanowhiskers/Au nanoparticles–hemoglobin composite electrodes for H2O2 detection via electrodeposition [126]. Copyright 2013, Royal Society of Chemistry.
Applsci 10 05889 g009
Figure 10. (a) The 3D PEDOT nanofiber network from the tetrabutylammonium perchlorate (TBAP) soft template for lactate detection [132]. Copyright 2020, Elsevier. (b) Multidimensional PEDOT nanofiber membranes via a modified vapor deposition polymerization method for dopamine detection [133]. Copyright 2016, American Chemical Society.
Figure 10. (a) The 3D PEDOT nanofiber network from the tetrabutylammonium perchlorate (TBAP) soft template for lactate detection [132]. Copyright 2020, Elsevier. (b) Multidimensional PEDOT nanofiber membranes via a modified vapor deposition polymerization method for dopamine detection [133]. Copyright 2016, American Chemical Society.
Applsci 10 05889 g010
Figure 11. (a) The PTTzFr film-based electrochemical sensors for glucose detection [134]. Copyright 2019, Elsevier. (b) The poly [3-(3′-N,N,N-triethylamino-1′-propyloxy)-4-methyl-2,5-thiophene hydrochloride] PMNT-based colorimetric thrombin aptasensors [135]. Copyright 2017, Elsevier.
Figure 11. (a) The PTTzFr film-based electrochemical sensors for glucose detection [134]. Copyright 2019, Elsevier. (b) The poly [3-(3′-N,N,N-triethylamino-1′-propyloxy)-4-methyl-2,5-thiophene hydrochloride] PMNT-based colorimetric thrombin aptasensors [135]. Copyright 2017, Elsevier.
Applsci 10 05889 g011
Table 1. Comparative sensing performance and limit of detection of polypyrrole nanostructure-based biosensors.
Table 1. Comparative sensing performance and limit of detection of polypyrrole nanostructure-based biosensors.
NanostructureReceptorAnalyteMechanismLOD 1Response TimeWorking TemperatureLinearityCycle StabilityReference
CPPyNP 2hPTHRhTPH 3FET 448 fM<10 s25 °C48 fM–480 pM2 weeks[79]
Pt_CPPyNP-DopamineFET100 fM<10 s25 °C0.1 pM–1 nM 4 weeks[80]
Ni-CPPyNPTAAR13c 5-embedded nanodiskCadaverineFET100 aM<5 s25 °C0.1 fM–100 µM5 weeks[81]
CPPyNThTAS2R38 6PTC 7 PROP 8FET1 fM 10 fM<5 s25 °C1 fM–1 µM1 week[84]
CPPyNT17β-estradiol binding aptamer17β-estradioFET1 fM<10 s25 °C1 fM–1 nM4 weeks[85]
Pt_CPPyNF-Oxalic acidFET100 fM<10 s25 °C10 fM–100 pM8 weeks[86]
PPy-PVS 9 filmCT-dsDNA 102-aminoantharcene o-chlorophenolAmperometric0.01 ppm 0.1 ppm<30 s25 °C0.01–20 ppm 0.1–30 ppm-[87]
PPy-DBS 11 filmChOx 12CholesterolAmperometric0.11 µM-25 °C0.11 µM–1.9 mM30 days[88]
PPy-PVAHRP 13 GoD 14H2O2 GlucoseAmperometric10 µM 1 mM<10 s25 °C10 µM–10 mM 1–5 mM-[89]
3D PPyNFF4P1A3 15CortisolFET100 aM<5 s25 °C100 aM–10 nM30 days[90]
3D CPPy plate-based filmPDGF-B binding aptamerPDGF-BB 16FET1.78 fM<10 s25 °C1.78 fM–17.8 pM4 weeks[91]
3D PPy filmHBsAg-binding aptamerHBsAg 17FET10 aM<10 s25 °C10 aM–0.1 µM500 cycles[92]
Urchin-like CPPyNPBPA binding aptamerBPA 18FET1 fM<10 s25 °C1 fM–10 pM4 weeks[93]
1 Limit of detection, 2 carboxylated polypyrrole nanoparticle, 3 human parathyroid hormone, 4 field effect transistor, 5 olfactory receptor, 6 bitter taste receptor, 7 phenylthiocarbamide, 8 propylthiouracil, 9 polypyrrole-polyvinyl sulphonate, 10 double stranded calf thymus DNA, 11 polypyrrole-dodecyl benzene sulphonate, 12 cholesterol oxidase, 13 horseradish peroxidase, 14 glucose oxidase, 15 cortisol antibody, 16 serum hepatis B surface antigen, 17 platelet-derived growth factor, 18 bisphenol A.
Table 2. Comparative sensing performance and limit of detection of polyaniline nanostructure-based biosensors.
Table 2. Comparative sensing performance and limit of detection of polyaniline nanostructure-based biosensors.
NanostructureReceptorAnalyteMechanismLOD 1Response TimeWorking TemperatureLinearityCycle StabilityReference
HCS@PANI 2AChE 3MalationElectrochemical2.15 pM-25 °C1.0 ng mL−1–10 µg mL−1-[97]
HCS-PANI 4Thiolated-probe DNAHepatitis B VirusElectrochemical3.62 fM-25 °C10 fM–1 nM-[98]
SiO2 (LuPc2) PANI (PVIA)-CNB 5GOxGlucoseElectrochemical0.1 mM<2 s25 °C1–16 mM45 days[99]
Au/PANI NWGOxGlucoseECL 60.05 µM-25 °C0.1–100 µM 40 cycles[100]
PEGlated PANI NFAptamerBRCA1 7Electrochemical3.8 fM-25 °C0.01 pM–1 nM-[101]
Electrospun PANI NFsCOX-2 polyclonal antibodyCOX-2 enzyme 8Electrochemical0.01 pg mL−1-25 °C0.01 pg mL−1–1 µg mL−1-[102]
PANI NTsHRP 9H2O2Electrochemical8.1 fM<200 s25 °C0.01–90 µM-[103]
PANI NTsAptamerCML 10Electrochemical0.1 fM-25 °C0.1 fM–1 µM 40 days[104]
2D PANIRGD 11DopamineElectrical2 nM-25 °C2 nM–1 µM48 h[105]
PANI nanosheet-GlucoseElectrochemical0.043 µM<1 s25 °C1 µM–1 mM-[106]
NiO/PANI nanosheets-GlucoseElectrochemical0.06 µM<10 s25 °C1 µM–1 mM60 days[107]
2D PANI layerAntibodyBNT 12FET 1350 pg mL−1-25 °C50–1000 pg mL−1-[108]
2D PANIAntibodyBNTFET100 pg mL−1-25 °C100–1000 pg mL−1-[109]
3D PANI hydrogel/Ag/ABEIXanthine oxidaseXanthineECL9.6 nM-25 °C0.01–200 µM-[110]
3D PANI hydrogel/PtB7-ThXanthine oxidaseGuaninePEC 140.02 µM<2 s25 °C0.1–80 µM2 days[111]
3D PANI hydrogel/PtUricaseUric acidElectrochemical1 µM<3 s25 °C0.07–1 mM-[112]
3D PANI-MoS2 nanoflowerAptamerCMLElectrochemical3 aM-25 °C0.01 fM–1 µM-[113]
PANI/graphene microflower-CholesterolElectrochemical1.93 mg dL−1-25 °C1.93–464.04 mg mL−1-[114]
1 Limit of detection, 2 hollow carbon spheres coated with needle-like polyaniline, 3 acetylcholine esterase, 4 hollow carbon spheres decorated with polyaniline, 5 lutetium phthalocyanine incorporated silica-polyaniline conducting nanobeads, 6 electrochemical luminescence, 7 breast cancer susceptibility gene, 8 cyclooxygenase-2, 9 horseradish peroxidase, 10 chronic myelogenous leukemia, 11 arginine-glycine-aspartate peptide, 12 B-type natriuretic peptide, 13 field effect transistor, 14 photoelectrochemical.
Table 3. Comparative sensing performance and limit of detection of PEDOT nanostructure-based biosensors.
Table 3. Comparative sensing performance and limit of detection of PEDOT nanostructure-based biosensors.
NanostructureReceptorAnalyteMechanismLOD 1Response TimeWorking TemperatureLinearityCycle StabilityReference
Pt/PEDOT nanosphere-H2O2Electrochemical2.84 µM<10 s25 °C2.5 µM–mM3 weeks[117]
Pt/PEDOT microsphereGOxGlucoseElectrochemical1.15 µM<10 s25 °C0.1–10 mM12 days[118]
PEDOT/MnO2 nanoellipsod-Catechol amineElectrical0.25 mM<20 s25 °C0.25–25 mM-[119]
BPEI-FC 2/PEDOT:PSS nanobeadGOxGlucoseElectrochemical2.4 mM<20 s25 °C0.5–5 mM-[120]
PEDOT:PSS/CS micelleHRPH2O2Electrochemical30 pM-25 °C0.1 nM–10 µM35 days[121]
PEDOT NThDRD1 3DopamineFET 410 fM<1 s25 °C10 pM–10 nM -[122]
PEDOT:PSS-silk fibroin core-sheath wires-Ascorbic acidElectrical1.14 µM<20 s25 °C1.14–800 µM4 weeks[123]
PEDOT:PSSPLL 5-g-OEG 6-BiotinStreptavidinElectrochemical1 fg mL−1-25°C1–1000 fg mL−1-[124]
PEDOT NFGOxGlucoseElectrochemical2.9 µM<3 s25 °C2.9 µM–25 mM 25 cycles[125]
PEDOT nanowhiskerAu-Hb 7H2O2Electrochemical0.6 µM<20 s25 °C1 µM–1.1 mM60 days[126]
PEDOT:PSS-based LBL filmGOxGlucoseElectrochemical41 µM<20 s25 °C0.1–1.4 mM-[127]
PEDOT/AuAntibodyVEGF 8Electrochemical0.5 pg mL−1-25 °C1–20 pg mL−1-[128]
Au-PEDOT-Graphene AerogelAntibodyPSA 9Electrochemical0.03 pg mL−1-25 °C0.1 pg mL−1–50 ng mL−1-[131]
3D PEDOT NF networkLDH 10LactateElectrochemical0.05 mM<10 s25 °C0.05–1.8 mM-[132]
3D PEDOT NF membranehDRD1DopamineFET100 fM<2 s25 °C0.1–100 pM-[133]
1 Limit of detection, 2 polyethylenimine binding with ferrocene, 3 human dopamine receptor nanovesicle, 4 field effect transistor, 5 biotin-derivatized poly(L-lysine), 6 grafted oligo-ethylene glycol, 7 hemoglobin, 8 vascular endothelial growth factor, 9 prostate specific antigen, 10 lactate dehydrogenase.
Table 4. Comparative sensing performance and limit of detection of other conducting polymer nanostructure-based biosensors.
Table 4. Comparative sensing performance and limit of detection of other conducting polymer nanostructure-based biosensors.
NanostructureReceptorAnalyteMechanismLOD 1Response TimeWorking TemperatureLinearityCycle StabilityReference
PTTzFr 2 filmGOx 3GlucoseElectrochromic12.8 µM<10 s25 °C12.8–500 µM-[134]
PMNT 4 filmAptamerThrombinColorimetric4 pM-25 °C0.01–0.1 nM-[135]
PT-based LBL 5 filmBiotinmiR-221Electrochemical0.7 pM-25 °C0.7–100 pM-[136]
1 Limit of detection, 2 poly(2,5-di(furan-2-yl) thiazolo[5,4-d]thiazole), 3 glucose oxidase, 4 poly[3-(3′-N,N,N-triethylamino-1′-propyloxy)-4-methyl-2,5-thiophene hydrochloride], 5 polythiophene-based layer-by-layer.

Share and Cite

MDPI and ACS Style

Cho, S.; Lee, J.S. Recent Development of Morphology Controlled Conducting Polymer Nanomaterial-Based Biosensor. Appl. Sci. 2020, 10, 5889. https://doi.org/10.3390/app10175889

AMA Style

Cho S, Lee JS. Recent Development of Morphology Controlled Conducting Polymer Nanomaterial-Based Biosensor. Applied Sciences. 2020; 10(17):5889. https://doi.org/10.3390/app10175889

Chicago/Turabian Style

Cho, Sunghun, and Jun Seop Lee. 2020. "Recent Development of Morphology Controlled Conducting Polymer Nanomaterial-Based Biosensor" Applied Sciences 10, no. 17: 5889. https://doi.org/10.3390/app10175889

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