0% found this document useful (0 votes)
8 views

Polymers 16 01706

Uploaded by

alfox0510
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
8 views

Polymers 16 01706

Uploaded by

alfox0510
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 17

Article

Electrochemical Investigation of PEDOT:PSS/Graphene Aging


in Artificial Sweat
Boriana Tzaneva 1,*, Valentin Mateev 2, Bozhidar Stefanov 1, Mariya Aleksandrova 3 and Ivo Iliev 4

1 Department of Chemistry, Faculty of Electrical Engineering and Technology, Technical University of Sofia,
Kliment Ohridski Blvd., 8, 1000 Sofia, Bulgaria; [email protected]
2 Department of Electrical Apparatus, Faculty of Electronic Engineering, Technical University of Sofia,

Kliment Ohridski Blvd., 8, 1000 Sofia, Bulgaria; [email protected]


3 Department of Microelectronics, Faculty of Electronic Engineering and Technology, Technical University of

Sofia, Kliment Ohridski Blvd., 8, 1000 Sofia, Bulgaria; [email protected]


4 Department of Electronics, Faculty of Electronic Engineering and Technology, Technical University of Sofia,

Kliment Ohridski Blvd., 8, 1000 Sofia, Bulgaria; [email protected]


* Correspondence: [email protected]

Abstract: Herein, we investigate the potential application of a composite consisting of PE-


DOT:PSS/Graphene, deposited via spray coating on a flexible substrate, as an autonomous conduct-
ing film for applications in wearable biosensor devices. The stability of PEDOT:PSS/Graphene is
assessed through electrochemical impedance spectroscopy (EIS), cyclic voltammetry (CV) and lin-
ear polarization (LP) during exposure to an artificial sweat electrolyte, while scanning electron mi-
croscopy (SEM) was employed to investigate the morphological changes in the layer following
these. The results indicate that the layers exhibit predominant capacitive behavior in the potential
range of −0.3 to 0.7 V vs. Ag/AgCl, with a cut-off frequency of approximately 1 kHz and retain 90%
capacity after 500 cycles. Aging under exposure to air for 6 months leads only to a minor increase in
impedance, demonstrating potential for storage under non-demanding conditions. However, pro-
longed exposure (>48 h) to the artificial sweat causes significant degradation, resulting in an imped-
ance increase of over 1 order of magnitude. The observed degradation raises important considera-
Citation: Tzaneva, B.; Mateev, V.; tions for the long-term viability of these layers in wearable biosensor applications, prompting the
Stefanov, B.; Aleksandrova, M.; need for additional protective measures during prolonged use. These findings contribute to ongoing
Iliev, I. Electrochemical Investigation efforts to enhance the stability and reliability of conducting materials for biosensors in health care
of PEDOT:PSS/Graphene Aging in and biotechnology applications.
Artificial Sweat. Polymers 2024, 16,
1706. https://doi.org/10.3390/ Keywords: aging; conductive polymer; spray coating; artificial sweat; electrochemical stability
polym16121706

Academic Editor: Hongsheng Yang

Received: 10 May 2024


1. Introduction
Revised: 9 June 2024
Accepted: 12 June 2024
The interest in conducting polymers for electrode fabrication on flexible substrates is
Published: 14 June 2024 driven by substantial advantages, including excellent mechanical properties and easy ap-
plication through various techniques, including printing on textiles, paper, and flexible
polymer substrates [1,2]. Some conductive polymers, such as poly(3,4-ethylenedioxythio-
phene):poly(styrenesulfonate) (PEDOT:PSS), offer additional benefits with proven bio-
Copyright: © 2024 by the authors.
Licensee MDPI, Basel, Switzerland.
compatibility, high electrochemical stability, high charge capacity, and commercial avail-
This article is an open access article
ability, making them highly suitable for in vitro, in vivo, and wearable sensors [3–8]. In
distributed under the terms and these systems, the conjugated polymer PEDOT provides electrical conductivity, while PSS
conditions of the Creative Commons conveys ionic transport, contributing to an overall capacitive behavior comparable to that
Attribution (CC BY) license of conventional metal and carbon-based electrodes [9]. PEDOT:PSS is recognized as a
(https://creativecommons.org/license leading organic electrode material, capable of replacing conventional carbon, metal (Au,
s/by/4.0/). Cu, Pt, etc.), and conductive glass-based electrodes, offering direct electron transfer at

Polymers 2024, 16, 1706. https://doi.org/10.3390/polym16121706 www.mdpi.com/journal/polymers


Polymers 2024, 16, 1706 2 of 17

lower cost and process temperatures. Additionally, the intrinsic brittleness in metallic con-
ductive layers, due to their high Young’s modulus, makes them unsuitable for flexible and
stretchable electronics, and thus incompatible with soft biological tissues [10]. The fabri-
cation of PEDOT:PSS films has been explored through various coating techniques, includ-
ing dip coating, spin coating, and spray coating. For cost-effective application of large-
area thin films over surfaces with a complex geometry, spray coating emerges as a viable
option. This deposition technique is based on spraying the PEDOT:PSS solution onto a
substrate, employing a nozzle at a precise distance and specific temperature [11]. While
this route easily affords an uniform coating, some challenges related to surface wettability
and substrate adhesion may arise. The technological and functional advantages of PE-
DOT:PSS-based layers have accelerated research on their properties over the last decade.
The effects on morphology, mechanical properties, and conductivity of PEDOT:PSS for
various deposition methods and conditions is often explored in the literature [12]. Charge
transfer in intrinsically conducting polymers like PEDOT:PSS is likely influenced by both
intra-grain and inter-grain structures due to their granular nature. The inter-grain struc-
ture is affected by grain elongation and stacking, primarily determined by the deposition
conditions, especially temperature. The amount of the secondary dopant and its removal
during the evaporation and annealing phases can impact the final alignment of grains
across the film [13]. Favorable alignment of PEDOT:PSS grains, particularly minimizing
the overall density of PSS regions along the charge pathway, contributes to the film con-
ductivity. Consequently, smoother samples with horizontally aligned grains are expected
to exhibit higher conductivity than the pristine sample at most temperatures.
Literature data on the electrochemical stability of these conductive layers, however,
are not unequivocal. Some authors report a reduction in their charging ability with con-
tinuous cycling [14,15], while others suggest that PEDOT:PSS films remain stable over one
hundred cycles [2,7]. Polymer materials are known to be susceptible to hydrolytic and
oxidative processes attacking specific bonds in their structure. As a hydrophilic polymer,
PSS is more sensitive to hydrolysis, as water molecules and ions can more freely penetrate
its internal structure. Nevertheless, some long-term tests demonstrate remarkable stability
of electropolymerized PEDOT:PSS films from 104 to 109 bipolar input pulses [8,16,17].
The long-term stability of the layers is directly related to their adhesion to the sub-
strate. In most studies, conductive substrates such as metallic (Au, Pt, Ag), carbon, or ITO
layers are used [2,18–20]. On some substrates, especially with thicker electrodeposition of
PEDOT:PSS, cracking and delamination from the electrode surface have been observed
due to weak adhesive forces and a lack of chemical interaction between them [2,18,19]. For
instance, Mousavi et al. [2] demonstrated that PEDOT:PSS exhibits the highest adhesion
to organic substrates with low conductivity and high roughness. Applying PEDOT:PSS
coatings on flexible polymer substrates has also been achieved following their prior met-
allization [6] and various intermediate adhesion layers have been proposed to improve
adhesion [8,17].
It is notable that electrochemical tests on PEDOT:PSS-based layers are usually con-
ducted in neutral phosphate buffer saline (PBS) solution [2,8,16,17], and in some cases in
media based on NaCl [12,21], KCl [22], or K4Fe(CN)6 solutions [23], albeit less commonly.
Notably, tests in environments simulating human sweat are not found in the literature,
despite the growing application of these polymers in wearable biosensors.
Conductive polymer-based electrodes are considered practical when they exhibit low
and frequency-independent impedance over a broad frequency range [24], a characteristic
that might be enhanced through PEDOT:PSS modification with various additives. E.g.,
Gold–silver core–shell (Au–Ag) nanoparticles have been incorporated into PEDOT:PSS
layers for paraoxon-ethyl detection, demonstrating a synergistic effect resulting in higher
conductivity and improved electrochemical stability [25]. However, this approach is com-
plex and costly, prompting exploration of simpler composites as alternatives. PEDOT:PSS
also acts as a surfactant for non-covalent functionalization of exfoliated graphene, which
Polymers 2024, 16, 1706 3 of 17

is employed for the preparation of hybrid inks with mixed conductivity [26]. The conju-
gated aromatic chains of PEDOT:PSS strongly anchor onto the graphene surface, due to
π–π interactions, without disrupting the electronic structure of graphene [26]. Combining
PEDOT:PSS with graphene allows for the modulation of electrical conductivity and ma-
terial functionality. These hybrid conductive inks PEDOT:PSS/Graphene, can be applied
conveniently through spray-coating method. The resulting films exhibit approximately
three orders of magnitude lower resistance compared to pure PEDOT:PSS [27], while
maintaining high transparency, chemical and thermal stability, stretchability, and low
contact resistance with organic materials [26]. Additionally, these polymer inks enable
precise and cost-effective deposition of conductive polymer layers on various substrates,
including flexible polymer substrates without the need for prior metallization [1,27,28].
Most reported application procedures of PEDOT:PSS/Graphene films include solar cells and
light-emitting diodes [29]. However, these structures lack well-defined electrochemical
characterization for wearable biosensing devices. Simultaneously, PEDOT:PSS-based poly-
mer layers exhibit pronounced hydrophilic properties, suggesting a significant impact of the
surrounding environment on their electrochemical characteristics. The promising conduc-
tivity and stability of the PEDOT:PSS/Graphene composite make it suitable for formation of
conductive patterns on dielectric substrates for biomedical applications. Clarifying both
their electrochemical stability and resilience in a simulated environment of human sweat is
crucial for incorporating conductive pattern made of PEDOT:PSS/Graphene layers into var-
ious wearable sensor structures, especially in electrochemical sensors.
This paper explores the potential application of a composite layer made of PE-
DOT:PSS/Graphene ink deposited through spray coating on polyethylene terephthalate
(PET) as an independent conductive pattern material. Stability of PEDOT:PSS/Graphene
is investigated through electrochemical impedance spectroscopy (EIS), linear polarization
(LP) and cyclic voltammetry (CV) during aging in ambient conditions and exposure to
artificial sweat electrolyte for nine days. Morphological changes in the layer post-tests are
examined using scanning electron microscopy (SEM). This study is a continuation of our
previous work [30] where we elucidated the influence of the spray deposition modes
(spray pressure, temperature, and number of passes) of PEDOT:PSS/Graphene ink on the
thickness, sheet resistance, and roughness of the layers as well as on its electrochemical
behavior in the artificial sweat. Here, we emphasize the aging and the long-term stability
of the hydrophilic PEDOT:PSS/Graphene composite ink, choosing the layer with the
smallest thickness (of 1 µm). Electrochemical studies of PEDOT:PSS/Graphene on insulat-
ing PET substrate provide advantages by avoiding interference at the polymer/metal in-
terface and potential issues like electrolyte penetration, delamination, and noise introduc-
tion from this interface to the useful signal. To address flexibility and substrate insulation
issues, we propose a test structure allowing electrochemical analysis of conducting poly-
mer electrodes using conventional electrochemical techniques. Additionally, assessing the
stability of polymer layers upon exposure to artificial sweat and extended air storage con-
tributes to defining the applicability boundaries of PEDOT:PSS/Graphene layers as elec-
trodes for routine applications.

2. Materials and Methods


2.1. Materials
The Graphene/PEDOT:PSS hybrid ink used in the present study was purchased as a
finished product from Sigma Aldrich—Merck KGaA (Darmstadt, Germany). According
to the production certificate, the conductive ink is prepared by dispersing 0.2 mg mL−1
PEDOT:PSS and 1 mg mL−1 electrochemically exfoliated graphene in dimethylformamide
(DMF). Sodium chloride (NaCl, ≥99.0%, CAS №: 7647-14-5), ammonium chloride (NH4Cl,
≥99.5%, CAS №: 12125-02-9), acetic acid (≥99.8%, CAS №: 64-19-7), D,L-lactic acid (CAS
№: 50-21-5) were used for artificial sweat preparation and were purchased from AlfaAesar
(Karlsruhe, Germany).
Polymers 2024, 16, 1706 4 of 17

2.2. Fabrication
The PET flexible (25 × 25 mm squares, 270 µm thick), underwent a cleaning process
by sonication in isopropyl alcohol, followed by UV–Ozone surface treatment (λ = 265 nm,
P = 350 W). Samples were fabricated employing a spray coating setup HS-AS18CK Ha-
osheng (Ningbo, China) equipped with a regulating nozzle (HS-30 0.01–0.1 mm diameter
range) and a heated substrate holder. The substrate temperature was maintained at 100
°C. The layers were deposited with an aerosol pressure of 3.5 mbar at a constant nozzle-
to-substrate distance of 12 cm and 5 spray passes. The pre-heated substrate was exposed
to the aerosol flow in 5-second cycles, allowing solvent evaporation and preventing mate-
rial leakage.

2.3. Material Characterization


Raman spectra of PEDOT:PSS/Graphene ink were obtained on a TO-ERS-532 spec-
trometer (Thunder Optics, Montpellier, France), based on a 532 nm laser source and a
Raman probe, equipped with a ×20 microscope objective lens. Approx. 50 µL of the ink
suspension were drop-casted on a glass substrate and dried at 80 °C for 15 min to remove
the solvent. The recorded spectrum is presented in Figure 1. In the Raman spectrum of
nanocomposite ink, the characteristic D, G and 2D peaks of graphene are seen at 1350,
1588 cm–1 and 2700 cm–1, respectively [27,31]. The peaks at ~1440 cm–1 and ~1500 cm–1 can
be assigned to C=C symmetric and C=C asymmetric stretch in PEDOT:PSS, respectively
[27,31,32]. The Raman spectrum of ink dropped on PET is a hybrid spectrum on which
both the peaks of the ink and those of the PET substrate can be observed (Figure S1a),
while the spectrum collected on tested sample with a thin PEDOT:PSS/Graphene layer
sprayed on PET presents only the strong peaks of PET substrate (Figure S1b).

Figure 1. Raman spectra of PEDOT:PSS/Graphene ink.

The sprayed PEDOT:PSS/Graphene film achieves thickness of 1.00 ± 0.05 µm meas-


ured by Alpha Step 100 (KLA-Tencor GmbH, Dresden, Germany). The obtained compo-
site layers were found to have a sheet resistance of 111.6 ± 6 Ω sq−1, as estimated by Veeco
FPP-5000 4 Point Prober (Veeco Instruments Inc., Plainview, NY, USA) and roughness of
9.2 ± 0.7 nm, as estimated by atomic force microscopy (Nanosurf, FlexAFM, Liestal, Swit-
zerland). Scanning electron microscopy (SEM) analysis was performed on a Hitachi
TM4000 (Hitachi, Tokyo, Japan) table-top microscope at accelerating voltage of 15 kV in
the back-scattered electrons (BSE) mode and by Tescan LYRA (Brno, Czech Republic) at
accelerating voltage of 20 kV.
Polymers 2024, 16, 1706 5 of 17

2.4. Electrochemical Measurements


Electrochemical impedance spectroscopy (EIS), linear polarization (LP), and cyclic
voltammetry (CV) were employed to investigate the electrochemical stability and aging
in artificial sweat. A potentiostat–galvanostat PalmSens4 (PalmSens BV, Houten, Nether-
lands) equipped with a frequency response analyzer and PSTrace 5.9 software were used.
For electrochemical tests, a model artificial sweat solution with a composition of 20 g L−1
NaCl, 17.5 g L−1 NH4Cl, 5 g L−1 acetic acid, and 15 g L−1 DL-lactic acid was used [33,34]. A
fresh electrolyte was prepared before each test from stock solutions of the individual com-
ponents, and the pH was adjusted to 4.7 using NaOH. The stock solutions were stored at
a temperature of 5 °C.
The electrochemical tests were conducted at a temperature of 25 °C in a three-elec-
trode cell, as shown in Figure 2. The working electrode was a sample with a nanocompo-
site PEDOT:PSS/Graphene layer deposited on PET with an exposed area of 0.1 cm2. A Pt-
plate (1 cm2) and Ag/AgCl electrode (3.0 M KCl, 0.21 V vs. SHE) were used as the counter
and reference electrodes, respectively. All potentials are presented versus the Ag/AgCl
electrode. EIS tests were performed at an open-circuit potential with an amplitude of 10
mV in a frequency range from 100 kHz to 10 mHz. The linear polarization (LP) was used
for the electrochemical loading of the composite layer during aging tests in the electrolyte
from −0.15 V to 1.0 V vs. Ag/AgCl at a scan rate of 1 mV s−1. CV with scan rates from 0.01
to 1.0 V was employed to determine the capacity of the tested layer. A total of 500 cycles
from −0.3 to 0.7 V with a potential scan rate of 0.1 V s−1 were conducted to investigate the
electrochemical stability of layer.
In the employed electrochemical cell design (Figure 2), the conductive polymer layer
contacts a metal current collector from the front side, connected to the potentiostat. The
current collector was fabricated from aluminum foil (38 µm thickness and 99.99% Al pu-
rity) with a 5 mm diameter opening, uniformly positioned around a silicone seal (1 mm
gap), determining the path of current lines and the main resistance for charge transfer
from the electrolyte/polymer boundary to the potentiostat. The voltametric characteristic
of the aluminum-conductive polymer contact in the range from −1 to 1 V is linear, exhib-
iting a 7.7 mA V−1 slope in both potential scanning directions, thus indicating a pure ohmic
contact between the two conductors. The small thickness of the aluminum foil allows its
maximum approach to the electrolyte without hindering sealing with the silicone seal.
Nevertheless, the 1 mm distance between the current collector and the electrolyte chamber
can be considered sufficiently large to exclude the possibility of the resultant electrochem-
ical signal being influenced by electrolyte penetration through the polymer to aluminum
boundary and deterioration of the contact at this boundary, a phenomenon encountered
in other arrangements of the current collector [24]. This would lead to liquid seepage be-
tween the insulating layers and internal corrosion [24,35] and the resulting decrease in
impedance might skew the interpretation of the electrochemical data.

Figure 2. Schematic overview and cross-section of the electrochemical cell setup employed in the
experiments with PEDOT:PSS electrodes in the current work.
Polymers 2024, 16, 1706 6 of 17

Under the above conditions, each of the 3 PEDOT:PSS/Graphene samples was tested
at least in 3 points in different areas of the surface. The samples regarded as “fresh” were
tested within a short timeframe (<1 week) after their deposition. The influence of two types
of aging on the electrochemical performance of the “fresh” samples was tested: (1) “aged
in air”, which were stored under exposure to air (50 ± 5% RH and 25 ± 1 °C) for 6 months,
prior testing; and (2) “aged in artificial sweat”, which were tested under exposure to a
simulated sweat liquid-phase environment, for up to 9 days.

3. Results
3.1. Electrochemical Stability
Cyclic voltammetry was employed to investigate charge transfer at the elec-
trode/electrolyte interface. In the potential range from −0.3 to 0.6 V, characteristic reaction
peaks were not registered either in the anodic or in the cathodic branch of the CV depend-
ences (Figure 3a). Therefore, CV profiles suggest predominantly capacitive charging of
the layer, and no faradaic reactions are observed [2]. The capacitive current linearly in-
creases with the rise in potential scan rate from 0.01 to 1.0 V (Figure 3b). The slope of the
linear fit to the dependence of the capacitive current density jc (mA m−2) on the scan rate ν
(V s−1) represents the voltametric capacitance (CCV) of the PEDOT:PSS/Graphene layer in
artificial sweat [2,36]. The determined CCV value stands at 111.6 mF m−2.

(a) (b)
Figure 3. Scan rate influence on polarization response of PEDOT:PSS/Graphene layer in artificial
sweat. (a) CV dependences at scan rate from 0.01 to 1.0 V cm−2; (b) dependence of the capacitive
voltametric current on scan rate.

Assessment of the charging ability during continuous cycling (500 cycles) was con-
ducted by calculating the voltametric charge, QCV, from the hysteresis area of the CV
curves (Figure S2) using the equation:
E
1 2
2v E1
QCV = j( E)dE (1)

where ν is the scan rate (0.1 V s−1), E1 and E2 are the potential window, and j is the current
density at each potential. Figure 4 present the calculated values, according to Equation (1)
and the capacity retention calculated as ratio of hysteresis area of a cycle n to that of the
first cycle. The average voltametric charge values decrease from 12.4 to 10.0 µC cm−2, in-
dicating a capacity retention after prolonged cycling of approximately 90.5 ± 5.1%. The
degradation of the layer is most pronounced in the initial 10 cycles (approximately 3.5%).
Polymers 2024, 16, 1706 7 of 17

Figure 4. Voltammetry charge and capacitance retention dependencies upon 500 CV cycles. Meas-
urements were performed between −0.3 and 0.7 V vs. Ag/AgCl at a scan rate of 0.1 V s−1.

The impact of electrochemical cycling on the impedance of the layer was investigated
through EIS tests. Figure 5 depicts Nyquist and Bode plots of tests before and after 500
scans in artificial sweat in a 1 V potential window (from −0.3 to 0.7 V). In both tests, the
Nyquist plot (Figure 5a) takes the form of an incomplete capacitive semicircle. After elec-
trochemical cycling, the dependencies at high frequencies overlap, with an observed in-
crease in the slope of the arc at low frequencies, reflecting an increase in the layer’s capac-
itance. The Bode plots presented in Figure 5b also show a slight increase in capacitive
behavior after cycling at frequencies below 10 Hz, demonstrated by the higher phase shift.
However, the observed differences can be considered negligible.

(a) (b)
Figure 5. EIS results for the PEDOT:PSS/Graphene layer before and after 500 cycles at 100 mV/s. (a)
Nyquist plot (inset: the equivalent circuit used to simulate the Nyquist plots); (b) Bode plot of mag-
nitude and phase of the impedance.

Low and frequency-independent impedance for PEDOT:PSS/Graphene layers is reg-


istered in the high-frequency range, where the PEDOT:PSS/Graphene layer exhibits a
purely resistive character (the phase shift does not exceed −10°) and remains unchanged in
magnitude after 500 cycles in artificial sweat. The cut-off frequency (fcut-off) for the transition
from resistive to capacitive behavior slightly decreases from 2.53 to 1.63 kHz after cycling.
These values are close to 1 kHz, which is a representative frequency in biosignals [2]. At low
and medium frequencies, the impedance increases linearly, indicating the dominance of ca-
Polymers 2024, 16, 1706 8 of 17

pacitive behavior in the layers. In this range, the magnitude dependence after electrochem-
ical aging lies just below that of a freshly deposited layer and demonstrates a slight decrease
in impedance (from 7.3 MΩ to 5.7 MΩ at 10−2 Hz) and an increase in the capacitive behavior
of the layer (additional phase shift of approximately −7° at 10−2 Hz). The layer exhibits stable
capacitive behavior in the frequency range of 0.1–300 Hz, as the phase angle remains more
negative than −75°. The reduction in phase shift at low frequencies to −61° for the fresh layer
is associated with parallel surface reactions contributing to charge transfer [37]. Presumably,
during electrochemical cycling, some functional groups in PEDOT:PSS exposed to the elec-
trolyte are deactivated, reducing the charge exchange, i.e., the resistive component of im-
pedance. This aligns with the CV results discussed earlier.
Various equivalent circuits have been proposed for the analysis of PEDOT:PSS-based
layers, ranging from the simplest serial RC circuit [21,38], Randles circuit, Rs(Cdl(RctW))
[2,7,39], to significantly more complex ones involving multiple RC blocks and Warburg im-
pedance (W) [2,8]. The choice of an equivalent circuit is significantly influenced by the sub-
strates used and additives to the conducting polymer. We experimented with different
equivalent circuits to analyze our results for PEDOT:PSS/Graphene on a PET insulating sub-
strate. The best fit was achieved with the equivalent circuit denoted by the description code
[Rs([Rct(RiQi)]Cdl)], represented as an inset in Figure 5a. In this circuit, Rs, Rct, and Ri represent
the solution resistance, charge transfer resistance of the external solid/liquid interface, and
the internal resistance of the composite layer, respectively. The Cdl present the contribution
of the double layer capacitance to impedance. The (RiQi) block connected in series with Rct
describes various transfer mechanisms within the interior of the PEDOT:PSS/Graphene
composite, such as ionic flow in the PSS ion channels and electrical conductivity in gra-
phene and PEDOT. Using a constant-phase element (Qi) reflects the surface inhomogene-
ity and porosity of the composite layer, resulting in a phase shift that does not reach the
typical −90° for pure capacitance. The discrepancies between experimental behavior and
combinations of purely electrical elements require the introduction of non-electrical ele-
ments such as Q into the equivalent electrical circuit to account for film properties, charge
carrier mobility, conduction path, and driving forces [40]. Using a constant-phase element
is especially necessary in cases of conducting polymers characterized by heterogeneity
and disorder (at least at the molecular level). In the literature, this equivalent electrical
circuit is used to describe the electrochemical behavior of organic layers with bipolar mol-
ecules that form a multilayered assembly, blocking the transfer of electron charge [41].
The fits achieved with the proposed equivalent electrical circuit are represented as
black lines on the Nyquist and Bode plots (Figure 5), and the values of individual elements
are given in Table 1. The χ2 values of both samples were obtained in the range between
10−4 and 10−3, indicating a sufficiently good fit. The most significant changes after 500 cy-
cles are observed in the values of Rct and Cdl, which increase by approximately two times,
from 100.6 to 207.7 Ω and from 257.6 to 450.6 nF, respectively. The capacitance could in-
crease either due to an increase in the active electrochemical surface of the electrode or the
volumetric capacitance of the PEDOT:PSS layer. Both pathways are associated with
changes in surface structure during electrochemical cycling [42].

Table 1. EIS data for PEDOT:PSS/Graphene of fresh layers and after aging through electrochemical
cycling in artificial sweat and exposure to air.

Sample fcut-off, kHz Rs, Ω Rct, Ω Ri, MΩ Qi, µS sn n Cdl, nF χ2


fresh 2.53 ± 0.50 97.1 ± 3.2 100.6 ± 21.4 30.6 ± 5.2 1.16 ± 0.12 0.87 ± 0.03 257.6 ± 42.5 0.0005 ± 0.0002
after 500 cycles 1.63 ± 0,11 100.1 ± 4.3 207.7 ± 45.9 36.7 ± 11.6 1.34 ± 0.18 0.83 ± 0.01 450.6 ± 60.1 0.0014 ± 0.0006
aged in air 1.94 ± 0.59 120.4 ± 9.1 115.1 ± 33.7 40.6 ± 7.4 1.19 ± 0.22 0.86 ± 0.01 269.3 ± 34.3 0.0002 ± 0.0001

The changes in the surface morphology of the conductive polymer layer were ob-
served through scanning electron microscopy. SEM images in the back-scattered electron
(BSE) mode, taken before and after 500 cycles in artificial sweat are presented in Figure 6.
Polymers 2024, 16, 1706 9 of 17

The BSE mode was chosen because the images are more contrasted, especially after aging
of the layer (Figure S3). The image of the untreated surface shows relatively evenly dis-
tributed bright globular aggregates of PEDOT molecules linked with PSS. Figure 6b de-
picts a more heterogeneous area of the PEDOT:PSS/Graphene layer’s surface after electro-
chemical aging. In the BSE mode, the surface appears more heterogeneous with bright
islands. The observed changes are likely a result of swelling and partial dissolution of the
PSS matrix, while the distribution and size of PEDOT aggregates appear unchanged.

Figure 6. SEM imagines in the BSE mode of PEDOT:PSS/Graphene layer. (a) Before electrochemical
testing; (b) after 500 scans in artificial sweat.

3.2. Air Aging


An essential characteristic of materials used in sensor devices is their stability during
storage. The physicochemical interaction of organic layers with the surrounding environ-
ment is inevitable. It manifests in aging processes within polymeric macromolecules, the
adsorption of oxygen, moisture, and contaminants, resulting in changes in the volume and
surface properties of composite layers. The occurring changes inevitably impact the imped-
ance of the layers and can, therefore, be adequately traced through electrochemical imped-
ance spectroscopy. Figure 7a illustrates the impedance characteristics obtained from fresh
PEDOT:PSS/Graphene layers and after exposure to air (RH approximately 50% and room
temperature) for 6 months. No significant alteration in the Nyquist and Bode plots’ shape is
observed, with partial overlap in the dependencies before and after air aging. The equivalent
electrical circuit [Rs([Rct(RiQi)]Cdl)] used for fresh layers is also applicable to air-aged layers.
The values of the elements from this circuit are presented in Table 1.

(a) (b)
Figure 7. EIS dependencies in artificial sweat for fresh deposited (green triangles) composite PE-
DOT:PSS/Graphene layers and after 6 months (red circles) of exposure to the air. (a) Nyquist plots;
(b) Bode plots.
Polymers 2024, 16, 1706 10 of 17

The primary trends during air aging include a slight increase in the capacitive semi-
circle radius in the Nyquist plots, reflecting a mild elevation in the values of Rct, Ri, and
Cdl. These results indicate that PEDOT:PSS/Graphene layers exhibit a weak tendency to
age and increase their impedance when exposed to ambient air. The aging processes may
be associated with drying and cracking of the layer over extended exposure periods [43].

3.3. Aging in Artificial Sweat Simultaneously with Electrochemical Loading


In addition to stability during sensor storage, it is crucial to determine the stability of
the conducting material in operational conditions. In this context, the impedance of a PE-
DOT:PSS/Graphene layer was periodically measured during prolonged exposure to arti-
ficial sweat. Electrochemical impedance spectroscopy (EIS) tests were alternated with
electrochemical loading of the layer through linear polarization within a wide potential
range from −0.2 to 1.0 V. Anodic polarization predominates since oxidative processes are
expected to exert a more pronounced negative influence on polymer aging processes. The
results of impedance tests and polarization dependencies are presented in Figures 8 and
9, respectively. Unlike air exposure, aging under operational conditions significantly al-
ters the course of impedance characteristics.

(a) (b) (c)


Figure 8. Aging of PEDOT:PSS/Graphene layer in artificial sweat. (a) Nyquist plot; (b) frequency
dependence of impedance magnitude; (c) phase shift from frequency.

Notably, Nyquist and Bode plots for the layer after 24 h of exposure (red inverted
triangles) are positioned close to those of the initial scan, maintaining stable capacitive
behavior at frequencies below 1 kHz. After 48 h, substantial changes in the frequency de-
pendencies of impedance magnitude and phase shift become evident. A significant in-
crease in impedance magnitude is observed across the entire frequency range during ex-
posure exceeding 24 h (Figure 8b). Higher impedance values indicate slower ion exchange
and a reduced layer capacitance compared to the fresh state [44]. This is further empha-
sized in Figure 8c, where a decrease in phase shift (below −70°) is registered in the low-
and mid-frequency range, accompanied by the formation of two peaks (at frequencies of
approximately 1 kHz and 0.1 Hz). These changes in Bode plots demonstrate an increase
in the resistive component’s contribution to the overall system impedance, possibly due
to thinning of the conducting layer [2] and the emergence of pores or new phase bounda-
ries with the electrolyte [41]. Furthermore, by the third day, the cut-off frequency (fcut-off)
shifts to higher frequencies, indicating the deterioration in the layer’s quality. Considering
that higher fcut-off values are characteristic of thinner layers with the same material, it can
be assumed that the occurring deteriorations during exposure to artificial sweat lead to
layer thinning.
The prolonged contact with the electrolyte (beyond 2 days) leads to a decrease in the
impedance magnitude in the mid- and low-frequency ranges (Figure 8b). This indicates
that changes in the composite are likely associated with an increase in ion content within
Polymers 2024, 16, 1706 11 of 17

the interior of the polymer layer. Simultaneously a reduction in the capacitive charging of
the layer is observed.
The deviation of the phase shift from −90° necessitates a change in the equivalent
circuit. In the proposed circuit for the fresh layer, Сdl is replaced with a constant-phase
element Qdl. The resultant fits from the equivalent circuit [Rs([Rct(RiQi)]Qdl)] are repre-
sented as black lines on the plots in Figure 8, and the values of the elements are presented
in Table 2.

Table 2. Data from electrochemical impedance spectra of PEDOT:PSS/Graphene layers after aging
in artificial sweat.
Soaking, fcut-off Rs Rct Ri Qi , Qdl
ni ndl χ2
Days kHz Ω kΩ MΩ µS sn nS sn
0 2.34 97.10 0.101 30.60 1.160 0.870 257.6 1.00 0.0005
1 2.57 231.5 1.522 65.03 0.448 0.875 236.0 0.96 0.0023
2 5.89 1885 967.0 329.5 0.053 0.646 38.03 0.87 0.0124
3 4.17 3165 648.7 791.9 0.082 0.649 0.049 0.81 0.0045
4 2.19 5604 922.5 1271 0.101 0.671 0.056 0.77 0.0051
7 2.40 5222 1077 4.01 × 012 0.052 0.843 0.153 0.64 0.0026
8 1.35 5160 547.0 5.469 × 1015 0.093 0.894 0.188 0.64 0.0037
9 0.76 4668 220.0 1.89 × 1015 0.121 0.801 0.209 0.67 0.0029

Contact with artificial sweat significantly elevates the values of all resistances in the
equivalent circuit, with the internal resistance Ri being the most affected. Structural and
chemical changes in the layer also lead to a decrease in capacitances, with Qi values de-
creasing by approximately 1 order of magnitude, and those of Qdl by more than 3 orders
of magnitude. The values of the frequency dispersion coefficient ndl also decrease, drop-
ping below 0.7 after 1 week.
Fitting some results with the proposed equivalent circuit is not sufficiently good, and
the χ2 values are not representative enough. Achieving a complete coincidence between
experimental results and fit of the equivalent circuit is a serious challenge for PE-
DOT:PSS/Graphene since this material encompasses phases with different conductivity
mechanisms, further complicating modeling due to different capacitive effects manifest-
ing in various frequency ranges [40]. Indeed, better fitting of experimental results of tests
over two days in artificial sweat is achieved using more complex equivalent electrical cir-
cuits such as hybrid series/parallel models, involving more elements corresponding to
distributed electronic phases in an ion-conductive environment [45]. Nevertheless, we
judged that using the same circuit for all electrolyte aging results has more advantages in
tracking degradation processes than reducing χ2 values at the expense of employing dif-
ferent and more complex equivalent electrical circuits in the modeling.
Between each EIS test, the sample is polarized in the range of −0.2 V to 1.0 V vs.
Ag/AgCl at a rate of 1 mV s−1 for the purpose of electrochemical loading of the polymer
layer. The obtained polarization dependencies are presented in Figure 9.

Figure 9. Polarization dependencies of PEDOT:PSS/Graphene at different soaking durations in arti-


ficial sweat.
Polymers 2024, 16, 1706 12 of 17

EIS spectra recorded before and after polarization of the layer from −0.2 to 1 V overlap
completely, indicating that polarization, even at these slow potential scanning rates over
a relatively wide potential range, does not significantly influence the properties of the
composite, unlike the duration of its exposure to artificial sweat.
From the polarization dependencies, the exchange current density (j0) and polariza-
tion resistance (Rp) were determined and are presented in Table 3. The results show a very
low exchange current (around and below nanoamperes) and a very high polarization re-
sistance (around and above 100 MΩ), which is expected for a layer that does not sponta-
neously interact with the surrounding electrolyte. The exchange current density (j0) of the
electrochemical system depends on various factors of the material and the environment
such as ion concentration, temperature, catalytic properties, and the specific surface area
of the electrode [46].
After one day of exposure, the exchange current decreases to approximately 3 nA
cm−2 and remains around this value until the end of the experiment. The lower exchange
current indicates an increase in the resistance of the conducting layer, most likely due to
the destruction and thinning (reduction in effective conductive cross-section) of the layer.
After 4 days of exposure to artificial sweat, the potential shifts in the negative direction.
This could be interpreted as an indication that the distance between the electrolyte and
the aluminum current collector has decreased to the point where the ion flow reaches the
aluminum surface, and the aluminum/conductive polymer interface begins to influence
the recorded polarization current signal.

Table 3. Electrochemical parameters extracted from the polarization dependencies of the PE-
DOT:PSS/Graphene layer aging in artificial sweat.

Soaking, days E, V j0, nA cm−2 Rp, MΩ Cathodic Tafel, V dec−1 Anodic Tafel, V dec−1
0 0.055 21.11 146.2 0.154 0.132
1 0.071 3.38 106.6 0.179 0.155
2 0.082 2.21 189.2 0.252 0.156
3 0.081 3.42 149.6 0.290 0.198
4 0.078 3.66 121.5 0.308 0.153
7 0.041 3.38 114.5 0.212 0.154
8 0.024 3.65 88.04 0.171 0.131
9 −0.001 3.85 75.33 0.162 0.113

SEM observations of the sample after prolonged exposure to artificial sweat show the
loss of the PSS phase, exposure of PEDOT:PSS clusters (Figure 10a, bottom left corner,
Figure 10b, Figure S4), with extensive areas in the middle of the test field completely de-
stroyed (Figure 10a, bottom right corner). In certain areas, the destruction of the composite
extends beyond the seal line (Figure 10a, yellow dashed line), primarily through cracks in
the layer.

Figure 10. SEM imagines in the BSE mode of the PEDOT:PSS/Graphene layer after 9 days in artificial
sweat and polarization tests at low magnification (a) and at high magnification (b).
Polymers 2024, 16, 1706 13 of 17

Unfortunately, although clearly visible by SEM layer destruction, our attempts to de-
termine with Fourier-transform infrared spectroscopy (FTIR) the chemical changes in the
polymer layer were unsuccessful. The obtained FTIR spectra are dominated by the peaks
of the PET substrate (Figure S5) and those characteristic of S–phenyl (at 1157, 1121 and
1012 cm−1), C=C bonds of the thiophene ring (at 1532 and 1521 cm−1), C–C bonds of the
thiophene ring (at 1356 and 1312 cm−1) and for the sulfonic acid group of the PSS (at 1196
cm−1) are not clearly registered [23].

4. Discussion
A specific feature of organic layers is their loose structure and the possibility of pen-
etration by various components from the surrounding environment. In this regard, PE-
DOT is considered a dense material, taking several days to wet with electrolyte [45]. How-
ever, its composites exhibit different behavior. Although PEDOT has high ion resistance,
the macroporous region around carbon particles, when included, allows ions to travel
long distances with relatively high ion mobility. The distribution of electronic phases fur-
ther reduces effective ion resistance [45]. Regarding PSS, acting as a connecting substance
between the other two components, it swells in a water environment due to water and ion
penetration, such as Na+, Cl−, and K+, reaching the internal structure of the hydrophilic
polymer. Therefore, the nanocomposite PEDOT:PSS/Graphene material is a complex sys-
tem where various physical processes contribute to charge transport. It can be assumed
that within the interior of the PEDOT-based system, there are two main types of phase
boundaries: PEDOT/PSS and PSS/Graphene. Studies of PEDOT:PSS and PEDOT:PSS/Gra-
phene systems have shown that graphene and PEDOT + PSS agglomerates are located in
the PSS matrix [32,47–49]. This suggests that the composite layer/electrolyte boundary can
be simplified to PSS/electrolyte, as the other phases mainly contact the electrolyte through
the PSS phase. Our EIS results show that the electrode material capacitive binds to the
surrounding electrolyte through the formation of a double electric layer, in line with other
studies [2,15,17,20]. This suggests hindered charge exchange at this boundary and initial
chemical stability. From this perspective, PEDOT:PSS/Graphene seems suitable for man-
ufacturing printed electrodes for capacitive sensors.
The process of wetting and destruction of the PSS phase is accelerated by electro-
chemical cycling as a result of potential-stimulated continuous absorption and release of
electrolyte ions and associated volume changes [15,17]. Ion transfer is associated with ex-
pansion in receiving zones and contraction on the donating side [45]. Thus, structural
changes occur in the layer during electrochemical loading, expressed in the penetration of
ions from the electrolyte into the depth of the layer [2]. The presented CV dependencies
in Figure S2 show that the reduction in the hysteresis area (voltametric charge) is primar-
ily the result of a decrease in the anodic slope at potentials above 0.3 V. Therefore, positive
polarization, which allows for oxidative processes, can be considered as more destruc-
tively acting electrochemical loading. It is known that in an oxidative environment (such
as H2O2), the thiophene sulfur atoms in PEDOT are over-oxidized and/or the double bonds
in the conjugated structure are broken, leading to a reduced conjugation length and, there-
fore, reduced conductivity. As a result, during electrochemical aging or contact with an
oxidative environment, the electrode becomes more insulating due to an increase in its
impedance [23,49].
The significant change in the impedance characteristics during prolonged contact of
the layer with the artificial sweat is probably caused by structural and chemical changes
in the PEDOT:PSS/Graphene composite. As mentioned above, the changes occurring
within three days result in a decrease in the capacitance of the double electric layer and
an increase in the layer resistance by over 3 orders of magnitude. This indicates that ho-
mogeneity within the film increases during exposure to the electrolyte, associated with
changes in thickness, ion adsorption on the film, and interior pores filled with electrolyte
[40]. Over time, the distribution of the phase with the highest ion conductivity becomes
uneven, decreasing and exposing PEDOT and graphene, while resistance values reach
Polymers 2024, 16, 1706 14 of 17

levels equivalent to layer breakage. Therefore, the results of aging in the electrolyte clearly
show that PSS should be treated as water-soluble under operational conditions for wear-
able biosensors, i.e., in contact with sweat and periodic electrochemical loading. During
this aging process, surface PSS molecules gradually detach from the composite, and after
a day or two, exposure of the other two phases begins. Electrolyte access to the internal
interphase boundaries of graphene-PSS leads to the breakdown of interphase bonds and
delamination [20]. With time, the electrolyte gains access to deeper structures, reaching
the boundary with the insulating substrate, resulting in the breakdown of the electrode.
The presented EIS results show that the duration of use of PEDOT:PSS/Graphene layers
as a conductive pattern material for monitoring electrochemical parameters in an electro-
lyte environment must be carefully determined. For longer tests, it is necessary to isolate
the layers both from the atmosphere and from the electrolyte.
The perspectives for development and further investigations following these findings
are multifaceted. Firstly, the development of additional protective measures or encapsu-
lation techniques to improve the stability of PEDOT:PSS/Graphene layers in saline-acidic
electrolytes could be explored. This would involve studying advanced coating materials
or barrier technologies to mitigate the observed degradation during prolonged exposure
to artificial sweat. Furthermore, it would be valuable to investigate a variety of additional
dopants to the existing PEDOT:PSS/Graphene composition that could enhance long-term
stability and performance. Examples of advanced coating materials could involve resilient
polymer blends, engineered nanocomposites, or functionalized nanostructures with spe-
cific protective properties. As for barrier technologies, these could encompass specialized
encapsulation methods, such as conformal coatings, or nanostructured encapsulants en-
gineered to shield the conductive polymer layers from the corrosive effects of the electro-
lyte. These barrier technologies may be designed to prevent moisture penetration, block
ion diffusion, and minimize chemical interactions that contribute to the observed degra-
dation of the PEDOT:PSS/Graphene layers in artificial sweat.

5. Conclusions
In this study, we investigated the electrochemically induced changes in PE-
DOT:PSS/Graphene ink layer on a flexible insulating PET substrate to demonstrate the
potential for use of this material as printed conductive patterns and electrodes. The pre-
sented results show that the layer exhibits predominantly capacitive behavior in the range
of −0.2 to 0.6 V vs. Ag/AgCl, with a cut-off frequency (fcut-off) of approximately 1 kHz, mak-
ing it suitable for high-frequency impedance measurements (as Electrical Impedance Ple-
thysmography and Impedance Cardiography) and as capacitive sensors in the low-fre-
quency range. During electrochemical cycling at 1 V potential window over the first 10
cycles, the capacitance decreases by approximately 10%, stabilizing at approximately 90%
of the initial value by the end of the 500-cycle test. Aging for 6 months under ambient
conditions leads to a slight increase in impedance, demonstrating the material’s potential
for storage under undemanding conditions.
The layer shows the most pronounced degradation during prolonged exposure to
artificial sweat. The impedance of the layer increases due to an increase in charge transfer
resistance (from 101 Ω to 1.1 MΩ) and internal resistance (from 30.6 MΩ to 5.5 × 1015 MΩ),
indicating severe damage and destruction of the polymer layer, as confirmed by SEM ob-
servations. The presented results for aging in artificial sweat question the possibility of
using spray-coated PEDOT:PSS/Graphene layers as the conductive pattern material for
long-term monitoring of electrochemical parameters in a weakly acidic electrolytic envi-
ronment. For longer tests, it is advisable to use additional protection to enhance its stabil-
ity in saline-acidic electrolytes.
Understanding the stability and performance of PEDOT:PSS/Graphene in various
conditions, such as exposure to artificial sweat and air, is crucial for determining its suit-
ability for biosensor applications. Additionally, the examination of morphological
Polymers 2024, 16, 1706 15 of 17

changes and the ability to mitigate potential issues related to electrode–substrate interac-
tions can contribute to the development of more effective and reliable biosensor technol-
ogies. Therefore, the findings of this study could have significant implications for the de-
sign and development of biosensors utilizing conductive polymer patterns.
Future research should focus on enhancing the stability of PEDOT:PSS/Graphene
layers in saline electrolytes by exploring advanced protective coatings and barrier tech-
nologies. Investigating new dopants to improve long-term stability is also important. Bar-
rier technologies like specialized encapsulation methods can prevent degradation in arti-
ficial sweat by shielding the layers from electrolyte corrosion and moisture penetration.

Supplementary Materials: The following supporting information can be downloaded at:


https://www.mdpi.com/article/10.3390/polym16121706/s1, Figure S1: Raman spectra of PE-
DOT:PSS/Graphene ink. (a) PEDOT:PSS/Graphene dropped on PET; (b) PEDOT:PSS/Graphene
sprayed on PET. Figure S2: Progressive shape change in CV plot of the PEDOT:PSS/Graphene layer
in artificial sweat. Figure S3. SEM raw imagines of PEDOT:PSS/Graphene layer. (a) before electro-
chemical testing in the SE and BES modes; (b) after 500 scans in artificial sweat in the SE and BES
modes. Figure S4. SEM imagines in the BSE mode of the PEDOT:PSS/Graphene layer after 9 days in
artificial sweat and polarization tests at low magnification (a) and at high magnification (b). Figure
S5. FTIR spectra of PET substrate, fresh and aged in artificial sweat PEDOT:PSS/Graphene sprayed
on PET.
Author Contributions: conceptualization, B.T.; methodology, B.T. and I.I.; investigation, B.T., V.M.
and M.A.; data curation, B.T., V.M., B.S. and I.I.; writing—original draft preparation, B.T. and M.A.;
writing—review and editing, I.I. and B.S.; supervision, I.I. All authors have read and agreed to the
published version of the manuscript.
Funding: This research was funded by the European Union-NextGenerationEU through the Na-
tional Recovery and Resilience Plan of the Republic of Bulgaria, grant number BG-RRP-2.004-0005.
Institutional Review Board Statement: Not applicable.
Data Availability Statement: The original contributions presented in the study are included in the
article/supplementary material, further inquiries can be directed to the corresponding author.
Conflicts of Interest: The authors declare no conflict of interest. The funders had no role in the de-
sign of the study; in the collection, analyses, or interpretation of data; in the writing of the manu-
script; or in the decision to publish the results.

References
1. Vacca, A.; Mascia, M.; Rizzardini, S.; Corgiolu, S.; Palmas, S.; Demelas, M.; Bonfiglio, A.; Ricci, P.C. Preparation and characteri-
sation of transparent and flexible PEDOT:PSS/PANI electrodes by ink-jet printing and electropolymerisation. RSC Adv. 2015, 5,
79600–79606. https://doi.org/10.1039/C5RA15295J.
2. Mousavi, H.; Ferrari, L.M.; Whiteley, A.; Ismailova, E. Kinetics and Physicochemical Characteristics of Electrodeposited PE-
DOT:PSS Thin Film Growth. Adv. Electron. Mater. 2023, 9, 2201282. https://doi.org/10.1002/aelm.202201282.
3. Donahue, M.J.; Sanchez-Sanchez, A.; Inal, S.; Qu, J.; Owens, R.M.; Mecerreyes, D.; Malliaras, G.G.; Martin, D.C. Tailoring PEDOT
properties for applications in bioelectronics. Mater. Sci. Eng. R Rep. 2020, 140, 100546. https://doi.org/10.1016/j.mser.2020.100546.
4. Sanchez-Sanchez, A.; del Agua, I.; Malliaras, G.G.; Mecerreyes, D. Conductive Poly(3,4-Ethylenedioxythiophene) (PEDOT)-
Based Polymers and Their Applications in Bioelectronics. In Smart Polymers and Their Applications, 2nd ed.; Aguilar, M.R.,
Román, J.S., Eds.; Woodhead: Sawston, UK, 2019; pp. 191–218. https://doi.org/10.1016/B978-0-08-102416-4.00006-5.
5. Kaur, G.; Kaur, A.; Kaur, H. Review on nanomaterials/conducting polymer based nanocomposites for the development of bio-
sensors and electrochemical sensors. Polym. Plast. Technol. Mater. 2021, 60, 504–521.
https://doi.org/10.1080/25740881.2020.1844233.
6. Lunghi, A.; Mariano, A.; Bianchi, M.; Dinger, N.B.; Murgia, M.; Rondanina, E.; Toma, A.; Greco, P.; Di Lauro, M.; Santoro, F.; et
al. Flexible neural interfaces based on 3D PEDOT:PSS micropillar arrays. Adv. Mater. Interfaces 2022, 9, 2200709.
https://doi.org/10.1002/admi.202200709.
7. Sakthinathan, I.; Yamasaki, N.; Barreca, D.; Maccato, C.; Ueda, T.; McCormac, T. Wells-Dawson type polyoxometalate,
[S2W18O62]4−-doped poly(3,4-ethylenedioxythiophene) films: Voltammetric behaviour and applications to selective bromate
detection. Electrochim. Acta 2023, 462, 142689. https://doi.org/10.1016/j.electacta.2023.142689.
8. Tian, F.; Yu, J.; Wang, W.; Zhao, D.; Cao, J.; Zhao, Q.; Wang, F.; Yang, H.; Wu, Z.; Xu, J.; et al. Design of adhesive conducting
PEDOT-MeOH:PSS/PDA neural interface via electropolymerization for ultrasmall implantable neural microelectrodes. J. Colloid
Interface Sci. 2023, 638, 339–348. https://doi.org/10.1016/j.jcis.2023.01.146.
Polymers 2024, 16, 1706 16 of 17

9. Hik, F.; Taatizadeh, E.; Takalloo, S.E.; Madden, J.D. Fast electrochemical response of PEDOT:PSS electrodes through large
combined increases to ionic and electronic conductivities. Electrochim. Acta 2023, 468, 143136.
https://doi.org/10.1016/j.electacta.2023.143136.
10. Seiti, M.; Giuri, A.; Corcione, C.E.; Ferraris, E. Advancements in tailoring PEDOT:PSS properties for bioelectronic applications:
A comprehensive review. Biomater. Adv. 2023, 154, 213655. https://doi.org/10.1016/j.bioadv.2023.213655.
11. Moniz, M.P.; Rafique, A.; Carmo, J.; Oliveira, J.P.; Marques, A.; Ferreira, I.M.M.; Baptista, A.C. Electrospray Deposition of
PEDOT:PSS on Carbon Yarn Electrodes for Solid-State Flexible Supercapacitors. ACS Appl. Mater. Interfaces 2023, 15, 30727–
30741. https://doi.org/10.1021/acsami.3c03903.
12. Castagnola, V.; Bayon, C.; Descamps, E.; Bergaud, C. Morphology and conductivity of PEDOT layers produced by different
electrochemical routes. Synth. Met. 2014, 189, 7–16. https://doi.org/10.1016/j.synthmet.2013.12.013.
13. Saha, A.; Ohori, D.; Sasaki, T.; Itoh, K.; Oshima, R.; Samukawa, S. Effect of Film Morphology on Electrical Conductivity of
PEDOT:PSS. Nanomaterials 2024, 14, 95. https://doi.org/10.3390/nano14010095.
14. Ge, Y.; Jalili, R.; Wang, C.; Zheng, T.; Chao, Y.; Wallace, G.G. A robust free-standing MoS2/poly(3,4-
ethylenedioxythiophene):poly(styrenesulfonate) film for supercapacitor applications. Electrochim. Acta 2017, 235, 348–355.
https://doi.org/10.1016/j.electacta.2017.03.069.
15. Wustoni, S.; Saleh, A.; El-Demellawi, J.K.; Koklu, A.; Hama, A.; Druet, V.; Wehbe, N.; Zhang, Y.; Inal, S. MXene improves the
stability and electrochemical performance of electropolymerized PEDOT films. APL Mater. 2020, 8, 121105.
https://doi.org/10.1063/5.0023187.
16. Schander, A.; Teßmann, T.; Strokov, S.; Stemmann, H.; Kreiter, A.K.; Lang, W. In-vitro evaluation of the long-term stability of
PEDOT:PSS coated microelectrodes for chronic recording and electrical stimulation of neurons. In Proceedings of the 38th
Annual International Conference of the IEEE Engineering in Medicine and Biology Society (EMBC), Orlando, FL, USA, 16–20
August 2016; pp. 6174–6177. https://doi.org/10.1109/EMBC.2016.7592138.
17. Boehler, C.; Oberueber, F.; Schlabach, S.; Stieglitz, T.; Asplund, M. Long-Term Stable Adhesion for Conducting Polymers in
Biomedical Applications: IrOx and Nanostructured Platinum Solve the Chronic Challenge. ACS Appl. Mater. Interfaces 2017, 9,
189–197. https://doi.org/10.1021/acsami.6b13468.
18. Ouyang, L.; Wei, B.; Kuo, C.-c.; Pathak, S.; Farrell, B.; Martin, D.C. Enhanced PEDOT adhesion on solid substrates with
electrografted P(EDOT-NH2). Sci. Adv. 2017, 3, e1600448. https://doi.org/10.1126/sciadv.160044.
19. Qu, J.; Garabedian, N.; Burris, D.L.; Martin, D.C. Durability of Poly(3,4-ethylenedioxythiophene) (PEDOT) films on metallic
substrates for bioelectronics and the dominant role of relative shear strength. J. Mech. Behav. Biomed. Mater. 2019, 100, 103376.
https://doi.org/10.1016/j.jmbbm.2019.103376.
20. Inoue, A.; Yuk, H.; Lu, B.; Zhao, X. Strong adhesion of wet conducting polymers on diverse substrates. Sci. Adv. 2020, 6,
eaay5394. https://doi.org/10.1126/sciadv.aay5394.
21. Koutsouras, D.A.; Gkoupidenis, P.; Stolz, C.; Subramanian, V.; Malliaras, G.G.; Martin, D.C. Impedance Spectroscopy of Spin-
Cast and Electrochemically Deposited PEDOT:PSS Films on Microfabricated Electrodes with Various Areas. ChemElectroChem
2017, 4, 2321–2327. https://doi.org/10.1002/celc.201700297.
22. Abdullayeva, N.; Sankir, M. Influence of Electrical and Ionic Conductivities of Organic Electronic Ion Pump on Acetylcholine
Exchange Performance. Materials 2017, 10, 586. https://doi.org/10.3390/ma10060586.
23. Sriprachuabwong, C.; Karuwan, C.; Wisitsorrat, A.; Phokharatkul, D.; Lomas, T.; Sritongkhamb, P.; Tuantranont, A. Inkjet-
printed graphene-PEDOT:PSS modified screen printed carbon electrode for biochemical sensing. J. Mater. Chem. 2012, 22, 5478–
5485. https://doi.org/10.1039/C2JM14005E.
24. Boehler, C.; Carli, S.; Fadiga, L.; Stieglitz, T.; Asplund, M. Tutorial: Guidelines for standardized performance tests for electrodes
intended for neural interfaces and bioelectronics. Nat. Protoc. 2020, 15, 3557–3578. https://doi.org/10.1038/s41596-020-0389-2.
25. Wahyuni, W.T.; Putra, B.R.; Rahman, H.A.; Anindya, W.; Hardi, J.; Rustami, E.; Ahmad, S.N. Electrochemical Sensors based on
Gold–Silver Core–Shell Nanoparticles Combined with a Graphene/PEDOT:PSS Composite Modified Glassy Carbon Electrode
for Paraoxon-ethyl Detection. ACS Omega 2024, 9, 2896–2910. https://doi.org/10.1021/acsomega.3c08349.
26. Liu, Z.; Parvez, K.; Li, R.; Dong, R.; Feng, X.; Müllen, K. Transparent conductive electrodes from graphene/PEDOT:PSS hybrid
inks for ultrathin organic photodetectors. Adv. Mater. 2014, 27, 669–675. https://doi.org/10.1002/adma.201403826.
27. Eawwiboonthanakit, N.; Jaafar, M.; Ahmad, Z.; Ohtake, N.; Lila, B. Fabrication of PEDOT: PSS/graphene conductive ink printed
on flexible substrate. Solid State Phenomen. 2017, 264, 70–73. https://doi.org/10.4028/www.scientific.net/SSP.264.70.
28. Popov, V.I.; Kotin, I.A.; Nebogatikova, N.A.; Smagulova, S.A.; Antonova, I.V. Graphene-PEDOT:PSS Humidity Sensors for High
Sensitive, Low-Cost, Highly-Reliable, Flexible, and Printed Electronics. Materials 2019, 12, 3477.
https://doi.org/10.3390/ma12213477.
29. Faruk, O.; Adak, B. Recent advances in PEDOT:PSS integrated graphene and MXene-based composites for electrochemical
supercapacitor applications. Synth. Met. 2023, 297, 117384. https://doi.org/10.1016/j.synthmet.2023.117384.
30. Tzaneva, B.; Aleksandrova, M.; Mateev, V.; Stefanov, B.; Iliev, I. Electrochemical Properties of PEDOT:PSS/Graphene
Conductive Layers in Artificial Sweat. Sensors 2024, 24, 39. https://doi.org/10.3390/s24010039.
31. Park, C.; Yoo, D.; Lee, J.J.; Choi, H.H.; Kim, J.H. Enhanced power factor of poly (3,4-ethyldioxythiophene):poly (styrene
sulfonate) (PEDOT:PSS)/RTCVD graphene hybrid films. Org. Electron. 2016, 36, 166–170.
https://doi.org/10.1016/j.orgel.2016.05.038.
Polymers 2024, 16, 1706 17 of 17

32. Yoo, D.; Kim, J.; Kim, J.H. Direct synthesis of highly conductive poly (3,4-ethylenedioxythiophene):poly (4-
styrenesulfonate)(PEDOT: PSS)/graphene composites and their applications in energy harvesting systems. Nano Res. 2014, 7,
717–730. https://doi.org/10.1007/s12274-014-0433-z.
33. Curto, V.F.; Fay, C.; Coyle, S.; Byrne, R.; O’Toole, C.; Barry, C.; Hughes, S.; Moyna, N.; Diamond, D.; Benito-Lopez, F. Real-time
sweat pH monitoring based on a wearable chemical barcode micro-fluidic platform incorporating ionic liquids. Sens. Actuators
B Chem. 2012, 171–172, 1327–1334. https://doi.org/10.1016/j.snb.2012.06.048.
34. Nasera, S.A.; Hameed, A.A.; Husseinc, M.A. Corrosion Behavior of Some Jewelries in Artificial Sweat. AIP Conf. Proc. 2020, 2213,
020030. https://doi.org/10.1063/5.0000111.
35. Vanhoestenberghe, A.; Donaldson, N. Corrosion of silicon integrated circuits and lifetime predictions in implantable electronic
devices. J. Neural Eng. 2013, 10, 031002. https://doi.org/10.1088/1741-2560/10/3/031002.
36. Da Silva, L.M.; De Faria, L.A.; Boodts, J.F.C. Determination of the morphology factor of oxide layers. Electrochim. Acta 2001, 47,
395–403. https://doi.org/10.1016/S0013-4686(01)00738-1.
37. Pan, Q.; Wu, Q.; Sun, Q.; Zhou, X.; Cheng, L.; Zhang, S.; Yuan, Y.; Zhang, Z.; Ma, J.; Zhang, Y.; et al. Biomolecule-friendly
conducting PEDOT interface for long-term bioelectronic devices. Sens. Actuators B Chem. 2022, 373, 132703.
https://doi.org/10.1016/j.snb.2022.132703.
38. Barron, S.L.; Oldroyd, S.V.; Saez, J.; Chernaik, A.; Guo, W.; McCaughan, F.; Bulmer, D.; Owens, R.M. A Conformable Organic
Electronic Device for Monitoring Epithelial Integrity at the Air Liquid Interface. Adv. Mater. 2023, 36, 2306679.
https://doi.org/10.1002/adma.202306679.
39. Magar, H.S.; Hassan, R.Y.; Mulchandani, A. Electrochemical impedance spectroscopy (EIS): Principles, construction, and
biosensing applications. Sensors 2021, 21, 6578. https://doi.org/10.3390/s21196578.
40. Rubinson, J.F.; Kayinamura, Y.P. Charge transport in conducting polymers: Insights from impedance spectroscopy. Chem. Soc.
Rev. 2009, 38, 3339–3347. https://doi.org/10.1039/B904083H.
41. Hernández, H.H.; Reynoso, A.M.R.; González, J.C.T.; Morán, C.O.G.; Hernández, J.G.M.; Ruiz, A.M.; Hernández, J.M.; Cruz,
R.O. Electrochemical Impedance Spectroscopy (EIS): A Review Study of Basic Aspects of the Corrosion Mechanism Applied to
Steels. In Electrochemical Impedance Spectroscopy; El-Azazy, M., Min, M., Annus, P., Eds.; IntechOpen: London, UK, 2020.
https://doi.org/10.5772/intechopen.94470.
42. Patra, S.; Munichandraiah, N. Supercapacitor studies of electrochemically deposited PEDOT on stainless steel substrate. J. Appl.
Polym. Sci. 2007, 106, 1160–1171. https://doi.org/10.1002/app.26675.
43. Guo, D.; Wang, L.; Wang, X.; Xiao, Y.; Wang, C.; Chen, L.; Ding, Y. PEDOT coating enhanced electromechanical performances
and prolonged stable working time of IPMC actuator. Sens. Actuators B Chem. 2020, 305, 127488.
https://doi.org/10.1016/j.snb.2019.127488.
44. Peringath, A.R.; Bayan, M.A.H.; Beg, M.; Jain, A.; Pierini, F.; Gadegaard, N.; Hogg, R.; Manjakkal, L. Chemical synthesis of
polyaniline and polythiophene electrodes with excellent performance in supercapacitors. J. Energy Storage 2023, 73, 108811.
https://doi.org/10.1016/j.est.2023.108811.
45. Dobashi, Y.; Fannir, A.; Farajollahi, M.; Mahmoudzadeh, A.; Usgaocar, A.; Yao, D.; Nguyen, G.T.M.; Plesse, C.; Vidal, F.;
Madden, J.D.W. Ion transport in polymer composites with non-uniform distributions of electronic conductors. Electrochim. Acta
2017, 247, 149–162. https://doi.org/10.1016/j.electacta.2017.06.141.
46. Lefrou, C.; Fabry, P.; Poignet, J.C. Electrochemistry: The Basics, with Examples; Springer: Berlin/Heidelberg, Germany, 2012.
47. Hong, W.; Xu, Y.; Lu, G.; Li, C.; Shi, G. Transparent graphene/PEDOT–PSS composite films as counter electrodes of dye-
sensitized solar cells. Electrochem. Comm. 2008, 10, 1555–1558. https://doi.org/10.1016/j.elecom.2008.08.007.
48. Zhou, J.; Anjum, D.H.; Lubineau, G.; Li, E.Q.; Thoroddsen, S.T. Unraveling the order and disorder in poly (3, 4-
ethylenedioxythiophene)/poly (styrenesulfonate) nanofilms. Macromolecules 2015, 48, 5688–5696.
https://doi.org/10.1021/acs.macromol.5b00851.
49. Thaning, E.M.; Asplund, M.L., Nyberg, T.A.; Inganäs, O.W.; von Holst, H. Stability of poly (3,4-ethylene dioxythiophene)
materials intended for implants. J. Biomed. Mater. 2010, 93B, 407–415. https://doi.org/10.1002/jbm.b.31597.

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury
to people or property resulting from any ideas, methods, instructions or products referred to in the content.

You might also like