Single Wall Carbon Nanotubes/Polypyrrole Composite Thin Film Electrodes: Investigation of Interfacial Ion Exchange Behavior

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Article

Single Wall Carbon Nanotubes/Polypyrrole Composite


Thin Film Electrodes: Investigation of Interfacial Ion
Exchange Behavior
Freddy Escobar-Teran 1,2 , Hubert Perrot 1 and Ozlem Sel 1, *

1 Laboratoire Interfaces et Systèmes Electrochimiques, LISE, UMR8235, Sorbonne Université, CNRS,


F-75005 Paris, France; [email protected] (F.E.-T.); [email protected] (H.P.)
2 Facultad de Ingeniería en Sistemas, Electrónica e Industrial, Universidad Técnica de Ambato,
Avenida los Chasquis y Río Payamino, Ambato 180103, Ecuador
* Correspondence: [email protected]

Abstract: Single-wall carbon nanotubes/polypyrrole (SWCNT/PPy) composite thin-film electrodes


were prepared by electrodeposition of the pyrrole monomer on a porous network made of SWCNT
bundles. Electrode/electrolyte interface, which is intimately related to the pseudocapacitive charge
storage behavior, is investigated by using coupled electrogravimetric methods (electrochemical
quartz crystal microbalance (EQCM) and its coupling with electrochemical impedance spectroscopy,
Ac-electrogravimetry), in a 0.5 M NaCl electrolyte (pH = 7). Our results show that the range of usable
potential is greater for composite SWCNT/PPy films than for SWCNT films, which should allow a
higher storage capacity to be obtained. This effect is also confirmed by mass variation measurements
via EQCM. The mass change (corresponding to the amount of (co)electroadsorbed species) obtained
with composite SWCNT/PPy films is four times greater than that observed for pristine SWCNT
films if the same potential range is examined. The permselectivity is also greatly improved in the
 case of composite SWCNT/PPy films compared to SWCNT films; the former shows mainly cation
 exchange preference. The quantities of anions estimated by Ac-electrogravimetric measurements are
Citation: Escobar-Teran, F.; Perrot, much lower in the case of composites. This corroborates the better permselectivity of these composite
H.; Sel, O. Single Wall Carbon SWCNT/PPy films even with a moderate amount of PPy.
Nanotubes/Polypyrrole Composite
Thin Film Electrodes: Investigation of Keywords: carbon/polymer composites; carbon nanotubes; conducting polymers; SWCNT; doped
Interfacial Ion Exchange Behavior. J. polypyrrole; electrochemical quartz crystal microbalance; EQCM; electrode/electrolyte interface;
Compos. Sci. 2021, 5, 25. https:// ion transfer
doi.org/10.3390/jcs5010025

Received: 16 December 2020


Accepted: 11 January 2021 1. Introduction
Published: 14 January 2021
Polymers and polymeric composites are a unique class of materials, extensively used in
numerous fields of applications such as electrochemical energy storage and conversion [1].
Publisher’s Note: MDPI stays neu-
tral with regard to jurisdictional clai-
Due to their peculiar physicochemical and mechanical properties, these materials have
ms in published maps and institutio-
been stimulating new research that aims at refining their structural, morphological and
nal affiliations. compositional characteristics to achieve the desired properties required for electrode or
electrolyte materials used in energy devices [2,3].
Among them, supercapacitors (SCs) can provide a large amount of power in a rela-
tively short time and can operate for a very high number of charge/discharge cycles and a
Copyright: © 2021 by the authors. Li- longer lifetime than batteries [4,5]. These versatile attributes have led to their incorpora-
censee MDPI, Basel, Switzerland. tion in a wide range of applications (such as city transit buses with stop-and-go driving,
This article is an open access article cranes and forklifts [6]). SC devices consist of two electrodes in contact with an electrolyte
distributed under the terms and con-
electrically isolated by a separator. During the charging process, the charges are stored and
ditions of the Creative Commons At-
separated across the two electrode/electrolyte interfaces, which provides electric energy
tribution (CC BY) license (https://
for the external load upon discharge. Briefly, SCs can be classified into two types with
creativecommons.org/licenses/by/
regard to their charge storage mechanism: electrical double-layer capacitors (EDLCs) and
4.0/).

J. Compos. Sci. 2021, 5, 25. https://doi.org/10.3390/jcs5010025 https://www.mdpi.com/journal/jcs


J. Compos. Sci. 2021, 5, 25 2 of 14

pseudocapacitors. EDLCs store charges by electrostatic adsorption of electrolyte ions at


the electrode–electrolyte interface, typically use carbon materials as electrodes [7–9]. As
for the pseudocapacitors, energy is stored by fast and reversible redox reactions between
the electrolyte and electroactive species on the electrode surface at characteristic potentials,
commonly employing transition metal oxides and conducting polymers (CPs) as electrode
materials [10,11].
Due to their high conductivity, porosity and superior electrochemical stability, carbon
materials are widely used as EDLC electrode materials [12], namely activated carbon
(ACs), carbon nanotubes (CNTs) and graphene are the most studied examples. CNTs
are of particular interest for the development of SC electrodes since they present unique
tubular structures and superior electrical conductivity, mechanical, thermal and chemical
stability [7]. CNTs have been considered as a promising candidate for high-power electrode
material because of the aforementioned properties. In spite of the fact that the surface area
of CNTs (800 m2 ·g−1 ) is small as compared to ACs (2000–3000 m2 g−1 ) [7,13], they can offer
a reasonable specific capacitance [14], which is speculated to stem from their aligned pore
structures, which, therefore, contribute to higher ion diffusion kinetics. Another option
is the use of CNTs as support for electroactive materials because of their high mechanical
resilience and open tubular network, which can form composite materials with metal
oxides and/or CPs [15,16].
Regarding the electrode materials for pseudocapacitors, CPs are a very attractive
solution due to their low price, non-toxicity and tunable chemical, electrical and phys-
ical properties. These polymers can be electrochemically doped in the presence of an
electrolyte to obtain very good electrical conductivity (10 to 100 S.cm−1 ) and a wide elec-
trochemical window. The most commonly used CPs for pseudocapacitive charge storage
involve poly(3,4-ethylenedioxythiophene) (PEDOT) [17], polypyrrole (PPy) [18], polyani-
line (PANI) [19], and polythiophene (PTh) [20]. Moreover, transitional metal oxides such
as ruthenium oxide (RuO2 ) [21], MnO2 [22], ZnO [23], and NiO [24] have also been widely
studied. Although the pseudo-capacitance can be much higher than EDL capacitance, it
suffers from low-power density and poor stability upon cycling, which further highlights
the interest of composite electrodes, uniting their benefits in a single electrode. In this
context, a mix of carbon materials and CPs, merging the advantages of the two components,
was evaluated in the field of pseudocapacitors. Indeed, a good electrical conductivity, a
high specific surface area, a combination of charge storage mechanisms (associated with
the electroadsorption and the Faradic reactions) and adequate control of the microstruc-
turation can lead to synergetic effects which can drastically improve the electrochemical
properties [25,26]. Different binary-composite films were assessed, employing various
methods of preparation: high specific capacitances were attained with PANI-CNTs [27] or
PEDOT-CNTs [28,29] composites; enhanced robustness of the flexible PPy-CNTs [30] or
improved capacitive property of PTh-CNTs films [31] has also been revealed. Nevertheless,
a deep understanding of the interfacial ionic transfer was not yet fully reached for these
binary-composite materials. This is the reason why new methods of investigation are
necessary to further improve their performances.
Taking into account the recent advances in the field of capacitive charge storage, on one
hand, fabrication of composite electrodes with high specific surface area appears to be a key
development path to achieve better SC performances. On the other hand, understanding
the principles governing the ionic exchange mechanisms in nanostructured supercapacitor
electrodes is equally important to shed light on this path and improve current devices.
Therefore, suitable morphological, compositional as well as electrochemical characteriza-
tion tools should be employed to assess composite electrode designs and the processes
occurring at the electrode/electrolyte interface. For this purpose, various sophisticated
analytical techniques have been developed to survey electrode–electrolyte interfaces that
mainly govern the electrochemical operation of energy devices. Among them, the elec-
trochemical quartz crystal microbalance (EQCM) [32–36] provides gravimetric changes of
electrodes during electrochemical characterizations (such as charge/discharge) and offers
J. Compos. Sci. 2021, 5, 25 3 of 14

valuable insights of the interfaces, once the required conditions of the deposited film in
terms of viscoelastic and hydrodynamic properties are included [37–39].
Therefore, this work focuses on the understanding of the processes occurring at the
electrode/–electrolyte interfaces of composite electrode materials by using EQCM-based
coupled electrochemical methods. Carbon- and CP-based composites (SWCNT/PPy) are
prepared by the electrochemical synthesis of doped polypyrrole on the single-wall CNT-
based networks, the latter providing a high specific surface area. Interfacial properties are
characterized not only with gravimetric EQCM but also with its coupling with electrochem-
ical impedance spectroscopy (EIS), the so-called Ac-electrogravimetry [40,41]. This comple-
mentary method can deliver unique species-selective and frequency-dependent informa-
tion about the ionic exchange mechanisms at the electrode/electrolyte interface [42–44].
The processes occurring at SWCNT/PPy composite electrode/electrolyte interface, which
are related to their pseudo-capacitive charge storage behavior, are discussed in comparison
to their pristine components (i.e., SWCNT [32] and PPy [45,46]).

2. Materials and Methods


2.1. Materials
Single-wall CNT (755117-1G, length: 300–2300 nm and diameter: 0.7–1.1 nm), sodium
dodecyl sulfate (NaDS), pyrrole monomer, sodium chloride and N-methylpyrrolidone
(NMP) solvent were acquired from Sigma-Aldrich. Poly(vinylidene fluoride)-co- hexafluo-
ropropylene (PVDF-HFP) is obtained from Solef® 21508, Solvay Solexis, Milan, Italy.

2.2. SWCNT/Polypyrrole Composite Thin Film Electrode Preparation


In the first step, the SWCNT film preparation was performed based on a method
described in the literature [47,48]. SWCNTs were deposited by the “drop-casting” on a
gold electrode (effective surface area of 0.2 cm2 ) of a quartz crystal resonator (9 MHz,
AW Sensors, Valencia, Spain), from a slurry containing 90% SWCNT powder and 10%
PVDF-HFP (Poly(vinylidene fluoride-hexafluoropropylene)) polymer binder in N-methyl-
2-pyrrolidone. A slurry containing 9 mg SWCNT powder, 1 mg of PVDF-HFP in 10 mL
of NMP was ultrasonicated for about 20 min. Around 8 µL of this slurry was deposited
on the gold electrode of the QCM. Then, the carbon films were subjected to a heat treat-
ment at 120 ◦ C for 30 min, with a heating rate of ~5 ◦ C·min−1 to evaporate the residual
solvent and improve the adhesion properties of the films on a gold electrode. The film
thickness was estimated to be around 500 nm (based on field emission gun scanning elec-
tron microscopy (FEG-SEM) analyses, Figure 1). In a second step, the composite films of
SWCNT/polypyrrole were obtained, following a method described in the literature [49].
The procedure consists of cycling the SWCNT electrode, previously prepared, in a solution
containing 0.1 M pyrrole (freshly distilled) and 0.05 M sodium dodecyl sulfate (NaDS), at a
scan rate of 0.01 V·s−1 in a potential range from 0 V to 0.675 V vs. Ag/AgCl during 2 cycles.
The resulting films are taken out of the 3-electrode cell, rinsed with double distilled water
and dried by purging with N2 flow.

2.3. Morphological and Physical Characterizations


Prior to the thin film preparation, the SWCNT powders were characterized using
nitrogen adsorption (the specific surface area was estimated at ~842 m2 ·g−1 , by using
Brunauer–Emmett–Teller (BET) analysis), X-ray diffraction (XRD) and high-resolution
transmission electron microscopy (HRTEM). The morphological and structural observa-
tions, not shown here, were already given in Reference [34]. The surface morphology of
the SWCNT/polypyrrole, after thin-film formation on the gold electrode of the resonators,
was investigated by field emission gun scanning electron microscopy (FEG-SEM) (Zeiss,
Supra 55, Oberkochen, Germany).
J. Compos. Sci. 2021, 5, 25 4 of 14
J. Compos. Sci. 2021, 5, x 4 of 14

Figure
Figure 1.
1. FEG-SEM
FEG-SEM images
images of
of single-wall
single-wall carbon nanotubes/polypyrrole (SWCNT
carbon nanotubes/polypyrrole /PPy) nanocomposite
(SWCNT /PPy) nanocomposite film
film deposited
deposited on
on
the gold electrode of a quartz resonator: (a) surface morphology with an inset showing the PPy clusters and (b) cross-
the gold electrode of a quartz resonator: (a) surface morphology with an inset showing the PPy clusters and (b) cross-section
section (red arrows point out the PPy clusters and orange colored arrows show the average film thickness).
(red arrows point out the PPy clusters and orange colored arrows show the average film thickness).

2.4. Electrochemical and Electrogravimetric Characterizations


2.4. Electrochemical and Electrogravimetric Characterizations
EQCM measurements were performed in 0.5 M NaCl (at pH = 7) in a three-electrode
EQCM measurements were performed in 0.5 M NaCl (at pH = 7) in a three-electrode
configuration. A lab-made QCM device was used to measure the microbalance frequency
configuration. A lab-made QCM device was used to measure the microbalance frequency
shift (∆f). Gold deposited on the quartz resonator was used as the working electrode. Plat-
shift (∆f ). Gold deposited on the quartz resonator was used as the working electrode.
inum grid and Ag/AgCl (3 M KCl) was used as a counter and a reference electrode, re-
Platinum grid and Ag/AgCl (3 M KCl) was used as a counter and a reference electrode,
spectively. The gravimetric conditions for the QCM analyses were assured by keeping
respectively. The gravimetric conditions for the QCM analyses were assured by keeping
film thickness acoustically thin (df < 500 nm) (Figure 1).
film thickness acoustically thin (df < 500 nm) (Figure 1).
Ac-electrogravimetric measurements were also performed in the same three-elec-
Ac-electrogravimetric measurements were also performed in the same three-electrode
trode configuration described above. A four-channel frequency response analyzer (FRA,
configuration described above. A four-channel frequency response analyzer (FRA,
Solartron 1254) and a lab-made potentiostat (SOLETEM-PGSTAT) were used. The QCM
Solartron 1254) and a lab-made potentiostat (SOLETEM-PGSTAT) were used. The QCM
was
was performed
performed under
under aa dynamic
dynamic regime,
regime, and
and the
the modified
modified working
working electrode
electrodewas
waspolar-
polar-
ized at selected potentials to which a sinusoidal small-amplitude potential perturbation
ized at selected potentials to which a sinusoidal small-amplitude potential perturbation
was
was superimposed.
superimposed. The The frequency
frequency range
range was
was between
between 63 63 kHz and 10
kHz and 10 mHz.
mHz. The mass
The mass
change, Δm, of the working electrode was measured simultaneously with the
change, ∆m, of the working electrode was measured simultaneously with the AC response, AC re-
sponse, ΔI,electrochemical
∆I, of the of the electrochemical
system,system, which
which led led electrogravimetric
to the to the electrogravimetric (mass/po-
(mass/potential)
∆𝑚 ∆𝐸
transfertransfer
tential) (TF), ∆m
functionfunction (ω ), and(𝜔)
∆E(TF), the electrochemical impedance,
, and the electrochemical ∆E
∆I ( ω ), to be
impedance, (𝜔), to
obtained
∆𝐸 ∆𝐼
simultaneously at a given potential and frequency modulation, f (pulsation, ω
be obtained simultaneously at a given potential and frequency modulation, f (pulsation, = 2πf). The
operation principle and the details of the Ac-electrogravimetric measurement
ω = 2πf). The operation principle and the details of the Ac-electrogravimetric measure-setup have
been described
ment setup haveinbeen
previous papers
described in [40,41,50].
previous papers [40,41,50].
3. Results and Discussion
3. Results and Discussion
3.1. Morphological Aspects of the SWCNT/PPy Composites
3.1. Morphological Aspects of the SWCNT/PPy Composites
The SWCNT/PPy composite electrodes are prepared in a two-step process. In a first
step,The SWCNT/PPy
a thin layer of SWCNTcomposite electrodes
is formed aregold
on the prepared in a of
electrode two-step
the quartzprocess. In a first
resonators by
step,
drop-casting a slurry containing SWCNT and PVDF-HFP as an active material and by
a thin layer of SWCNT is formed on the gold electrode of the quartz resonators as
drop-casting a slurry containing
a binder, respectively. SWCNT and
The morphological PVDF-HFPof
characteristics asthe
an SWCNT-modified
active material andQCM as a
binder,
resonatorsrespectively. The morphological
were previously discussed [34]. characteristics of the SWCNT-modified
Briefly, N2 adsorption analysis indicates QCMthatres-
the
onators were previously discussed [34]. Briefly, N adsorption analysis
SWCNTs have a pore diameter of 1–2 nm and a high specific surface area (~842 m ·g 1 ).
2 indicates that
2 −
the
SWCNTs
Thin layers have a poreusing
formed diameter
theseofSWCNTs
1–2 nm and a high
present specific of
a network surface
CNT area (~842made
bundles, m2·g−1of).
Thin layers
at least formed
~10–20 CNTs using
[34].these
In the SWCNTs
second present a network of CNT
step, SWCNT-modified QCMbundles, made were
resonators of at
least
used~10–20 CNTs [34].
as a working In thefor
electrode second step, SWCNT-modified
the electropolymerization ofQCM
pyrroleresonators
monomer, were used
leading
as
to athe
working
SWCNT/PPyelectrode for the electropolymerization
composites. The surface morphology of pyrrole monomer, leading to
of SWCNT/polypyrrole the
(PPy)
SWCNT/PPy
thin films wascomposites.
characterized Thebysurface
FEG-SEM,morphology
as shown ofinSWCNT/polypyrrole
Figure 1. The images (PPy)
reveal thina
films was characterized
high-density of CNT bundlesby FEG-SEM,
as it wasas already
shown in seenFigure
under 1. the
Thesame
images reveal
film a high-
preparation
density
conditionsof CNT
[34].bundles
After twoas itcycles
was already
(cyclicseen under the same
voltammetry) of PPyfilm preparation conditions
electrodeposition on the
[34]. After two cycles
SWCNT-based (cyclicthe
electrodes, voltammetry)
formation ofofpolypyrrole
PPy electrodeposition
clusters with on typical
the SWCNT-based
cauliflower
or nodularthe
electrodes, morphology
formation is ofobserved
polypyrrole in agreement
clusters with with the previous
typical cauliflower studies [45], which
or nodular mor-
phology is observed in agreement with the previous studies [45], which is more evident
mpos. Sci. 2021, 5, x 6 of 14

J. Compos. Sci. 2021, 5, case


25 because the charge compensation can also be due to the concomitant insertion of anions 5 of 14
or to solvated cations or to a mixture of these species. Dynamic approaches have the poten-
tial to reach a fair separation of these processes (identification by the kinetic transfers). For
this reason, Ac-electrogravimetric investigations were performed to complement the EQCM
is more evident on certain locations on the film surface (Figure 1a inset and Figure 2b,
and will be discussed in Section 3.3.
indicated with red arrows).

500 a 8 b 25 mV.s
-1
-2
Current / µA.cm

-1
50 mV.s

-2
6 -1

m/µg.cm
0 100 mV.s

4
-500
-1
25 mV.s
-1 2
50 mV.s
-1
-1000 100 mV.s
0
-1.2 -0.8 -0.4 0.0 0.4 -1.2 -0.8 -0.4 0.0 0.4
E/V vs. Ag/AgCl E/V vs. Ag/AgCl

-1
Specific Capacitance / F.g
0 50
c d
-1
Fdm/dq / g.mol

-50 25

0
-100
-25
-150
25 mV.s
-1 -50 -1
25 mV.s
-1 -1
-200 50 mV.s -75 50 mV.s
-1 -1
100 mV.s 100 mV.s
-100
-250
-1.0 -0.8 -0.6 -0.4 -0.2 0.0 0.2 0.4 -1.2 -0.8 -0.4 0.0 0.4
E/V vs. Ag/AgCl E/V vs. Ag/AgCl
Figure 2. EQCM results
Figure of SWCNT/PPy
2. EQCM results of composite
SWCNT/PPy thincomposite
films: (a) current vs. (a)
thin films: potential;
current(b)
vs.mass variation
potential; vs. potential;
(b) mass
(c) calculated F dm values vs. potential; (d) specific 𝑑𝑚
capacitance values (calculated from panel a) vs. potential. Measurements
dq
variation vs. potential; (c) calculated 𝐹 values vs. potential; (d) specific capacitance values (cal-
in aqueous electrolyte 0.5 M NaCl pH 7.0 at 3 different𝑑𝑞 scan rates.
culated from panel a) vs. potential. Measurements in aqueous electrolyte 0.5 M NaCl pH 7.0 at 3
different scan rates.The cross-section of the SWCNT/PPy shown in Figure 1b permits the composite
thin-film thickness (df ) to be estimated. The thickness is slightly inhomogeneous along
Figure 2d
thedepicts the and
thin film variation of specific
presents capacitance
a variation from 450 values as anm.
to 550 function
The of the ap-
electrodeposited PPy
plied potential under the(only
component sametwoconditions
cycles ofaselectrodeposition)
mentioned previously. on theThe gravimetric
SWCNTs spe-significantly
does not
cific capacitance
changevalues (Cs) thickness,
the film are estimated and to be around
therefore, the 25 to 30 F·g−1conditions
gravimetric in the range arefrom
kept -for the QCM
1.0 V to +0.45based
V vs.analysis.
Ag/AgCl.ItFor very cathodic potentials, reduction of water interferes
is important to note that the composite film thickness is mainly in governed
the CV curves and thus, strongly modify the C values. Otherwise, these C are in
by the SWCNT layer under these experimental conditions. Although the exact location of
s s the same
order of magnitude as that obtained
PPy electrodeposition with pristine
is difficult SWCNT it
to pronounce, thin-film electrodes
likely occurs on the[34,51,52].
porous surface of the
A smallnetwork
dependence
formedon bythethe
scan rate values
SWCNTs, is than
rather observed
formingin the EQCMcompact
a second responses,
layer on top of
the former. 𝑑𝑚
which is more pronounced in the 𝐹 the function is shown in Figure 2c. This result
𝑑𝑞
suggests that3.2.
depending
EQCM Study on the scan
of the rate, different
SWCNT/PPy species may contribute to the electro-
Composites
adsorption and/or faradaicEQCM
Classical processes. Indeed,
studies were the information
already performedprovided by the
for pristine classical
SWCNT films in NaCl
EQCM is insufficient for a thorough explanation of the transfer mechanism of the different
electrolytes, as shown in [32,34]. In these experiments, the appropriate potential range
species. As Ac-electrogravimetry
was selected from –0.4 wasVproven to be a strong
vs. Ag/AgCl to 0.4complementary
V vs. Ag/AgCl. tool The
to EQCM
current response
in the previous CNT studies [32–34], the composite films were also analyzed
reveals a classic square shape with bumps, while the mass change corresponds with this to a V-shape
technique. As mentioned
type: earlier,part,
in the cathodic this acoupled tool cancontribution
cation/solvent deliver unique species-selective
is observed, and on the contrary,
and frequency-dependent
an anion/solvent information
contributionabout
wasthe ion exchange
distinguished in mechanisms at theAelec-
the anodic region. permselectivity
trode/electrolyte interface.
failure, indicating a mixed contribution of cations and anions, is also seen around the point
of zero charges (PZC)/point of zero mass (PZM), corresponding to a potential of +0.1 V vs.
Ag/AgCl. The maximum mass variation is less than 1 µg·cm−2 in the case of the cation
response and starting from this potential reference. In light of these results, an EQCM
J. Compos. Sci. 2021, 5, 25 6 of 14

analysis of the SWCNT/PPy composites was performed. The SWCNT/PPy composite


film is oxidized and reduced between 0.45 V and −1.2 V vs. Ag/AgCl at three different
scan rates of 25 mV·s−1 , 50 mV·s−1 and 100 mV·s−1 in 0.5 M NaCl solution (pH = 7).
Figure 2a,b show the corresponding current and the mass changes, respectively. A slight
contribution of the PPy component is apparently observed in the CV responses (Figure 2a),
as compared to that of pristine SWCNTs [32]. Indeed, the current evolution is smoother or
less bumpy in the case of the composite film, for example, when the region around −0.2 V
vs. Ag/AgCl is examined. On the contrary, the mass response of the film is totally different
in terms of amplitude and shape. Previous EQCM studies on pristine SWCNT in the same
electrolyte have shown a V-shaped mass response around the PZC, which is around 0.1 V
vs. Ag/AgCl [32]. Instead, here, for the composite film, the mass response shows only
one slope; even a larger potential range has been examined. When the composite film
was reduced, a larger mass increase was observed: a factor of four is observed between
the amplitudes of composite and pristine SWCNTs for the same potential range. On the
other hand, the anion contribution is completely removed here for the same potential
range (Figure 2b), which was observed for the pristine SWCNT film from +0.2 to +0.4 V
vs. Ag/AgCl. When the composite film is oxidized, the mass decreases all the time, while
in the case of pristine SWCNTs, the mass increases for the higher potentials revealing an
anion contribution, as already demonstrated [32]. These qualitative remarks already show
that the presence of PPy changes
 the interfacial
 ion-exchange behavior of the SWCNTs.
Figure 2c shows the F dm dm 1
dq = F dt × i function calculated from the EQCM data as a
function of potential, which is equivalent to the global molar mass of the species obtained
from the reduction branch in the range from 0.4 V to −1.2 V vs. Ag/AgCl in 0.5 M NaCl. At
the cathodic potential region, the values vary in the range of 50–200 g·mol−1, which could
correspond to Na+ with a high number of hydration or Na+ accompanied by free solvent
molecules. However, at the anodic potentials, the values are smaller, and they vary in the
range 0–50 g·mol−1 , which is also difficult to be attributed to a single ion contribution
(i.e., 23 g·mol−1 and 35 g·mol−1 , for Na+ and Cl− , respectively). Unfortunately, using
EQCM alone to understand the interfacial process appears to be limited in the present case
because the charge compensation can also be due to the concomitant insertion of anions or
to solvated cations or to a mixture of these species. Dynamic approaches have the potential
to reach a fair separation of these processes (identification by the kinetic transfers). For this
reason, Ac-electrogravimetric investigations were performed to complement the EQCM
and will be discussed in Section 3.3.
Figure 2d depicts the variation of specific capacitance values as a function of the
applied potential under the same conditions as mentioned previously. The gravimetric
specific capacitance values (Cs ) are estimated to be around 25 to 30 F·g−1 in the range
from −1.0 V to +0.45 V vs. Ag/AgCl. For very cathodic potentials, reduction of water
interferes in the CV curves and thus, strongly modify the Cs values. Otherwise, these
Cs are in the same order of magnitude as that obtained with pristine SWCNT thin-film
electrodes [34,51,52].
A small dependence on the scan rate values is observed in the EQCM responses,
which is more pronounced in the F dm dq the function is shown in Figure 2c. This result
suggests that depending on the scan rate, different species may contribute to the electroad-
sorption and/or faradaic processes. Indeed, the information provided by the classical
EQCM is insufficient for a thorough explanation of the transfer mechanism of the dif-
ferent species. As Ac-electrogravimetry was proven to be a strong complementary tool
to EQCM in the previous CNT studies [32–34], the composite films were also analyzed
with this technique. As mentioned earlier, this coupled tool can deliver unique species-
selective and frequency-dependent information about the ion exchange mechanisms at the
electrode/electrolyte interface.
The measurements were performed at each 200 mV in the range from 0.4 V to −1.2 V
vs. Ag/AgCl in 0.5 M NaCl electrolyte. Figure 3 shows an example of the experimental
and theoretical electrochemical transfer functions (TFs) obtained from Ac-electrogravim-
etry of an SWCNT/PPy thin film in 0.5 M NaCl electrolyte at −0.4 V vs. Ag/AgCl. The
J. Compos. Sci. 2021, 5, 25 7 of 14
experimental data were fitted according to the model presented briefly in Appendix A
(theoretical part). More details about the theory can be found in previous works [40,41].
A Mathcad software® was used to perform this step, assuring a good fitting for all the
transfer functions3.3. Ac-Electrogravimetric
measured Study of the
at a given potential andSWCNT/PPy
by using the Composites
same set of parameters.
With the Ac-electrogravimetric approach,
A good agreement between experimental data and theoretical curves an EQCM response aredeconvolution
evident in was per-
formed with the pristine SWCNT, ∆𝐸 as shown in references [32,34]. Briefly, the contribution
Figure 3. The electrochemical impedance, (𝜔) (Figure 3a), present a slightly distorted
of Na+ .H2 O, H+ and Cl− were ∆𝐼 detected when the surface is negatively and positively
straight line at thecharged,
lower frequency
respectively.domain,
The fasterindicating
interfacialthat there
transfer is a multi-ion
is obtained for the transfer
Na+ .H2 O and Cl−
contribution. Therefore,
species. it
H2isO difficult to extract
was observed at lowerinformation from this
kinetics of transfer, transfer
whereas +
H wasfunction.
seen at very low
values. In terms of ∆𝑞
quantities, major contributions are due to hydrated sodium
The charge/potential transfer function, (𝜔) (Figure 3b), permit an easier separation of with a
∆𝐸
quasi-equivalent contribution of water in the cathodic part and a lesser proton and chloride
the ionic contributions, however, without any possibility to identify the ionic species in-
contributions. These findings are compared below to those obtained with the SWCNT/PPy
volved. Figure 3bcomposite
shows two loops, which can be attributed to at least two species, where
electrodes.
their time constants are not sufficiently
The measurements were different from at
performed each
eachother
200 mVto obtain perfectly
in the range sep-V to −1.2 V
from 0.4
arated loops. Here,vs.aAg/AgCl
big loop in is 0.5
observed
M NaCl around a characteristic
electrolyte. Figure 3 showsfrequency
an exampleofof10 theHz, fol-
experimental and
lowed by a smaller one appearing
theoretical at low frequencies.
electrochemical Equations
transfer functions (A1) and
(TFs) obtained (A2)
from given in
Ac-electrogravimetry
Appendix A wereofused an SWCNT/PPy thin film
here for the fitting in 0.5 assuming
process, M NaCl electrolyte at −0.4 V
the contribution ofvs.twoAg/AgCl.
ions The
∆𝐸 ∆𝑞
named c1 and c2. experimental
At this level,data
these were
twofitted according
transfer to the model
functions, (𝜔) presented
and briefly
(𝜔), in Appendix A
are simi-
(theoretical part). More details about the theory∆𝐼can be found ∆𝐸in previous works [40,41].
lar to those obtained with pristine
A Mathcad software SWCNTs
® was used under the same
to perform thisexperimental
step, assuring conditions
a good fittingof for all the
measurements [32]. transfer functions measured at a given potential and by using the same set of parameters.

3. Experimental
FigureFigure and theoretical
3. Experimental and electrochemical data of the SWNT/PPy
theoretical electrochemical data of thin
the film in 0.5 M NaCl
SWNT/PPy thinmeasured at −M
film in 0.5 0.4 V vs.
∆E ∆q
Ag/AgCl: (a) electrochemical impedance, ∆I (ω ) and (b) charge/potential transfer function,
∆𝐸 (ω ). Theoretical functions
NaCl measured at −0.4 V vs. Ag/AgCl: (a) electrochemical impedance,
were calculated with the following parameters: df = 0.4 µm, Kc1 = 9.50 × 10−5 cm·s−1 , G∆𝐼
(𝜔)∆Eand −(b) charge/po-
9 mol ·s−1 ·cm−2 ·V−1 ,
c1 = 7.98 × 10
− 3 − 1 ∆𝑞 − 7 − 1 − 2 − 1
Kc2 = 3.46 × 10transfer
tential cm·s function,
and Gc2 = 9.3(𝜔)
× 10 mol·s ·cm
. Theoretical ·V . were calculated with the following param-
functions
∆𝐸
eters: df = 0.4 µ m, Kc1 = 9.50 × 10−5agreement
A good cm·s−1, Gc1 between
= 7.98 × 10 −9 mol·s−1·cm−2·V−1, Kc2 = 3.46 × 10−3 cm·s−1 and
experimental data and theoretical curves are evident in
Gc2 = 9.3 × 10 mol·sFigure
−7 −1 ·cm ·3.
−2 −1. electrochemical impedance, ∆E (ω ) (Figure 3a), present a slightly distorted
V The
∆I
straight line at the lower frequency domain, indicating that there is a multi-ion transfer
contribution. Therefore, it is difficult to extract information from this transfer function.
∆q
The charge/potential transfer function, ∆E (ω ) (Figure 3b), permit an easier separation
of the ionic contributions, however, without any possibility to identify the ionic species
involved. Figure 3b shows two loops, which can be attributed to at least two species,
where their time constants are not sufficiently different from each other to obtain perfectly
separated loops. Here, a big loop is observed around a characteristic frequency of 10 Hz,
followed by a smaller one appearing at low frequencies. Equations (A1) and (A2) given
in Appendix A were used here for the fitting process, assuming the contribution of two
∆q
ions named c1 and c2. At this level, these two transfer functions, ∆E∆I ( ω ) and ∆E ( ω ), are
similar to those obtained with pristine SWCNTs under the same experimental conditions
of measurements [32].
8 of 14
J. Compos. Sci. 2021, 5, 25 8 of 14

∆𝑚
In the mass/potential transfer function, (𝜔), one big loop appears in the third
∆𝐸
In the mass/potential transfer function, ∆m ∆E ( ω ), one big loop appears in the third
quadrant (Figure 4), which is characteristic for cation/solvated cation contributions or free
quadrant (Figure 4), which is characteristic for cation/solvated cation contributions or free
solvent molecules in the same flux direction. In the fitting process of the experimental
∆𝑚 solvent molecules in the same flux direction. In the fitting process of the experimental
(𝜔), three species are considered:
∆m Na+.nH2O at higher frequencies, free H2O molecules
∆𝐸
∆E ( ω ) , three species are considered: Na+ .nH2 O at higher frequencies, free H2 O molecules
at intermediate frequencies at
and H at lowerfrequencies
+
intermediate frequencies.and
TheHidentification of these species
+ at lower frequencies. The identification of these species
was achieved by the determination
was achieved by the determination of their molar mass, Mi , in
of their molar mass, M i, using Equation (A3) given using Equation (A3) given in
the theoretical part of Appendix A. The two
the theoretical partparameters
of AppendixKi and Gi, were
A. The already given
two parameters by Gi, were already given by
Ki and
the electrochemical response thefor the two ionic species.
electrochemical responseForforthe
thefree
two solvent, Ks andFor
ionic species. Gs the
are free solvent, Ks and Gs are
∆𝑚
only estimated through theonlymass/potential transfer the
estimated through function, (𝜔). transfer function, ∆m
mass/potential ∆E ( ω ).
∆𝐸

∆m
Figure 4. Experimental and theoretical electrogravimetric ∆𝑚 transfer function, ∆E (ω ), of the SWNT/PPy
Figure 4. Experimental and thin theoretical
film in 0.5 electrogravimetric
M NaCl measuredtransfer at −0.4 function,
V vs. Ag/AgCl. (𝜔), of the function was calculated with
Theoretical
∆𝐸
SWNT/PPy thin film in 0.5 M the NaCl measured at -0.4 V vs. Ag/AgCl. Theoretical − 5
following parameters: df = 0.4 µm, Kc1 = 9.50 × 10 cm·s , Gc1cal-
function − 1
was = 7.98 × 10−9 mol·s−1 ·cm−2 ·V−1 ,
culated with the following parameters:
Kc2 = 3.46 × 10 d f = 0.4
− 3 µ m, K
−1 = 9.50 ×
cm·s , Gc2 = 9.3 × 10
c1 10 −5 cm·
− s
7 −1, Gc1 −
= 1 7.98 ×2 10 −9
mol·s ·cm ·V−1 , Ks = 1.97 × 10−3 cm·s−1 ,

mol·s−1·cm−2·V−1, Kc2 = 3.46 × 10−3Gcm· s −1, Gc2 = 9.3 × 10−7 mol·s−1·cm−2·V−1, Ks = 1.97 × 10−3 cm·s−1, Gs = 1.99
− 6 mol·s , Mc1 = 1 g·mol , Mc2 = 23 + 18 g·mol and Ms = 18 g·mol−1 .
− 1 − 1 − 1
s = 1.99 × 10
× 10−6 mol·s−1, Mc1 = 1 g·mol−1, Mc2 = 23 + 18 g·mol−1 and Ms = 18 g·mol−1.
The contribution of three different species in the charge compensation process, es-
The contribution of three different
timated species
by fitting theinexperimental
the charge compensation
data, was furtherprocess, esti-
confirmed by carefully analyzing
mated by fitting the experimental data, was further confirmed by carefully
the partial electrogravimetric transfer function. For example, analyzing the the cation 1, c1, the con-
partial electrogravimetric transfer function. For example, ∆m the
c2s cation 1, c1, the contribution
tribution is removed and ∆E ( ω ) is calculated or the cation 2, c2, the contribution is

∆𝑚 𝑐2𝑠
is removed and | (𝜔) is calculated or the c1scation 2, c2, the contribution is removed
∆𝐸
∆𝑚 𝑐1𝑠 removed and ∆m ∆E ( ω ) is calculated (Equations (A3) and (A4)). Figure 5a,b exhibit a

and | (𝜔) is calculated (Equations (A3) and (A4)). Figure 5a,b exhibit a good agree-
∆𝐸 good agreement between the theoretical and experimental data. These partial electrogravi-
ment between the theoretical andtransfer
metric experimental
functions data. Theseapartial
provide electrogravimetric
crosscheck for validating the hypothesis involving
transfer functions provide athree
crosscheck
different species and a better separation of thethree
for validating the hypothesis involving various dif-contributions. If the molar
ferent species and a better masses
separation of the various contributions. If the molar masses
chosen to identify the species are different from reality, a good agreement cannot
chosen to identify the species are different from reality, a good agreement ∆m
c1scannot be ob-
be obtained.
∆𝑚 𝑐1𝑠 It should be noted that for ∆E ( ω ), a small loop appears in the fourth
tained. It should be noted that for | (𝜔), a small loop appears in the fourth quadrant
quadrant ∆𝐸 at low frequencies. According to our calculation, this response is related to the
at low frequencies. Accordingproton to our calculation,
contribution, and this responsethe
normally, is cation
relatedresponse
to the protonis located at the third quadrant
contribution, and normally,of the cation response is located at the third
the Cartesian coordinates system. This difference quadrant ofisthe Car-
due to the mathematic treatment
tesian coordinates system. This difference
observed is due to
in Equation (A4)thewhere
mathematic
a factortreatment
of Mc1 -Mc2 observed
appears;inin our case, a negative value
Equation (A4) where a factor of Mc1-Mc2asappears;
is obtained the molar in mass
our case,
of c2a(hydrated
negative valuesodium) is obtained
is higher than the molar mass of c1
as the molar mass of c2 (hydrated sodium) is higher than the molar mass of c1 (proton).
(proton). Thus, the response is located at this part of the complex plane, which appears as
Thus, the response is located at this part
a virtual anionofcontribution.
the complex Since plane,wewhich
haveappears
a perfect asagreement
a virtual for all the TFs, our model
anion contribution. Since we have a perfect agreement for all the TFs,
used in the fitting process is confirmed to be valid at this our model used in
potential (−0.4 V vs. Ag/AgCl).
the fitting process is confirmed to be validwas
This procedure at this potential
followed (−0.4 V vs.
for different Ag/AgCl).
potentials This
in the potential range used for the
procedure was followed forEQCM different potentials in the
measurements. Thepotential range used
key parameters, for
Ki , G the M
i and EQCM
i, are obtained for all the species
measurements. The key parameters,
involved K i, Gi and
directly orMindirectly
i, are obtained
in thefor all thecompensation
charge species involved process.
directly or indirectly in the charge compensation process.
pos. Sci. 2021, 5, x 9 of 14

J. Compos. Sci. 2021, 5, 25 9 of 14

Figure 5. Partial electrogravimetric transfer function of the SWNT-PPy thin film in 0.5 M NaCl measured at −0.4 V vs.
Figure 5.c1sPartial electrogravimetric transfer function of the SWNT-PPy
c2s
∆m thin film in 0.5 M NaCl

Ag/AgCl: (a) ∆m ∆E ( ω ) cation1-solvent partial∆𝑚 transfer
𝑐1𝑠 function and (b) ∆E (ω ) cation2-solvent partial transfer
measuredfunctions
function. Theoretical at −0.4 Vwere
vs. Ag/AgCl:
calculated
∆𝐸
(𝜔)setcation1-solvent
(a)with |the same of parameterspartial transfer function
used previously and presented in
for the data
Figures 3 and 4. ∆𝑚 𝑐2𝑠
(b) | (𝜔) cation2-solvent partial transfer function. Theoretical functions were calculated
∆𝐸
For potentials
with the same set of parameters usedmore cathodic
previously than
for the −0.3
data V vs. Ag/AgCl,
presented in Figures the contribution
3 and 4. of only three
species is found mixing the two cations, Na+ .nH2 O and H+ , and free solvent molecules
following
For potentials morethe same transfer
cathodic than −0.3direction. On thethe
V vs. Ag/AgCl, contrary, for more
contribution anodic
of only potential values,
three
species is found the mixing
proton contribution
the two cations, vanishes,
Na+.nH and
2Ochloride
and H+, species
and free appear
solventat very low frequencies.
molecules
following the sameFigure transfer 6a,b display On
direction. howthe thecontrary,
kinetic transfer
for more rates, Ki , and
anodic values,resistance, Rti ,
the transfer
potential
of the species evolve as a function of the applied potential.
the proton contribution vanishes, and chloride species appear at very low frequencies. The potential range corresponds
to the conditions where the electroadsorption/desorption or insertion/expulsion
Figure 6a,b display how the kinetic transfer rates, Ki, and the transfer resistance, Rti, processes
of the thin film can be observed.
of the species evolve as a function of the applied potential. The K i values presented in Figure 6a
The potential range corre- indicate that the
+
sponds to the Na .nH2 O ion
conditions is the
where thefastest of the four species. This
electroadsorption/desorption orobservation may be correlated to
insertion/expulsion
processes of the an thin
easier dehydration
film process
can be observed. Theof K
the sodium
i values species in
presented compared
Figure 6atoindicate
the protons. For the
Sci. 2021, 5, x latter ions, it is noted that their corresponding interfacial transfer
that the Na .nH2O ion is the fastest of the four species. This observation may be correlated
+ occurs at low
10frequencies.
of 14
Additionally, the concentration of the protons in the electrolyte
to an easier dehydration process of the sodium species compared to the protons. For the is much lower compared to
that of sodium species at this pH value.
latter ions, it is noted that their corresponding interfacial transfer occurs at low frequen-
cies. Additionally, the concentration of the protons in the electrolyte is much lower com-
pared to that of sodium species at this pH value.
Furthermore, the kinetic transfer rates, Ks, of free water molecules are somewhat close
to the values of the Na+.nH2O ions. As for the Cl− ions, their transfer occurs at low frequen-
cies and only at more anodic potentials.
In SWCNT/PPy composites, the kinetic transfer rates, Ki, of Na+.nH2O are signifi-
cantly lower than the Ki of the Na+.nH2O in pristine SWCNTs [32] and higher than the Ki
of the Na+ species detected in pristine PPy [45,49], measured under similar conditions.
These results indicate that the composite structure does not hinder the sodium species
transfer, and probably it is governed by the insertion process in PPy rather than a faster
electroadsorption/desorption process. Additionally, the contribution of the sodium spe-
cies to the charge compensation process (both electrostatic and faradaic) is enlarged to the
whole potential range, i.e., 0.4 V to −1.2 V vs. Ag/AgCl (larger than that used for SWCNTs
alone), but at the expense of their kinetics.
Figure 6. The kinetic
Figure transfer
6. The kinetic transferKiconstants,
constants, , and transfer resistance,
Ki, and Rtiresistance,
transfer , for all species
Rti, estimated from estimated
for all species the fitting of the
Ac-electrogravimetric data of
from the fitting of SWCNT/PPy composite measured
the Ac-electrogravimetric data of in 0.5 M NaCl:composite
SWCNT/PPy (a) kinetic measured in 0.5 MKi and
transfer constants,
(b) transferNaCl:
resistance, Rti . transfer constants, Ki and (b) transfer resistance, Rti.
(a) kinetic

Furthermore, the kinetic transfer rates, K , of free water molecules are somewhat
s
For the other species, the following +trends were observed: (i) the kinetic transfer rates,
close to the values of the Na .nH 2 O ions. As for the Cl− ions, their transfer+ occurs at low
Kc, of the H+ in the SWCNT/PPy are faster than the
frequencies and only at more anodic potentials.
kinetic transfer rates of H in the
SWCNT and (ii) the kinetic transfer rates, Ka, of the Cl in the SWCNT/PPy are more or
-

less equivalent to the kinetic transfer rates of the Cl- in the SWCNT [32] and in the pristine
doped PPy [46].
In view of the transfer resistance values, Rti, of the ions, the Na+.nH2O species from
the SWCNT/PPy are slightly more difficult to be transferred compared with the Na+.nH2O
J. Compos. Sci. 2021, 5, 25 10 of 14

In SWCNT/PPy composites, the kinetic transfer rates, Ki , of Na+ .nH2 O are signifi-
cantly lower than the Ki of the Na+ .nH2 O in pristine SWCNTs [32] and higher than the
Ki of the Na+ species detected in pristine PPy [45,49], measured under similar conditions.
These results indicate that the composite structure does not hinder the sodium species
transfer, and probably it is governed by the insertion process in PPy rather than a faster
electroadsorption/desorption process. Additionally, the contribution of the sodium species
to the charge compensation process (both electrostatic and faradaic) is enlarged to the
whole potential range, i.e., 0.4 V to −1.2 V vs. Ag/AgCl (larger than that used for SWCNTs
alone), but at the expense of their kinetics.
For the other species, the following trends were observed: (i) the kinetic transfer rates,
Kc , of the H+ in the SWCNT/PPy are faster than the kinetic transfer rates of H+ in the
SWCNT and (ii) the kinetic transfer rates, Ka , of the Cl− in the SWCNT/PPy are more or
less equivalent to the kinetic transfer rates of the Cl− in the SWCNT [32] and in the pristine
doped PPy [46].
In view of the transfer resistance values, Rti , of the ions, the Na+ .nH2 O species from
the SWCNT/PPy are slightly more difficult to be transferred compared with the Na+ .nH2 O
species in the SWCNT: 10 ohms·cm2 in Figure 6b versus 2 ohms·cm2 , previously given
in [32]. The Rti of H+ in the SWCNT/PPy composite is more or less equivalent to that
in pristine SWCNT. The Rti of Cl− in the SWCNT/PPy composite is comparable to that
obtained in pristine SWCNT [32] and that found in doped PPy electrodes [46].
Taking into consideration of the Ki and Rti values of various species detected, one can
suggest that there is no major hindrance of the ions transfer when doped PPy is deposited
on the SWCNT electrodes. The species detected are related to the electroactivity of the
two components. For example, at the anodic potentials, there is no cation contribution in
pristine SWCNTs, whereas it exits in the composite structure due to the PPy.
Figure 7 displays the relative concentration change for the ionic species detected in
the Ac-electrogravimetric measurements. The Ci -C0 values for H2 O are significantly higher
than the Ci -C0 values of Na+ .nH2 O, H+ and Cl− . The relative concentration change of the
Na+ .nH2 O in SWCNT/PPy is equivalent to the Ci -C0 values of the Na+ .nH2 O in pristine
SWCNT. These values agree well with the Ci -C0 values of Na+ detected in doped PPy
structures [46]. The same trend was observed for H+ and Cl− anions, Ci -C0 values of the
H+ and Cl− did not significantly vary in the SWCNT/PPy composite compared to pristine
i. 2021, 5, x SWCNTs [32]. The Ci -C0 values of the free solvent in the SWCNT–PPy 11 of 14 composite are
significantly higher than the Ci -C0 values of the free solvent in the SWCNT. These values
are rather similar to that obtained for doped PPy thin films studied in our group [46],
compared to thosewhich indicates
of pristine that the
SWCNTs, water
which transfer
could is principally
be revealed thanks governed by the doped PPy of the
to the advanced
electrogravimetric composite
tools (EQCM electrode.
and its EIS coupling) used in our study.

Figure
Figure 7. Evolution of 7. Evolution
the relative of the relative
concentration, Ci-C0 ofconcentration, Ci -Cthe
each species over 0 ofpotential
each species over the potential applied
applied
measured in 0.5 M NaCl SWCNT/PPy.
measured in 0.5 M NaCl SWCNT/PPy.

4. Conclusions
Globally, the signature of PPy in the SWCNT/PPy composite is not very visible, both
in current and mass response. Only after a closer examination, it can be seen that this
J. Compos. Sci. 2021, 5, 25 11 of 14

Solely evaluating the electrochemical signatures of the pristine SWCNT and SWCNT/PPy
composites does not allow us to distinguish the improvements in terms of interfacial species
transfer behavior. Therefore, the main novelty of the present study is the way of charac-
terization of the electrode/electrolyte interface by combining EQCM with simultaneous
EIS measurements, providing a gravimetric and dynamic deconvolution of the interfacial
processes. The results of the advanced electrogravimetric study revealed the differences
between the pristine and composite electrodes relative to their charge compensation mech-
anism. The clarified points include the enlargement of the usable potential window and the
enhanced permselectivity of the SWCNT/PPy composites compared to those of pristine
SWCNTs, which could be revealed thanks to the advanced electrogravimetric tools (EQCM
and its EIS coupling) used in our study.

4. Conclusions
Globally, the signature of PPy in the SWCNT/PPy composite is not very visible, both
in current and mass response. Only after a closer examination, it can be seen that this
composite SWCNT/PPy film has a much-improved response compared to pristine SWCNT
and PPy films tested under the same conditions. These subtleties can be summarized
as follows:
• The range of usable potential is greater for composite SWCNT/PPy films than for
SWCNT films, particularly for the cathodic part, since it is possible to push this
limit further (more cathodic potentials than −0.6 V vs. Ag/AgCl) before reaching
solvent degradation under the same experimental conditions. This should allow a
higher storage capacity to be obtained. This effect is also confirmed by mass variation
measurements via the responses given by the quartz microbalance tool;
• The mass change obtained with composite SWCNT/PPy films is much higher than
that observed for pristine SWCNT films if this mass variation is considered as a whole.
However, this observation should be moderated because the most cathodic potential
that can be achieved with these composite films is much lower compared to the tests
on SWNCT films. Nevertheless, if the same potential range is examined, classically
from −0.4 V vs. Ag/AgCl to +0.4 V vs. Ag/AgCl, the amount of (co)electroadsorbed
species is still four times greater;
• The permselectivity is also greatly improved in the case of composite SWCNT/PPy
films compared to SWCNT films. This is, on one hand, confirmed by the shape of the
∆m-E response, given by the EQCM measurements, which is monotonically decreasing
for composite SWCNT/PPy films while it has a classical “V” shape around the PZC for
SWCNT films. The part corresponding to the most anodic potentials is mainly related
to the electroadsorbed anions. In the case of the composite SWCNT/PPy films, this
contribution is clearly very small. On the other hand, the quantities of anions estimated
by Ac-electrogravimetric measurements are much lower in the case of composites.
This corroborates the better permselectivity of these composite SWCNT/PPy films
even with a moderate amount of PPy.

Author Contributions: F.E.-T. performed the experiments; H.P. and O.S. planned the experiments;
F.E.-T., H.P and O.S analyzed the data; H.P. and O.S. contributed to the original draft preparation
and revision; O.S. wrote, revised and edited the paper with discussions mainly with H.P. All authors
have read and agreed to the published version of the manuscript.
Funding: F.E.-T. thanks the (Secrétariat National pour l’Enseignement Supérieur, la Science, la
Technologie et l’Innovation (SENESCYT)-Equateur for the financial support during his Ph.D. thesis.
Acknowledgments: The authors thank Francoise Pillier for the FEG-SEM measurements.
Conflicts of Interest: The authors declare no conflict of interest.
J. Compos. Sci. 2021, 5, 25 12 of 14

Appendix A
∆E
The experimental data of the electrochemical impedance, (ω), and the charge/potential
∆I
∆q
TF, ∆E(ω ), were fitted using theoretical functions given in Equations (A1) and (A2). It is
considered here that only two ions, c1 and c2, are involved in the charge compensation process.
−1
∆E
 
Gc1 Gc2
(ω ) = jωdf + (A1)
∆I th jωdf + Kc1 jωdf + Kc2

∆q
 
Gc1 Gc2
( ω ) = df F + (A2)
∆E th jωdf + Kc1 jωdf + Kc2
where df is the film thickness, Ki represents the kinetics rate of transfer, whereas Gi describes
the level of difficulty corresponding to this transfer for each ionic species, c1 and c2 in our
case, transferred at the electrode/electrolyte interface. For the parameter Gi , an analogy
1
with a transfer resistance is generally used: Rti = FG i
.
The theoretical electrogravimetric transfer function, ∆m
∆E ( ω ), can be calculated, taking
into account the charged/uncharged species contribution:

∆m
 
Gc1 Gc2 Gs
( ω ) = d f Mc1 + Mc2 + M s (A3)
∆E th jωdf + Kc1 jωdf + Kc2 jωdf + Ks

In Equation (A3), additional parameters, Ks and Gs for the solvent molecules are
added as the molar mass (Mi ) of each species.
From the theoretical overall electrogravimetric transfer function (Equation (A3)),
it is possible to calculate the theoretical partial transfer functions by removing the c2
c1s c2s
contribution, calculating ∆m ( ) ; or the c1 contribution, calculating ∆m
(ω ), as

∆E ω ∆E
th th
shown in the following equations:

∆m c1s
 
Gc1 Gs
(ω ) = df ( Mc1 − Mc2 ) + Ms (A4)
∆E th jωdf + Kc1 jωdf + Ks

∆m c2s
 
Gc2 Gs
( ω ) = d f ( Mc2 − Mc1 ) + Ms (A5)
∆E th jωdf + Kc2 jωdf + Ks

References
1. Xue, Y.; Chen, S.; Yu, J.R.; Bunes, B.R.; Xue, Z.X.; Xu, J.K.; Lu, B.Y.; Zang, L. Nanostructured conducting polymers and their
composites: Synthesis methodologies, morphologies and applications. J. Mater. Chem. C 2020, 8, 10136–10159. [CrossRef]
2. Nair, S.S.; Mishra, S.K.; Kumar, D. Review—Polymeric materials for energy harvesting and storage applications. Polymer-Plastics
Technol. Mater. 2020. [CrossRef]
3. Simon, P.; Gogotsi, Y. Materials for electrochemical capacitors. Nat. Mater. 2008, 7, 845–854. [CrossRef] [PubMed]
4. Miller, J.R.; Simon, P. Electrochemical Capacitors for Energy Management. Science 2008, 321, 651–652. [CrossRef] [PubMed]
5. Wang, G.; Zhang, L.; Zhang, J. A review of electrode materials for electrochemical supercapacitors. Chem. Soc. Rev. 2012, 41,
797–828. [CrossRef] [PubMed]
6. Miller, J.R.; Burke, A. Electrochemical Capacitors: Challenges and Opportunities for Real-World Applications. Electrochem. Soc.
Interface 2008, 17, 53–57. [CrossRef]
7. Zhang, L.L.; Zhao, X.S. Carbon-based materials as supercapacitor electrodes. Chem. Soc. Rev. 2009, 38, 2520–2531. [CrossRef]
8. Béguin, F.; Presser, V.; Balducci, A.; Frackowiak, E. Carbons and Electrolytes for Advanced Supercapacitors. Adv. Mater. 2014, 26,
2219–2251. [CrossRef]
9. Shao, H.; Wu, Y.-C.; Lin, Z.; Taberna, P.-L.; Simon, P. Nanoporous carbon for electrochemical capacitive energy storage. Chem. Soc.
Rev. 2020, 49, 3005–3039. [CrossRef]
10. Han, L.; Tang, P.; Zhang, L. Hierarchical Co3O4@PPy@MnO2 core–shell–shell nanowire arrays for enhanced electrochemical
energy storage. Nano Energy 2014, 7, 42–51. [CrossRef]
11. Xia, X.; Chao, D.; Fan, Z.; Guan, C.; Cao, X.; Zhang, H.; Fan, H.J. A New Type of Porous Graphite Foams and Their Integrated Com-
posites with Oxide/Polymer Core/Shell Nanowires for Supercapacitors: Structural Design, Fabrication, and Full Supercapacitor
Demonstrations. Nano Lett. 2014, 14, 1651–1658. [CrossRef] [PubMed]
J. Compos. Sci. 2021, 5, 25 13 of 14

12. Simon, P.; Gogotsi, Y. Capacitive Energy Storage in Nanostructured Carbon–Electrolyte Systems. Acc. Chem. Res. 2013, 46,
1094–1103. [CrossRef] [PubMed]
13. Davies, A.; Yu, A. Material advancements in supercapacitors: From activated carbon to carbon nanotube and graphene. Can. J.
Chem. Eng. 2011, 89, 1342–1357. [CrossRef]
14. An, K.H.; Kim, W.S.; Park, Y.S.; Choi, Y.C.; Lee, S.M.; Chung, D.C.; Bae, D.J.; Lim, S.C.; Lee, Y.H. Supercapacitors Using
Single-Walled Carbon Nanotube Electrodes. Adv. Mater. 2001, 13, 497–500. [CrossRef]
15. Li, P.; Yang, Y.; Shi, E.; Shen, Q.; Shang, Y.; Wu, S.; Wei, J.; Wang, K.; Zhu, H.; Yuan, Q.; et al. Core-Double-Shell, Carbon
Nanotube@Polypyrrole@MnO2 Sponge as Freestanding, Compressible Supercapacitor Electrode. ACS Appl. Mater. Interfaces
2014, 6, 5228–5234. [CrossRef]
16. Zeng, S.; Chen, H.; Cai, F.; Kang, Y.; Chen, M.; Li, Q. Electrochemical fabrication of carbon nanotube/polyaniline hydrogel film
for all-solid-state flexible supercapacitor with high areal capacitance. J. Mater. Chem. A 2015, 3, 23864–23870. [CrossRef]
17. Mo, D.; Zhou, W.; Ma, X.; Xu, J.; Zhu, D.; Lu, B. Electrochemical synthesis and capacitance properties of a novel poly(3,4-
ethylenedioxythiophene bis-substituted bithiophene) electrode material. Electrochim. Acta 2014, 132, 67–74. [CrossRef]
18. Feng, H.; Wang, B.; Tan, L.; Chen, N.; Wang, N.; Chen, B. Polypyrrole/hexadecylpyridinium chloride-modified graphite oxide
composites: Fabrication, characterization, and application in supercapacitors. J. Power Sources 2014, 246, 621–628. [CrossRef]
19. Liu, T.; Finn, L.; Yu, M.; Wang, H.; Zhai, T.; Lu, X.; Tong, Y.; Li, Y. Polyaniline and Polypyrrole Pseudocapacitor Electrodes with
Excellent Cycling Stability. Nano Lett. 2014, 14, 2522–2527. [CrossRef]
20. Ambade, R.B.; Ambade, S.B.; Shrestha, N.K.; Nah, Y.-C.; Han, S.-H.; Lee, W.; Lee, S.-H. Polythiophene infiltrated TiO2 nanotubes
as high-performance supercapacitor electrodes. Chem. Commun. 2013, 49, 2308–2310. [CrossRef]
21. Shen, J.; Li, T.; Huang, W.; Long, Y.; Li, N.; Ye, M. One-pot polyelectrolyte assisted hydrothermal synthesis of RuO2 -reduced
graphene oxide nanocomposite. Electrochim. Acta 2013, 95, 155–161. [CrossRef]
22. Long, X.; Zeng, Z.; Guo, E.; Shi, X.; Zhou, H.; Wang, X. Facile fabrication of all-solid-state flexible interdigitated MnO2
supercapacitor via in-situ catalytic solution route. J. Power Sources 2016, 325, 264–272. [CrossRef]
23. Bae, J.; Song, M.K.; Park, Y.J.; Kim, J.M.; Liu, M.; Wang, Z.L. Fiber supercapacitors made of nanowire-fiber hybrid structures for
wearable/flexible energy storage. Angew. Chem. Int. Ed. Engl. 2011, 50, 1683–1687. [CrossRef]
24. Wang, D.-W.; Li, F.; Cheng, H.-M. Hierarchical porous nickel oxide and carbon as electrode materials for asymmetric supercapaci-
tor. J. Power Sources 2008, 185, 1563–1568. [CrossRef]
25. Meng, Q.; Cai, K.; Chen, Y.; Chen, L. Research progress on conducting polymer based supercapacitor electrode materials. Nano
Energy 2017, 36, 268–285. [CrossRef]
26. Shown, I.; Ganguly, A.; Chen, L.-C.; Chen, K.-H. Conducting polymer-based flexible supercapacitor. Energy Sci. Eng. 2015, 3,
2–26. [CrossRef]
27. Imani, A.; Farzi, G. Facile route for multi-walled carbon nanotube coating with polyaniline: Tubular morphology nanocomposites
for supercapacitor applications. J. Mater. Sci. Mater. Electron. 2015, 26, 7438–7444. [CrossRef]
28. Zhou, Y.; Xu, H.; Lachman, N.; Ghaffari, M.; Wu, S.; Liu, Y.; Ugur, A.; Gleason, K.K.; Wardle, B.L.; Zhang, Q.M. Advanced
asymmetric supercapacitor based on conducting polymer and aligned carbon nanotubes with controlled nanomorphology. Nano
Energy 2014, 9, 176–185. [CrossRef]
29. Tahir, M.; He, L.; Haider, W.A.; Yang, W.; Hong, X.; Guo, Y.; Pan, X.; Tang, H.; Li, Y.; Mai, L. Co-Electrodeposited porous
PEDOT–CNT microelectrodes for integrated micro-supercapacitors with high energy density, high rate capability, and long
cycling life. Nanoscale 2019, 11, 7761–7770. [CrossRef] [PubMed]
30. Chen, Y.; Du, L.; Yang, P.; Sun, P.; Yu, X.; Mai, W. Significantly enhanced robustness and electrochemical performance of flexible
carbon nanotube-based supercapacitors by electrodepositing polypyrrole. J. Power Sources 2015, 287, 68–74. [CrossRef]
31. Fu, C.; Zhou, H.; Liu, R.; Huang, Z.; Chen, J.; Kuang, Y. Supercapacitor based on electropolymerized polythiophene and
multi-walled carbon nanotubes composites. Mater. Chem. Phys. 2012, 132, 596–600. [CrossRef]
32. Escobar-Teran, F.; Arnau, A.; Garcia, J.V.; Jiménez, Y.; Perrot, H.; Sel, O. Gravimetric and dynamic deconvolution of global EQCM
response of carbon nanotube based electrodes by Ac-electrogravimetry. Electrochem. Commun. 2016, 70, 73–77. [CrossRef]
33. Escobar-Teran, F.; Perrot, H.; Sel, O. Charge storage properties of single wall carbon nanotubes/Prussian blue nanocube
composites studied by multi-scale coupled electrogravimetric methods. Electrochim. Acta 2018, 271, 297–304. [CrossRef]
34. Escobar-Teran, F.; Perrot, H.; Sel, O. Ion Dynamics at the Single Wall Carbon Nanotube Based Composite Electrode/Electrolyte
Interface: Influence of the Cation Size and Electrolyte pH. J. Phys. Chem. C 2019, 123, 4262–4273. [CrossRef]
35. Hillman, A.R. The EQCM: Electrogravimetry with a light touch. J. Solid State Electrochem. 2011, 15, 1647–1660. [CrossRef]
36. Tsai, W.-Y.; Taberna, P.-L.; Simon, P. Electrochemical Quartz Crystal Microbalance (EQCM) Study of Ion Dynamics in Nanoporous
Carbons. J. Am. Chem. Soc. 2014, 136, 8722–8728. [CrossRef]
37. Shpigel, N.; Levi, M.D.; Sigalov, S.; Daikhin, L.; Aurbach, D. In Situ Real-Time Mechanical and Morphological Characterization of
Electrodes for Electrochemical Energy Storage and Conversion by Electrochemical Quartz Crystal Microbalance with Dissipation
Monitoring. Acc. Chem. Res. 2018, 51, 69–79. [CrossRef]
38. Shpigel, N.; Levi, M.D.; Aurbach, D. EQCM-D technique for complex mechanical characterization of energy storage electrodes:
Background and practical guide. Energy Storage Mater. 2019, 21, 399–413. [CrossRef]
39. Levi, M.D.; Shpigel, N.; Sigalov, S.; Dargel, V.; Daikhin, L.; Aurbach, D. In Situ Porous Structure Characterization of Electrodes for
Energy Storage and Conversion by EQCM-D: A Review. Electrochim. Acta 2017, 232, 271–284. [CrossRef]
J. Compos. Sci. 2021, 5, 25 14 of 14

40. Gabrielli, C.; Garcia-Jareno, J.J.; Keddam, M.; Perrot, H.; Vicente, F. Ac-electrogravimetry study of electroactive thin films. II.
Application to polypyrrole. J. Phys. Chem. B 2002, 106, 3192–3201. [CrossRef]
41. Gabrielli, C.; García-Jareño, J.J.; Keddam, M.; Perrot, H.; Vicente, F. Ac-Electrogravimetry Study of Electroactive Thin Films. I.
Application to Prussian Blue. J. Phys. Chem. B 2002, 106, 3182–3191. [CrossRef]
42. Arias, C.R.; Debiemme-Chouvy, C.; Gabrielli, C.; Laberty-Robert, C.; Pailleret, A.; Perrot, H.; Sel, O. New Insights into Pseudoca-
pacitive Charge-Storage Mechanisms in Li-Birnessite Type MnO2 Monitored by Fast Quartz Crystal Microbalance Methods. J.
Phys. Chem. C 2014, 118, 26551–26559. [CrossRef]
43. Lemaire, P.; Dargon, T.; Alves Dalla Corte, D.; Sel, O.; Perrot, H.; Tarascon, J.-M. Making Advanced Electrogravimetry as an
Affordable Analytical Tool for Battery Interface Characterization. Anal. Chem. 2020, 92, 13803–13812. [CrossRef]
44. Lemaire, P.; Sel, O.; Alves Dalla Corte, D.; Iadecola, A.; Perrot, H.; Tarascon, J.-M. Elucidating the Origin of the Electro-
chemical Capacity in a Proton-Based Battery HxIrO4 via Advanced Electrogravimetry. ACS Appl. Mater. Interfaces 2020, 12,
4510–4519. [CrossRef]
45. Gao, W.L.; Sel, O.; Perrot, H. Electrochemical and viscoelastic evolution of dodecyl sulfate-doped polypyrrole films during
electrochemical cycling. Electrochim. Acta 2017, 233, 262–273. [CrossRef]
46. Benmouhoub, C.; Agrisuelas, J.; Benbrahim, N.; Pillier, F.; Gabrielli, C.; Kadri, A.; Pailleret, A.; Perrot, H.; Sel, O. Influence
of the Incorporation of CeO2 Nanoparticles on the Ion Exchange Behavior of Dodecylsulfate Doped Polypyrrole Films: Ac-
Electrogravimetry Investigations. Electrochim. Acta 2014, 145, 270–280. [CrossRef]
47. Levi, M.D.; Salitra, G.; Levy, N.; Aurbach, D.; Maier, J. Application of a quartz-crystal microbalance to measure ionic fluxes in
microporous carbons for energy storage. Nat. Mater. 2009, 8, 872–875. [CrossRef]
48. Sigalov, S.; Levi, M.D.; Salitra, G.; Aurbach, D.; Jänes, A.; Lust, E.; Halalay, I.C. Selective adsorption of multivalent ions into
TiC-derived nanoporous carbon. Carbon 2012, 50, 3957–3960. [CrossRef]
49. Gabrielli, C.; Garcia-Jareño, J.J.; Perrot, H. Charge compensation process in polypyrrole studied by ac electrogravimetry.
Electrochim. Acta 2001, 46, 4095–4103. [CrossRef]
50. Gabrielli, C.; Perrot, H.; Rose, D.; Rubin, A.; Toque, J.P.; Pham, M.C.; Piro, B. New frequency/voltage converters for ac-
electrogravimetric measurements based on fast quartz crystal microbalance. Rev. Sci. Instrum. 2007, 78. [CrossRef]
51. Pan, H.; Li, J.; Feng, Y. Carbon Nanotubes for Supercapacitor. Nanoscale Res. Lett. 2010, 5, 654. [CrossRef]
52. Pan, H.; Poh, C.K.; Feng, Y.P.; Lin, J. Supercapacitor Electrodes from Tubes-in-Tube Carbon Nanostructures. Chem. Mater. 2007, 19,
6120–6125. [CrossRef]

You might also like