Вы находитесь на странице: 1из 8

Electrochimica Acta 83 (2012) 463–470

Contents lists available at SciVerse ScienceDirect

Electrochimica Acta
journal homepage: www.elsevier.com/locate/electacta

Electrochemical incorporation of graphene oxide into conducting polymer films


Anna Österholm a,∗,1 , Tom Lindfors a,∗,1 , Jussi Kauppila b , Pia Damlin b , Carita Kvarnström b
a
Process Chemistry Centre, Laboratory of Analytical Chemistry, Åbo Akademi University, 20500 Turku, Finland
b
Turku University Centre for Materials and Surfaces (MATSURF), Laboratory of Materials Chemistry and Chemical Analysis, University of Turku, 20014 Turku, Finland

a r t i c l e i n f o a b s t r a c t

Article history: In this paper we report a simple, one-step method for direct electrochemical incorporation of graphene
Received 24 April 2012 oxide (GO) into conducting polymer films without the use of any additional dopants. Polypyrrole (PPy)
Received in revised form 8 July 2012 and poly(3,4-ethylenedioxythiophene) (PEDOT) were successfully electropolymerized in as-prepared
Accepted 9 July 2012
aqueous GO dispersions using potentiostatic polymerization. When dispersed in water, the GO flakes
Available online 14 August 2012
carry a negative surface charge which allows them to serve as counter ions and hence be incorporated into
the polymer films during electropolymerization. The incorporation of GO was verified using XPS whereas
Keywords:
vibrational spectroscopy was used to confirm the successful formation of the two conducting polymers.
Graphene oxide
Electrically conducting polymers
The surface morphologies of the polymer–GO films were investigated with SEM. The incorporation of
Composites GO into PPy resulted in a distinctly different and more porous surface morphology than that observed
Electrochemistry in PPy films synthesized in the presence of conventional counter ions. The PEDOT–GO film, on the other
Electropolymerization hand, had a very smooth surface morphology similar to that of PEDOT films polymerized in the presence of
other polyanions. As these composite films are aimed for electrochemical device applications, their redox
behavior and electrochemical capacitance properties were also investigated using cyclic voltammetry and
electrochemical impedance spectroscopy (EIS). The polymer–GO films exhibited good electrochemical
reversibility during cycling in 0.1 M KCl. The PPy films exhibited high redox capacitance regardless of
the counter ion used in electropolymerization whereas PEDOT–GO had lower redox capacitance than
PEDOT–PSS.
© 2012 Elsevier Ltd. All rights reserved.

1. Introduction ISEs, a high capacitance is a crucial material property. In theory,


the capacitance of ECP films can be increased just by increasing the
In the field of electrochemical device applications there is a amount of deposited polymer, but in practice, the film thickness
continuous need for new electrode materials. Electrically con- cannot be increased indefinitely since the ion transport is limited
ducting polymers (ECPs) are unique since they are both ionic in very thick films. Thus, one way to improve the capacitance
and electronic conductors. Some ECPs also exhibit high redox without the need for very thick films is to fabricate composite
capacitance, making them interesting for supercapacitor applica- materials of ECPs and different carbon materials.
tions and as ion-to-electron transducer materials in all-solid-state From the point of view of scalability and cost, graphene oxide
ion-selective electrodes (ISEs) [1,2]. To further improve material (GO) has become increasingly important as it forms stable dis-
properties like stability, conductivity, capacitance, and mechanical persions in water and can easily be synthesized from natural
strength, ECPs have been fabricated as composites with a wide graphite [16,17]. The good dispersibility is a consequence of its
variety of materials like metal nanoparticles [3], conventional highly oxidized structure with a large number of oxygen con-
polymers [4], and carbon materials like nanotubes and fullerenes taining functional groups (alkoxy, epoxy, carbonyl, and carboxyl
[5,6]. For the same reasons, the possibility to combine ECPs with groups) that results in the GO flakes having a negative surface
graphene and its derivatives has gained a lot of scientific ground charge [17–19]. Recently, there have been many reports on the
during the last few years, especially in the field of supercapacitors co-deposition of ECPs and GO. Most of them are focused on chem-
[7–15]. For applications like supercapacitors and all-solid-state ical polymerization of ECPs on GO sheets and papers [8,14,20] or
using chemical co-polymerization of GO with pyrrole, aniline, and
3,4-ethylenedioxythiophene (EDOT) [10,11,13,21].
Due to its anionic character, GO should be able to act as
∗ Corresponding authors. Tel.: +358 2 215 4419; fax: +358 2 215 4479.
dopant during electrochemical polymerization making it possible
E-mail addresses: anna.osterholm@chemistry.gatech.edu (A. Österholm),
tom.lindfors@abo.fi (T. Lindfors).
to deposit ECP–GO composite films directly on a variety of elec-
1
ISE member. trode substrates. The electrochemical co-polymerizations of ECPs

0013-4686/$ – see front matter © 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.electacta.2012.07.121
464 A. Österholm et al. / Electrochimica Acta 83 (2012) 463–470

with GO [13] and sulfonated graphene [12] have been reported, 2.2. Electrode preparation
but in both cases additional dopant ions, like polystyrene sulfonate
(PSS− ) or dodecyl benzene sulfonic acid (DBSA), were added to The electrochemical polymerizations were performed using an
the polymerization solutions. The drawback of polymerizing in Autolab PGSTAT20 equipped with GPES software and an impedance
presence of additional dopants is that a mixed material is formed module (FRA). Glassy carbon (GC) and tin-oxide (TO) were used
which makes it difficult to determine if GO is incorporated as a as working electrodes and a GC rod as a counter electrode. The
dopant or if it is merely physically entrapped in the polymer matrix. GC working electrodes were polished with 0.3 and 0.05 ␮m Al2 O3
Furthermore, by using additional dopants it becomes difficult to and the TO glasses were rinsed with deionized water and ethanol
control the amount of incorporated GO and hence to fabricate these and then cleaned in a plasma cleaner. All potentials reported are
films reproducibly. For example, it has been shown [13] that when given vs a Ag/AgCl (3 M KCl) reference electrode. Pyrrole (Aldrich)
polypyrrole (PPy) is electropolymerized in the presence of GO and was distilled before use whereas EDOT (Aldrich) was used as
PSS the polymerization parameters, like deposition charge and cur- received. Pyrrole (0.2 M) and EDOT (0.01 M) were added to the
rent density, had a great affect on the GO content and distribution above mentioned aqueous graphite oxide dispersions and the mix-
in the composite film. tures were ultrasonicated for 45 min to exfoliate the graphite oxide
In this paper, we report the one-step electrochemical deposition to GO. Reference films without GO were prepared using the same
of poly(3,4-ethylenedioxythiophene) (PEDOT)–GO and PPy–GO monomer concentration in 0.1 M KCl (PPy–Cl) and in 0.1 M NaPSS
composite films from aqueous GO dispersions without any addi- (PEDOT–PSS). All solutions were thoroughly purged with N2 gas.
tional dopants. In the absence of other dopants, GO flakes should The PPy and PEDOT films were deposited on the working elec-
be incorporated into the ECP films as negative counter ions. An trodes (active area: 7.1 mm2 ) using potentiostatic polymerization.
advantage of electrochemical polymerization is that the films are PPy–GO and PPy–Cl were polymerized at 0.88 V vs Ag/AgCl whereas
deposited directly on the electrode substrates. In comparison to PEDOT–GO and PEDOT–PSS were polymerized at 1.05 V and 0.92 V,
chemical oxidative polymerization, the film thickness can easily be respectively. These values were chosen from initial tests using gal-
controlled and the films are free from impurities (e.g. the oxidant vanostatic polymerization. Potentiostatic polymerization ensures
and its reaction products). PEDOT and PPy were chosen for this that similar film thicknesses are obtained and it also minimizes the
study mainly because their electropolymerization can be carried risk of unwanted side reactions that can occur during galvanostatic
out in aqueous electrolytes, even at neutral pH. The PEDOT–GO and polymerization. All polymerizations were continued until a total
PPy–GO films were characterized with cyclic voltammetry, elec- charge of 35.2 mC (0.5 C/cm2 ) had passed. It is assumed that all the
trochemical impedance spectroscopy (EIS), Raman, FTIR, energy charge is consumed by the film formation.
dispersive X-ray (EDX) and scanning electron spectroscopy (SEM).
X-ray photoelectron spectroscopy (XPS) was used to confirm that
2.3. Film characterization
GO was incorporated in the composite films during electropoly-
merization.
After polymerization the films were investigated by cyclic
voltammetry using the PGSTAT20 potentiostat. Five consecutive
2. Experimental cyclic voltammograms (CV) were recorded in oxygen free 0.1 M
KCl between −0.5 and 0.5 V using a 50 mV/s scan rate. Impedance
2.1. Preparation of graphite oxide
spectra in the frequency range 100 kHz to 10 mHz were measured
in the same electrolyte solution. Before recording the EIS spectra,
The preparation of graphite oxide was carried out with a modi-
the films were conditioned at 0.25 V for 60 s. All EIS spectra were
fied Hummers’ method [16]. In brief, 2 g of natural graphite flakes
recorded at 0.25 V with a 5 mV amplitude (E).
(Alfa Aesar, mesh 325) were dispersed in concentrated H2 SO4
A LEO 1530 Gemini FEG-SEM and a Thermo-Noran Vantage
(68 ml) and NaNO3 (1.5 g) in a three-necked flask equipped with a
X-ray Microanalysis System were used for the morphological char-
thermometer and a gas trap. KMnO4 (9 g) was then gradually added
acterization and the energy dispersive X-ray analyses. XPS spectra
to the mixture over a period of 30 min. During this step, the reac-
were collected using a Phi Quantum 2000 instrument (Physical
tion mixture was kept on an ice water bath due to the exothermic
Electronics) with monochromatized Al K␣ as the radiation source.
nature of the oxidation reaction. Following the addition of KMnO4 ,
For the FTIR and Raman measurements the polymer–GO and pure
the mixture was allowed to cool for 2 h in the ice water bath.
polymer films were deposited on TO glass. The ex situ Raman mea-
After cooling, the mixture was removed from the ice water bath
surements were performed using a Renishaw Ramascope (system
and was allowed to stand for five days. During this time, the reac-
1000 B, exc : 514 nm) equipped with a Leica microscope and a CCD
tion temperature was maintained below 35 ◦ C. Next, diluted H2 SO4
detector. The scattering signal was collected at 180◦ . FTIR spectra
was added, and the mixture was stirred overnight. Then, unreacted
were recorded from KBr pellets with a Bruker IFS 66/S spectrom-
KMnO4 was removed using 30% H2 O2 (6 ml). The final mixture
eter using a DTGS detector. For each spectrum 64 interferograms
was centrifuged and the graphite oxide precipitate collected and
were recorded. The resolution was 4 cm−1 .
re-dispersed in H2 O and centrifuged again. This procedure was
repeated several times to raise the pH. Finally, the dispersion was
placed in cellulose dialysis tubing (Thermo Scientific, MWCO 3500) 3. Results and discussion
and dialyzed against double quartz distilled H2 O. The outer solution
was changed twice daily until no significant UV–vis absorbance was 3.1. Electrochemical deposition of polymer-graphene oxide films
detected in outer solution. The final pH of the dispersion was 2.5 and
the concentration of graphite oxide was approximately 4.4 mg/ml. The electropolymerization of PPy–GO as well as PEDOT–GO
It has been shown [17] that GO dispersed in water has a negative resulted in homogeneous films with good substrate adhesion on GC
surface charge even at this low pH. Elemental analyses (XPS and and satisfactory adhesion on TO. PPy–GO could be electropolymer-
EDX) revealed that the as-prepared dispersion contains 2.4% sul- ized at the same potential as PPy–Cl whereas the polymerization of
fur, most likely originating from residual H2 SO4 . This could, in part, EDOT required a slightly higher potential in the GO dispersion than
contribute to the low pH value. H2 SO4 is fully dissociated and at pH in 0.1 M NaPSS (H2 O). As there were no additional dopants added
2.5 the concentration ratio [HSO4 − ]/[SO4 2− ] is approximately 0.2 to the GO dispersion, GO should be incorporated into the polymer
(pKa = 1.8 (HSO4 − )). The C:O ratio was determined to be 2.28. films during oxidative electropolymerization as the main counter
A. Österholm et al. / Electrochimica Acta 83 (2012) 463–470 465

ion. The necessity to use a slightly higher potential during elec-


tropolymerization of EDOT in the GO dispersion is probably a result
of the low pH. Polyanions, like PSS− and dodecyl sulfate, interact
electrostatically with the EDOT radicals during electropolymeriza-
tion. This facilitates the polymerization and lowers the potential
required for successful film formation [22,23]. The surface charge
of GO is highly pH dependent and at low pH the majority of car-
boxylic acid groups and all of the phenolic hydroxyl groups will
be protonated. As the surface of GO is chemically heterogeneous,
the surface functional groups have varying acidity and different pKa
values. This has made it challenging to determine the degree of pro-
tonation/surface charge on GO [24]. However, it is well known that
several acids like mellitic/graphitic acid and other benzene polycar-
boxylic acids have pKa values below and around 2 [25]. Increasing
the pH of the GO dispersion will increase the number of negative
groups due to a higher degree of deprotonation. This will increase
the anionic character of GO which, in turn, could have a positive
influence on the polymerization of EDOT in accordance with the
polymerization in presence of, e.g. PSS [22].
As mentioned, the GO dispersion used in this study contains
small amounts of sulfur, probably originating from residual H2 SO4
used in the GO synthesis. As the pH of the as-prepared dispersions
is 2.5, the majority of the sulfur should be in SO4 2− form; hence,
it is possible that SO4 2− and some HSO4 − could act as additional
counter ions during electropolymerization. However, from the EDX,
XPS, and vibrational analyses the amounts of sulfur were found to
be very small and, according to the evidence presented below, GO
should still act as the main counter ion.
The differences in elemental compositions of the films were
determined by EDX. The PPy–GO film consisted of carbon, oxygen,
as well as a small amount of sulfur. Thus, the elemental analysis
supported the presence of GO in the PPy–GO films as the spectra
showed a clear oxygen peak and a small sulfur peak along with
the expected peaks from carbon and nitrogen. As expected, the
EDX spectrum of the PPy–Cl film did not contain any detectable
amounts of oxygen or sulfur. Even though EDX has been used to Fig. 1. (a) Raman spectra of GO and PEDOT–GO using exc : 514 nm (inset:
study PEDOT–GO composites [21] it is not well suited for verify- PEDOT–PSS) and (b) FTIR spectra of GO, PEDOT–GO, and PEDOT–PSS.

ing the presence of GO since the elements used for the detection of
GO are also found in the PEDOT itself. It is therefore necessary to the PEDOT film [21], but also that the GO dispersion is a suitable
use alternative methods to verify the presence of GO in the PEDOT electropolymerization medium for PEDOT, as the PEDOT–GO and
composites. PEDOT–PSS have almost identical Raman spectra. As ECPs typically
show strong Raman spectra in the same wavenumber range as the
G and D bands of GO and related compounds, Raman spectroscopy
3.2. Vibrational analysis may not be the most suitable method for confirming the incorpora-
tion of GO in all ECP–GO composites since the spectral differences
Vibrational spectroscopy, especially Raman and FTIR, are often between ECP–GO films and pure ECP films are typically very subtle.
used to verify that GO, graphene, or graphite has been incorpo- In the Raman spectrum of PPy–GO it was not possible to distin-
rated into ECPs films [7,8,10–12,14,21]. Fig. 1a shows the Raman guish the GO bands (not shown). However, the PPy–Cl and PPy–GO
spectrum of PEDOT–GO, GO, and PEDOT–PSS (inset). The GO spec- spectra were almost identical which confirms that it possible to
trum consists of two bands centered at 1596 and 1353 cm−1 that electropolymerize PPy in the GO dispersion. It should also be men-
are attributed to the well-documented G and D modes [26,27]. tioned that SO4 2− and HSO4 − have several Raman bands between
The Raman spectrum of the PEDOT–GO film has two strong bands 1200 cm−1 and 400 cm−1 . Bands originating from SO4 2− and HSO4 −
at 1502 and 1436 cm−1 attributed to the asymmetric and sym- were not found in the Raman spectra of GO, PEDOT–GO, or PPy–GO.
metric C C stretch. The band at 1364 cm−1 originates from the This leads us to conclude that the amount of incorporated sulfur
C␤ C␤ stretch, the band at 1260 cm−1 from the C␣ C␣ inter-ring groups in PEDOT–GO and PPy–GO is very small.
stretch, the band at 1113 cm−1 from the C O C deformation, and Fig. 1b shows the FTIR transmission spectra of PEDOT–GO, GO,
finally the band at 988 cm−1 from the oxoethylene ring deforma- and PEDOT–PSS. The GO spectrum shows a broad band with a max-
tion [28]. The band pattern in the Raman spectrum of PEDOT–GO imum at 3424 cm−1 originating from structural OH groups and
is almost identical to that of PEDOT–PSS apart from small shifts adsorbed H2 O molecules. The narrow band at 1732 cm−1 origi-
in some band positions. In Fig. 1a the strong Raman bands orig- nating from the C O stretch is often assigned to carboxylic acid
inating from PEDOT almost overlap the G and D bands in GO. groups but according to a study by Szabó et al. [24,29] this band can
However, small shoulders are found on the high energy side of also originate from other carbonyl moieties like single ketones or
the 1563 cm−1 band and on the low energy side of the 1364 cm−1 quinones. In addition, IR bands from C OH deformation (shoul-
band in the PEDOT–GO spectrum. These shoulders are not found der from ca. 1395 cm−1 ), C O C (1220 cm−1 ), and C O groups
in the Raman spectrum of PEDOT–PSS (see inset in Fig. 1a). This (1051 cm−1 ) are found in the GO spectrum. The assignment of
suggests, not only that GO has been successfully incorporated in the IR band at 1620 cm−1 seems to be rather controversial and
466 A. Österholm et al. / Electrochimica Acta 83 (2012) 463–470

it is assigned to both skeletal vibrations of graphitic domains or proceed with caution when using these methods for determining
to bending vibrations in H2 O molecules [29–32]. For highly oxi- the presence of, e.g. GO in ECP films.
dized GO samples the latter assignment seems more probable.
The PEDOT–GO spectrum consists of a combination of IR bands 3.3. X-ray photoelectron spectroscopy
from PEDOT and from GO. IR bands at 1732 cm−1 , 1620 cm−1 , and
1220 cm−1 are found in the PEDOT–GO spectrum but are absent XPS is a surface technique that is able to provide information
in the PEDOT–PSS spectrum. The IR band at 1732 cm−1 assigned on the chemical and electronic structure of polymer–GO films.
to C O groups can also be found in FTIR spectra of over-oxidized The C 1s core level spectra of the PPy–GO and PEDOT–GO films
PEDOT (and other polymers), so this band alone should not be used and their corresponding reference films (PPy–Cl and PEDOT–PSS)
as evidence for GO incorporation as is often done. The main IR bands are shown in Fig. 2. The strongest peak in the PPy–Cl spectrum
from PEDOT are found below 1550 cm−1 . A very similar pattern of (Fig. 2a) is centered at 284.72 eV and originates from the C C/C C
IR bands are found in the PEDOT–GO spectrum as in PEDOT–PSS bonds. From the curve fit it was possible to distinguish the C N
spectrum, again confirming the successful formation of PEDOT in peak at 286.3 eV and small peaks at 288.1 and 289.3 eV from C O
the GO dispersion. The bands between 1512 and 1170 cm−1 are typ- and O C O groups, respectively [11,29,36]. The C O and O C O
ically assigned to the stretching of C C and C C in the thiophene peaks in the PPy–Cl spectrum suggest that some over-oxidation
ring whereas the IR bands between 1140 and 1050 are believed to must have occurred during the electropolymerization. The above
originate from the alkylenedioxy group. The bands at 920, 830, and mentioned peaks are also found in the XPS spectrum of the
685 cm−1 are vibrations from the C S bond. In addition to these PPy–GO film (Fig. 2b), but there is an additional strong peak
bands, both the PEDOT–PSS and the PEDOT–GO spectra show IR centered at 286.8 eV originating from C O (epoxy and hydroxyl)
bands at 1358, 1287, 1090, and 970 cm−1 ; these bands are believed groups [11,26,29,37]. It is well known that GO contains a large
to be so called doping induced bands [33–35]. The PEDOT–GO and number of epoxy groups and hydroxyl on the basal planes of
the PEDOT–PSS were made into KBr pellets immediately after syn- the GO sheets [17–19,24]. This is clear evidence that we have
thesis and so it is expected that they are in their oxidized and successfully incorporated GO into the PPy film. The XPS results
conducting forms. The spectra of PEDOT–PSS and PEDOT–GO also also indicate that the majority of the functional groups on our GO
show an IR band at 3441 and 3438 cm−1 , respectively; this band are epoxide and C OH groups [18,29]. These groups are polar due
could originate from OH groups in residual water. As we have to the C O bond and it is therefore assumed that they can function
shown, Raman and FTIR are useful methods for confirming the as charge compensating sites (electrophilic centers) during the
successful formation of ECPs in the GO dispersion but one should electropolymerization similar to charge compensating dopant ions

Fig. 2. C 1s XPS spectra of (a) PPy–Cl, (b) PPy–GO, (c) PEDOT–PSS, and (d) PEDOT–GO. The 2 values for the curve fittings were between 1.7 and 3.9.
A. Österholm et al. / Electrochimica Acta 83 (2012) 463–470 467

Fig. 3. SEM micrographs of (a) PPy–Cl and (b) PPy–GO; magnification 1000×. Fig. 4. SEM micrographs of (a) PEDOT–PSS and (b) PEDOT–GO; magnification
10,000×.

[18]. This is a reasonable assumption because water forms strong


hydrogen bonds with GO through the hydroxyl and epoxy groups film (Fig. 3a) has a nodular structure typical for PPy. The surface
[38]. Consequently, GO should also be able to act as big counter morphology of the PPy–GO (Fig. 3b) is flaky as opposed to nodular
anion with multiple charge compensating sites even if the degree and hence distinctly different from PPy–Cl. This surface morphol-
of deprotonation of carboxylic acid and phenolic groups would be ogy is also very different from the PPy–GO films reported by Deng
low in this GO dispersion. et al. [13]. The most probable reason is that we did not use any
The conclusions that can be drawn from the XPS spectra of the additional dopant ion during electropolymerization as was done in
PEDOT films (Fig. 2c and d) are in good agreement with those Ref. [13] and therefore the nodular PPy morphology is not as evi-
from PPy. In the PEDOT–PSS spectrum the peak at 286.7 eV orig- dent in our films. Our PPy–GO films were very uniform and had
inates from the C O C bond in the dioxyethylene ring. In the XPS formed evenly over the entire active area of the GC electrode. Parts
spectrum of PEDOT–GO the peak is centered at 286.9 eV and has of the film surface appear to be quite smooth but the film also has
significantly higher intensity as a result of the additional contri- rather big pores between the flakes. From the SEM image obtained
butions from the C O bonds in hydroxyl and epoxy groups of GO. using higher magnification (10,000×, not shown), it is possible to
From the S 2p3/2 XPS analyses (not shown) we could also conclude distinguish that the large flakes are covered with what appears to
that the residual sulfur was not covalently bound to the GO flakes or be PPy.
to the polymer structure as no S C peaks (at ca. 164 eV) were found The difference in surface morphology is not as pronounced
in the XPS spectrum of PPy. However, two low-intensity peaks that between the PEDOT films. Both the PEDOT–PSS film (Fig. 4a) and the
could be assigned to S O bonds (at ca. 168 eV) were found in both PEDOT–GO film (Fig. 4b) were electropolymerized in the presence
the PEDOT–GO and in the PPy–GO S 2p spectra. This supports our of large dopant ions, as opposed to PPy where one dopant ion was
conclusion that the residual sulfur exists as HSO4 − and SO4 2− . small (Cl− ) and the other large (GO). Electropolymerizing PEDOT
in presence of large dopant ions is known to result in a compact
3.4. Surface morphology surface morphology [38]. Interestingly, according to Fig. 4a and
b, the surface morphology of PEDOT–GO appears to be smoother
The surface morphology is an important parameter if the films than the surface morphology of the PEDOT–PSS film. From the SEM
are to be used as electrode materials in electrochemical devices. image recorded using the higher magnification (10,000×) a nodu-
The surface morphology of ECP films depends on the electrolyte, lar surface morphology is observed for the PEDOT–PSS film with
solvent, deposition method, monomer concentration, etc. Here all an average grain size of 100–200 nm. This nodular surface mor-
films were prepared in the same way and using the same set-up, phology is not seen in the PEDOT–GO film, not even using the
only the dopant ion was different. Fig. 3 shows the SEM images of highest magnification (50,000×, not shown). Fig. 4b also indicates
PPy films taken with the 1000× magnification which shows best the presence of larger flakes in the film which is in good agree-
the differences in the surface morphology of the films. The PPy–Cl ment with the PPy–GO film in Fig. 3b. The uniform and smooth
468 A. Österholm et al. / Electrochimica Acta 83 (2012) 463–470

reduction (Qred ) was 1.03 for PPy–Cl and 1.07 for PPy–GO (Fig. 5,
v = 50 mV/s). In the CV of PPy–GO there is a cathodic peak at −0.38 V
that decreases during redox cycling; this peak is not found in the
CV of PPy–Cl. This could be a consequence of morphological chain
reorganization in the PPy–GO film that occurs during the first redox
cycles as a result of ion exchange when the film is cycled in 0.1 M
KCl. This ion exchange would not occur in the PPy–Cl film as it
was prepared and therefore already in contact with 0.1 M KCl dur-
ing synthesis. A cathodic peak has also been observed for pure GO
at similar potentials and it has been ascribed to a redox pair of
some electrochemically active oxygen-containing groups on the GO
flakes [39].
The PEDOT films gave rise to a broad capacitive-like current
response (Fig. 5b). A broad oxidation peak is seen between 0 and
0.1 V but no clear reduction peak was observed during the reverse
scan. This is in good agreement with previously reported data for
PEDOT–PSS in aqueous electrolytes [22]. As for PPy, both PEDOT
films exhibited excellent reversibility as Qox /Qred was almost unity
for v = 50 mV/s (Fig. 5b). Compared to PEDOT–PSS, the redox current
and the charge values were significantly lower for PEDOT–GO. The
difference in electroactivity between PEDOT–PSS and PEDOT–GO in
Fig. 5b can be explained by several factors. First, the polymerization
of PEDOT–GO was carried out at a higher potential than PEDOT–PSS
which results in an increased risk of partial over-oxidation of PEDOT
and other side-reactions like cross-linking of polymer chains. Both
reactions would contribute to a lower redox current. The formation
of C O bonds is a typical sign of over-oxidation in PEDOT and other
conducting polymers [40]. According to the XPS spectra in Fig. 2,
PEDOT–PSS and PEDOT–GO both contain C O groups, but also GO
itself should contain some C O groups. It is therefore reasonable
Fig. 5. CVs recorded in 0.1 M KCl between −0.5 V and 0.5 V for (a) PPy–Cl and PPy–GO to assume that if PEDOT–GO would be highly over-oxidized, the
and (b) PEDOT–PSS and PEDOT–GO. The scan rate was 50 mV/s and the potentials peak intensity of the C O peak in the XPS spectrum of PEDOT–GO
are given vs Ag/AgCl/3 M KCl. (Fig. 2d) would be of significantly higher intensity than the cor-
responding peak in the XPS spectrum of PEDOT–PSS. As this does
surface morphology makes PEDOT–GO a promising material for not appear to be the case it seems reasonable to assume that the
further characterization and as ion-to-electron transducers in all- higher polymerization potential does not result in the formation of
solid-state ISEs. any significant amount of C O groups on the PEDOT chains. A sec-
ond possibility for the lower electroactivity of PEDOT–GO is that it
3.5. Electrochemical properties is a consequence of the different ions incorporated into the film
during polymerization and hence differences in chain morphol-
If these composite films are intended to be used in electro- ogy. The SEM images suggest that incorporation of GO results in
chemical applications it is necessary to investigate the effect of a smoother surface morphology. If this is true also in the bulk of
the incorporated GO on the electroactivity of the polymer. The the film, it could result in a lower effective surface area resulting in
polymer–GO and the reference films were characterized by cyclic a lower redox current. It has been reported that PEDOT–PSS forms
voltammetry in the potential range −0.5 V to 0.5 V in 0.1 M KCl. a highly swollen polymer hydrogel with a very high effective sur-
The ion transport occurring simultaneously during electrochem- face area that allows for fast ion diffusion within the film despite
ical oxidation and reduction of the polymer will depend on the its compact morphology [41]. The exact reason for the lower elec-
dopant ion incorporated into the film during electropolymeriza- troactivity is still unclear but an optimization of the polymerization
tion. PPy–Cl contains a small, and therefore mobile, counter ion parameters is currently underway to rule out the effect of possible
whereas PPy–GO contains a large, immobile counter ion. During cross-linking or other side-reactions. In both CVs in Fig. 5b there is
redox cycling of PPy–Cl, both anions and cations can contribute to a small decrease in redox current during the first few cycles; the
the ion transport. However, since no detectable amounts of potas- most likely and common explanation is a change in chain confor-
sium were found in the EDX spectra of neither reduced nor oxidized mation and film morphology as a result of an initial ion exchange
PPy–Cl, it seems like the ion transport under these experimental that occurs during the so called break-in period when the films are
conditions is dominated by Cl− . In PPy–GO we can assume that GO moved to 0.1 M KCl.
is immobilized in the PPy matrix as counter ions and, as a result, Fig. 6 shows the CVs of the PPy–GO and PEDOT–GO recorded at
cation ingress/egress is responsible for the ion transport. The CVs different scan rates. The PPy–GO and PEDOT–GO films were poly-
recorded during redox cycling of the PPy films are shown in Fig. 5a. merized on GC electrode with the diameter of 1.6 mm. The currents
The CVs look very similar indicating that neither the anions used are therefore lower than in Fig. 5 (d = 3 mm). The oxidation and
during the two polymerizations, nor the type of ion dominating the reduction peak separation of PPy–GO increased with increasing
charge transport, have a significant influence on the shape of the scan rate indicating slow electron transfer kinetics (Fig. 6a). It may
redox response of PPy. The integrated area of the CV, which is a be due to the surface morphology of PPy–GO and the incorporation
measure of the charge capacity (i.e. total amount of charge trans- of rather large electrically non-conducting GO flakes into the film,
ferred during one redox cycle), was slightly higher for PPy–GO than which not only can have a negative effect on the electrical con-
for PPy–Cl. Both films exhibited good electrochemical reversibil- ductivity of PPy but they can also sterically hinder the diffusion of
ity as the ratio of the charge passed during oxidation (Qox ) and charge compensating counter ions in the PPy matrix. The oxidation
A. Österholm et al. / Electrochimica Acta 83 (2012) 463–470 469

Fig. 7. EIS spectra of (a) PPy–Cl, (b) PPy–GO, (c) PEDOT–PSS, and (d) PEDOT–GO.

Fig. 6. CVs recorded in 0.1 M KCl for (a) PPy–GO and (b) PEDOT–GO with the scan
rates of 10, 20, 50, 100 and 200 mV/s. The potentials are given vs Ag/AgCl/3 M KCl.

peak potential of PEDOT–GO (ca. 0.045 V) is practically independent


of the scan rate showing the good reversibility of the oxidation and
reduction process (Fig. 6b). In comparison to PPy–GO, it is possible
that the electron transfer kinetics are faster in the more compact
Fig. 8. The absolute values of −Z vs f−1 plotted for the frequencies 10, 15.85 and
PEDOT–GO film.
100 mHz: (a) PEDOT–GO, (b) PEDOT–PSS, (c) PPy–GO and (d) PPy–Cl. PPy–GO and
PPy–Cl are overlapping.
3.6. Redox capacitance properties

To get some insight into the capacitive properties of these films of PEDOT–GO (r2 = 0.999). The calculated redox capacitances for
impedance spectra were recorded in 0.1 M KCl immediately fol- PPy–Cl and PPy–GO were both 2003 ␮F. For the PEDOT–PSS and
lowing the cyclic voltammetry analyses. Typical impedance spectra PEDOT–GO the capacitances were 1060 ␮F and 390 ␮F, respec-
for ECP–GO films at 0.25 V are shown in Fig. 7. For both PPy films tively. Since the EIS spectra of PEDOT–GO deviate to some extent
and the PEDOT–PSS film, the impedance plots are dominated by from a capacitive vertical line this approximation of the capac-
almost vertical capacitive lines that extend to low frequencies, itance contains some uncertainty. However, it is clear that this
which is characteristic for ideal capacitors. Only a slight devia- film has a much lower capacitance than PEDOT–PSS. Since the
tion from the capacitive line was observed at higher frequencies same deposition charge was used for PEDOT–PSS and PEDOT–GO,
indicating fast electronic and ionic charge transfer. For PEDOT–GO and by assuming a polymerization yield of 100%, we presume
there is a clear deviation from the ideal capacitive behavior at low that the amount of deposited polymer is similar. Thus, the lower
frequencies which can indicate a slow ion diffusion process or an capacitance suggests that some of the PEDOT in PEDOT–GO is not
increasing resistive behavior. If there is significant cross-linking or electroactive or that the PEDOT–GO film has a significantly lower
other side reactions occurring during polymerization it is possi- effective surface area.
ble that it results in a film morphology/chain packing that restricts 1
C= (1)
the ion movement in the bulk of the film. The redox capacitances 2f | − Z  |
(C) were determined according to Eq. (1) by line fitting (y = ax) of
the EIS data plotted as |−Z | vs f−1 and by calculating the capac- 4. Conclusions
itances from the slope (a) of the straight lines (C = (2a)−1 ). The
frequencies of 10, 15.85 and 100 mHz were used for the line fittings. In this paper we have shown that electrochemical polymeriza-
The line fits in Fig. 8 indicate that Eq. (1) gives a good approxima- tion is an easy, one-step route for synthesizing ECP–GO composite
tion of the redox capacitances. The correlation coefficients (r2 ) of films directly onto different electrode substrates in as-prepared
the line fits were higher than 0.9999 for all polymer films except GO dispersions without any additional dopant ions. Although there
470 A. Österholm et al. / Electrochimica Acta 83 (2012) 463–470

have been other reports on electrochemical one-step syntheses of [8] S. Biswas, T. Drzal, Chemistry of Materials 22 (2010) 5667.
ECP–GO films, additional dopants were typically added to the poly- [9] H. Wang, Q. Hao, X. Yang, L. Lu, X. Wang, ACS Applied Materials & Interfaces 2
(2010) 821.
merization solutions [13]. Even chemically polymerized ECP–GO [10] S. Konwer, S.K. Dolui, Materials Chemistry and Physics 124 (2010) 738.
films are often synthesized in the presence of dopant or surfactant [11] S. Bose, T. Kuila, M.E. Uddin, N.H. Kim, A.K.T. Lau, J.H. Lee, Polymer 51 (2010)
ions [7,12,24]. The drawback of using additional dopants is that it 5921.
[12] A. Liu, C. Li, H. Bai, G. Shi, Journal of Physical Chemistry C 114 (2010) 22783.
results in films that are mixed materials where the GO content and [13] M. Deng, X. Yang, M. Silke, M. Qiu, G. Borghs, H. Chen, Sensors and Actuators B
distribution can be difficult to control. 158 (2011) 176.
In this study, the incorporation of GO into PPy resulted in a [14] J. Zhang, S. Zhao, Journal of Physical Chemistry C 116 (2012) 5420.
[15] R. Bisseur, P.K.Y. Liu, S.F. Scully, Synthetic Metals 156 (2006) 1023.
distinctly different and more porous surface morphology without
[16] M. Hirata, T. Gotou, S. Horiuchi, M. Fujiwara, M. Ohba, Carbon 42 (2004)
compromising the polymer’s electroactivity or redox capacitance. 2929.
This is very important since the positive attributes of the ECP [17] D. Li, M.B. Müller, S. Glije, R.B. Kaner, G.G. Wallace, Nature Nanotechnology 3
(2008) 101.
should be maintained also in the composite material. The spec-
[18] A. Lerf, H. He, M. Forster, J. Klinowski, Journal of Physical Chemistry B 102 (1998)
troscopic evidence from several different techniques support our 4477.
claim that we have incorporated GO as a counter ion during elec- [19] D.R. Dreyer, S. Park, C.W. Bielawski, R. Ruoff, Chemical Society Reviews 39
trochemical polymerization, both in PPy and in PEDOT, and that (2010) 228.
[20] X. Yan, J. Chen, J. Yang, Q. Xue, P. Miele, ACS Applied Materials & Interfaces 2
the as-prepared GO dispersion can be used as electropolymeriza- (2010) 2521.
tion medium for these ECPs. Incorporation of GO into PEDOT films [21] S. Liu, J. Tian, L. Wang, Y. Luo, X. Sun, Analyst 136 (2011) 4898.
resulted in highly uniform and smooth films which is important [22] J. Bobacka, A. Lewenstam, A. Ivaska, Journal of Electroanalytical Chemistry 489
(2000) 17.
for preparing ion-to-electron transducer layers with reproducible [23] N. Sakmeche, J.J. Aaron, M. Fall, S. Aeiyach, M. Jouini, J.C. Lacroix, P.C. Lacaze,
response characteristics for all-solid-state ISEs. This, in combina- Langmuir 15 (1999) 2566.
tion with PEDOT’s well-known environmental stability, is a great [24] T. Szabó, E. Tombácz, E. Illés, I. Dékány, Carbon 44 (2006) 537.
[25] H.C. Brown, D.H. McDaniel, O. Häfliger, in: E.A. Baude, F.C. Nachod (Eds.), Deter-
advantage for PEDOT–GO which is the reason why further opti- mination of Organic Structures by Physical Methods, Academic Press, New York,
mization of the film deposition conditions is currently carried out 1955 (Chapter 14).
at our lab in order to improve the redox behavior and capacitance [26] K.N. Kundin, B. Ozbas, H.C. Schniepp, R.K. Prud’homme, I.A. Aksay, R. Car, Nano
Letters 8 (2008) 36.
properties.
[27] G.K. Ramesha, S. Sampath, Journal of Physical Chemistry C 113 (2009)
7985.
Acknowledgements [28] S. Garreau, G. Louarn, J.P. Buisson, G. Froyer, S. Lefrant, Macromolecules 32
(1999) 6807.
[29] T. Szabó, O. Berkesi, P. Forgó, K. Josepovits, Y. Sanakis, D. Petridis, I. Dékány,
Financial support from the Finnish National Graduate School Chemistry of Materials 18 (2006) 2740.
in Nanoscience (J.K.) and the Academy of Finland (T.L. and A.Ö.) [30] G. Socrates, Infrared and Raman Characteristic Group Frequencies – Tables and
Charts, 3rd ed., John Wiley & Sons, Chichester, 2001.
are gratefully acknowledged. Assoc. Prof. Róbert E. Gyurcsányi at
[31] M. Acik, G. Lee, C. Mattevi, M. Chhowalla, K. Cho, Y.J. Chabal, Nature Materials
the Budapest University of Technology and Economics is gratefully 9 (2010) 840.
acknowledged for fruitful discussions. [32] Y. Si, E.T. Samulski, Nano Letters 8 (2008) 1679.
[33] C. Kvarnström, H. Neugebauer, S. Blomquist, H.J. Ahonen, J. Kankare, A. Ivaska,
Electrochimica Acta 44 (1999) 2793.
References [34] C. Kvarnström, H. Neugebauer, A. Ivaska, N.S. Sariciftci, Journal of Molecular
Structure 521 (2000) 271.
[1] M. Rudge, J. Davey, I. Raistrick, S. Grottesfeld, J.P. Ferraris, Journal of Power [35] A. Dkhissi, F. Louwet, L. Groenendaal, D. Beljonne, R. Lazzaroni, J.L. Brédas,
Sources 47 (1994) 89. Chemical Physics Letters 359 (2002) 466.
[2] J. Bobacka, A. Ivaska, A. Lewenstam, Chemical Reviews 108 (2008) 329. [36] http://srdata.nist.gov/xps/
[3] T. Tsakova, Journal of Solid State Electrochemistry 12 (2008) 1421 (references [37] J. Zhang, H. Yang, G. Shen, P. Cheng, J. Zhang, S. Guo, Chemical Communications
therein). 46 (2010) 1112.
[4] C. Lu, Y. Tang, S. Huan, S. Li, L. Li, Y. Wang, Advanced Functional Materials 20 [38] A. Lerf, H. He, T. Riedl, M. Forster, J. Klinowski, Solid State Ionics 857 (1997)
(2010) 1714. 101.
[5] G.Z. Chen, M.S. Shaffer, D. Coleby, G. Dixon, W. Zhou, D.J. Fray, A.H. Windle, [39] L. Chen, Y. Tang, K. Wang, C. Liu, S. Luo, Electrochemistry Communications 13
Advanced Materials 12 (2000) 522. (2011) 133.
[6] N.S. Sarisciftci, L. Smilowitz, A.J. Heeger, F. Wudl, Science 258 (1992) 1474. [40] P. Tehrani, A. Kanciurewska, X. Crispin, N.C. Robinson, M. Fahlman, M. Berggren,
[7] F. Alvi, M.K. Ram, P.A. Basnayaka, E. Stefanakos, Y. Goswami, A. Kumar, Elec- Solid State Ionics 177 (2007) 3521.
trochimica Acta 56 (2011) 9406. [41] S. Ghosh, O. Inganäs, Advanced Materials 11 (1999) 1214.

Вам также может понравиться