You are on page 1of 10

Electrochimica Acta 52 (2007) 5294–5303

Synthesis of fully and partially sulfonated polyanilines


derived from ortanilic acid: An electrochemical and
electromicrogravimetric study
Abraham Guadalupe Cano Márquez, Luz Marı́a Torres Rodrı́guez ∗ , Antonio Montes Rojas
Laboratorio de Electroquı́mica, CIEP-Facultad de Ciencias Quı́micas, Universidad Autónoma de San Luis Potosı́,
Avenida Dr. Manuel Nava No. 6, Zona Universitaria, C.P 78210, San Luis Potosı́, S.L.P, Mexico
Received 28 November 2006; received in revised form 14 February 2007; accepted 14 February 2007
Available online 20 February 2007

Abstract
The electrochemical polymerization of 2-aminobenzene sulfonic acid, also called ortanilic acid (o-ASA), on a gold electrode precoated with
polyaniline (PANI), has been carried out. We proved that the electropolymerization of o-ASA is enhanced on PANI electrodes, resulting in thicker
films obtained in aqueous media at room temperature. The electrosynthesized film (P(o-ASA)) was characterized by cyclic voltammetry, FTIR and
nuclear magnetic resonance. The compensation of P(o-ASA) charge was evaluated using electrochemical quartz crystal microbalance combined
with cyclic voltammetry, which showed that the electroneutralization process mainly involves cations. Additionally, copolymers of aniline and
o-ASA were electrosynthesized, using a metallic electrode modified with PANI also as a working electrode. The degree of sulfanation of copolymers
has been modulated with the proportions of monomers in the electrosynthesis solution. The studies reveal a more important participation of cations
in fully sulfonated polyaniline than in partially sulfonated polyaniline.
© 2007 Elsevier Ltd. All rights reserved.

Keywords: Polyaniline; Sulfonated; EQCM study; Charge compensation process; 2-Aminobenzene sulfonic acid

1. Introduction [12]. In addition, some SPANIs are n-dopable [13,14]. Such


properties make them a suitable choice, better than PANI, for
Self-doped polyanilines represent one of the most interesting applications, such as biosensors [15] due to physiological pH
types of polyaniline (PANI) derivatives. These polymers bear values; or rechargeable batteries, since SPANIs are capable of
aniongenic functional groups, such as –COOH [1–4], –OH [5], storing more specific energy than PANI [16] as a function of
–SO3 H [1,6–16] directly or spacer bound to the aromatic ring. self-doping.
Among self-doped PANIs, special interest has been devoted to SPANIs have been synthesized by both chemical and elec-
sulfonated polyaniline (SPANI) because the sulfonic acid group trochemical methods. In the second case, the synthesis has been
is a strong acid, and because there is a wide variety of sul- performed mainly by copolymerization of aniline and sulfonated
fonated aromatic amines commercially available. Introducing aniline monomers, such as: orthanilic acid [17,18], metanilic
these groups in the PANI chain determines many distinctive acid [1,19–22], sulfanilic acid [18], aniline-2, 5-disulfonic acid
properties of SPANI, different from those of the parent polymer [23], and more recently, disulfonated aminonaphtyl derivatives
[6,7], such as: high solubility in aqueous media [8,9], elec- [24].
trochemical activity in alkaline and neutral media [1,10], the Homopolymerization of sulfonated aromatic amines has
property of self-doping [11], and the fact that its conductivity also been reported, either by casting the electrosynthesized
does not change when treated with aqueous solutions at pH ≥ 4 homopolymer from solution [25,26] or by electrodeposition
[13,14,18,27,28]. Nevertheless, in the latter case difficulties,
such as formation of thin films and low polymerization rates
∗ Corresponding author. Tel.: +52 48 26 24 40x561; fax: +52 48 26 23 71. were noted. Thiemann and Brett [18] observed that a particularly
E-mail address: luzmaria@uaslp.mx (L.M. Torres Rodrı́guez). important parameter at the initial stages of electrodeposition of

0013-4686/$ – see front matter © 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.electacta.2007.02.048
A.G. Cano Márquez et al. / Electrochimica Acta 52 (2007) 5294–5303 5295

poly(4-aminobenzene sulfonic acid) is the electrode material. It The sensitivity constant of the quartz crystal electrode was
was found that, whereas the monomer polymerizes on glassy car- determined by potentiostatic deposition of silver, and was found
bon, its electrodeposition on ITO electrodes is extremely slow. In to be 1.714 × 108 Hz g−1 cm2 . This data is comparable to the
this respect, it has been reported that the electropolymerization theoretical value of 1.81 × 108 Hz g−1 cm2 established by the
of substituted anilines is more effective on polyaniline modified Sauerbrey equation [34]. In order to verify that the film behaves
electrodes [29,30] than on bare surfaces. Thus, the electrosyn- as a rigid layer, the polymer oxidative charge and total mass
thesis of fully sulfonated polyanilines can be improved by using change (m) were evaluated from the polymer growth curves.
a precoated polyaniline electrode. The m–Q graph was linear showing that the equation of Sauer-
As a consequence of the difficulties of the electrodeposi- brey could be applied.
tion of fully sulfonated PANIs, the properties of these films
have not been studied enough, e.g. to our knowledge there are 2.3. Spectroscopy
no studies on the charge compensation process. Contrarily, for
partially sulfonated PANIs, this process has been evaluated in For IR measurements, the polymer samples were mixed and
different works [11,31–33]. These studies have revealed that pressed with KBr. The FTIR spectra were obtained by means
the electroneutralization of copolymers is mainly achieved by of a Termo Nicolet 470 FTIR. 1 H RMN and 13 C RMN spectra
cation participation, which suggests that in the case of fully were recorded using a Brüker DMX-500 Hz spectrometer having
sulfonated PANIs cation participation can be increased, due deuterated solvents as internal references.
to the presence of more sulfonic groups for electroneutraliza-
tion. 3. Results and discussions
Hence, the aim of this study is: (1) to establish a more advan-
tageous method of electrodeposition of fully sulfonated PANI 3.1. Electrosynthesis of a homopolymer
than those previously reported, based on electrochemical oxi-
dation of 2-aminobenzene sulfonic acid, also called ortanilic We tried to grow a homopolymer film, hereinafter referred to
acid (o-ASA), onto a gold substrate previously modified with as P(o-ASA), onto three different substrates: bare Au, aniline-
PANI, as well as on the electrochemical and spectroscopic char- aminobenzene sulfonic acid copolymer/Au, as well as PANI/Au.
acterization of the obtained film. (2) To determine the influence The electroxidation of o-ASA onto gold was performed either
of the film sulfonation degree on the charge compensation dur- by controlled potential electrolysis or potential sweeping. In
ing the redox process by studying electrochemical quartz crystal agreement with references [25,26], no characteristic growth was
microbalance (EQCM) measurements, the charge compensation observed. As a matter of fact, no film deposition was obtained.
process of the obtained homopolymer and partially sulfonated In the case of copolymer/Au substrate, the copolymer was
polyaniline (copolymers of aniline and o-ASA). electrosynthesized as reported by Barbero and Kötz [17] by
cycling the electrode potential between 200 and 900 mV at
2. Experimental 50 mV s−1 in a 0.05 M o-ASA and 1 × 10−3 M aniline aqueous
solution. These conditions were chosen because they proved to
2.1. Chemicals produce a film with a high sulfonation degree, resulting in a
substrate similar to the homopolymer. Deposition of P(o-ASA)
Aniline was distilled and kept in dark, all the other chemi- on a copolymer precoated electrode was achieved by repeated
cals were reagent grade and used without further purification. linear potential scanning, though repeated scans over the range
The aqueous solutions were prepared using deionized water −180 to 1020 mV at 50 mV s−1 in a 0.05 M o-ASA plus 0.5 M
(1.1 ␮−1 cm−1 ), and the solutions were deoxygenated by purg- H2 SO4 aqueous solution did not result in homogeneous film for-
ing with nitrogen gas. After this, a nitrogen atmosphere was kept mation. Finally, at the PANI/Au substrate a successful growth
over the solution during each run. of P(o-ASA) was achieved by repeated scans over the range
−200 mV to 1200 mV at 50 mV s−1 , as can be seen in Fig. 1. Two
2.2. EQCM measurement important facts can be emphasized. First, the current intensity
increases steadily at each scan, which is characteristic of poly-
Electrochemical measurements were conducted using a PAR mer growth during deposition. Second, during the first cycles
273A (Princeton Applied Research) potentiostat–galvanostat (Fig. 1A(a)), the voltammogram resembles that of PANI. Nev-
coupled to an electrochemical quartz crystal microbalance Seiko ertheless, at subsequent cycles (Fig. 1A(b)) the redox peaks get
model QCA922, both controlled by WinEchem V. 1.5 installed closer, until they reach a steady value (Fig. 1A(c)). It should
on a personal computer. be noticed that the small separation between peaks is a dis-
A typical three electrode cell was used consisting of a 9 MHz tinctive feature for self-doped PANIs [6,7]. The obtained film
AT-cut quartz crystal coated with gold (0.1963 cm2 ) as working was adherent enough to resist mechanical stress without peeling
electrode, platinum wire as counter electrode, and Ag/AgCl in off when washed with water, and was visible to the naked eye.
3 M NaCl as reference electrode. On the other hand, the quartz In order to have an idea of the amount of deposited film, the
crystal was mounted in a home-made cell, in which one side peak current density of potentiodynamic curves in monomer-
of the crystal was kept out of the solution to prevent capacitive free solution of P(o-ASA) was compared to that obtained by
shunting. Zhang et al. [13]. The values for the same film grown in dif-
5296 A.G. Cano Márquez et al. / Electrochimica Acta 52 (2007) 5294–5303

Fig. 1. (A) Cyclic voltammograms of poly(o-ASA) electrodeposition on a PANI modified Au electrode (0.1963 cm2 ): (a) 3 cycles, (b) 8 cycles, (c) 17 cycles. Working
solution: o-ASA (0.07 M) and H2 SO4 (0.5 M). Scan rate: 50 mV s−1 . (B) Change of mass during electrodeposition of P(o-ASA) recorded simultaneously with the
cyclic voltammograms shown in (A).

ferent synthesis conditions at the same scan rate, were around explanation is consistent with the fact that no noticeable film
6.1 mA cm−2 and 0.9 mA cm−2 , respectively. It should be men- formation was obtained on an aniline-aminobenzene sulfonic
tioned that, for the latter case, the film was obtained by applying acid copolymer/Au electrode, because the electroneutralization
a constant current for 6.5 h. It means that the electrodeposition in the oxidation process for this film is mainly caused by proton
method reported here has a much higher polymerization rate ejection. As a result, the group –SO3 H cannot be anchored to
of P(o-ASA) (19.5 mC cm−2 h−1 ). This value is also greater the copolymer film.
than that reported for aniline-o-aminobenzene sulfonic acid
copolymer (0.1 mC cm−2 h−1 ) [17]. However, during P(o-ASA) 3.1.1. Effect of substrate thickness on P(o-ASA)
polymerization, colored soluble oligomeric products were seen polymerization
near the electrode, which indicates an efficiency yield below In the course of this study, we observed the effect of thickness
100%. of the predeposited PANI substrate on the electropolymerization
Furthermore, the mass change was simultaneously recorded of P(o-ASA). Therefore, several potential sweeps ranging from
during electropolymerization. As shown in Fig. 1B, a uniform 1 to 10 cycles were performed for the electrodeposition on a
mass increment occurs at each cycle reaching a value of ca. PANI substrate.
6000 ng. EQCM measurements also show that the deposition Electrodeposition of P(o-ASA) on a PANI substrate, syn-
rate is lower for the first cycles. This is probably due to par- thesized by application of ten repetitive scans, resulted in a
tial covering of the electrode’s surface with P(o-ASA), which thick film (Fig. 1). On the other hand, during potentiodynamic
results in a cyclic voltammogram similar to the voltammogram electroxidation of o-ASA over a thin PANI film produced by
of PANI growth at this stage of polymerization. Afterwards, the performing only one scan of potential, the voltammograms
mass increase is higher, and the cyclic voltammetry response no resembled those obtained for sulfonated anilines on a bare metal-
longer resembles to the response of PANI, being characteristic lic electrode [25], characterized by an oxidation peak placed
of SPANIs. We think that this is due to the full coverage of the at about 1 V attributed to monomer oxidation, and a redox
electrode with the sulfonated polymer. At this point, deposition process around 0.6 V due to the product formed during the elec-
of P(o-ASA) on P(o-ASA) takes place. troxidation. The peak current of this process hardly increased
The enhanced electrodeposition of substituted PANI on a with successive potential cycling, although no polymer film was
precoated polyaniline electrode has been associated with the found on the electrode’s surface. These results show that the
formation of electrocatalytic sites at the PANI surface [29]. thickness of PANI substrate is decisive in the electrochemical
We propose that this could be related to the charge compen- deposition of P(o-ASA).
sation process of PANI. In fact, during electroxidation of PANI In order to find the optimum quantity of PANI substrate, we
positive charges are generated at the polymer chain, which are calculated the anodic charge of P(o-ASA) from the voltammetric
compensated by the incorporation of anions from the solution. responses obtained both during deposition (last cycle) and in
Consequently, the –SO3 group of o-ASA is incorporated into monomer-free 0.5 M H2 SO4 media. The obtained results are
the PANI film as a dopant anion, anchoring o-ASA monomers shown in Fig. 2 as a function of the number of cycles applied
on the PANI surface. Then, when the oxidation potential of for the electropolymerization of PANI.
o-ASA is attained, the o-ASA monomer is polymerized on It was observed that the charge obtained for the last cycle
PANI. During reduction, the anion cannot be ejected from PANI of polymerization is higher than the obtained for the film in
because P(o-ASA) is trapped within the PANI chains. This 0.5 M H2 SO4 . This result confirms that the polymerization yield
A.G. Cano Márquez et al. / Electrochimica Acta 52 (2007) 5294–5303 5297

are closer. The reduced separation of the two peaks for SPANI
has been associated with steric effects caused by the bulky sul-
fonic acid substituent [6,7]. This behavior contrasts with that
reported by Thiemann and Brett [18] for electropolymeriza-
tion of o-ASA on glassy carbon electrodes, which reports three
anodic current maxima at 0.3, 0.49 and 0.66 V versus SCE at the
cyclic voltammogram of P(o-ASA); furthermore, the shape of
the curves is different. This discrepancy could be attributed to
the different conditions of electrosynthesis, as well as to the elec-
trode material, different upper limit of oxidation, and monomer
concentration.
Finally, the response of P(o-ASA) was compared with the
cyclic voltammogram of PANI used for precoating the electrode.
As shown in Fig. 3, the curves are different. The current is higher
for P(o-ASA) and the second oxidation couple is shifted nega-
tively even more than the peak of PANI, whereas the first couple
Fig. 2. Integrated anodic charge of P(o-ASA): (䊉) during polymerization (last is shifted positively. These differences show that the film at the
cycle); (×) in H2 SO4 solution (0.5 M) as a function of cycle number applied electrode’s surface is mainly P(o-ASA).
during electrosynthesis of PANI substrate. Additionally, cyclic voltammetry was performed at different
scan rates. The peak heights scale linearly with the sweep rate
was less than 100%. The obtained curves also show that, when in the range 5–250 mV s−1 , but only in the case of the anodic
electroxidation takes place on thinner PANI films (one to four current of process I, the linear dependence between these vari-
cycles) no noticeable film deposition is obtained, whereas PANI ables remains no more than 25 mV s−1 . For this process, a plot
precoated electrodes synthesized from a minimum of five cycles of the peak current versus the square root of the scan rate was
generate a uniform P(o-ASA) film. The deposited charge of P(o- linear with zero different intercept, indicating that the current is
ASA) increases gradually with the amount of PANI substrate, controlled by diffusion. This deviation from the expected behav-
and attains a steady value. Therefore, in order to obtain P(o- ior of a surface-confined electroactive species can be originated
ASA) films a minimum of five cycles must be performed during by mass transport limitations within the film or charge transfer
PANI synthesis. limitations at the substrate film interface [35].
Moreover, the solubility of P(o-ASA) was evaluated and the
3.2. Electrochemical characterization of films results show that the film is highly soluble both in neutral and
alkaline media, barely soluble in acidic media and completely
After electrosynthesis, the film was analyzed by cyclic insoluble in organic media, such as acetonitrile and acetone. This
voltammetry in 0.5 M H2 SO4 P(o-ASA). The resulting voltam- behavior is in agreement with a high molecular weight polymer
mogram was stable and the film could be cycled repeatedly linked to sulfonic acid groups.
without considerable change in the electrochemical response.
As can be seen in Fig. 3a, the cyclic voltammogram has two 3.2.1. Redox processes at P(o-ASA) modified electrodes
overlapping oxidation peaks (I, II) as well as their cathodic The electrochemical behavior of the hydroquinone–
counterparts at around 425 and 595 mV. This response is simi- benzoquinone (HQ/BQ) redox couple was tested using a P(o-
lar to the parent PANI, except that the two sets of redox peaks ASA) film modified electrode as working electrode. Cyclic
voltammetric response obtained from scans between −100
and 950 mV at 50 mV s−1 , shows two redox processes, one
attributed to the process of P(o-ASA), and the other asso-
ciated with the quasireversible redox transition between HQ
and BQ. For the HQ/BQ redox process, both the observed
E as well as the current for the P(o-ASA) modified elec-
trode were higher than those obtained on bare gold. These
evidences show that the quinone–hydroquinone couple acts
more irreversibly for the P(o-ASA) modified electrode, which
indicates that P(o-ASA) inhibits the electrochemical reaction.
This behavior is contrary to that reported for PANI [36], and
copolymer aniline/m-aminobenzene sulfonic acid (metanilic
acid) [22] coated electrodes, which show an electrocatalytic
ability toward this redox couple. The opposite behavior of the
Fig. 3. Cyclic voltammetry of (a) electrochemically deposited P(o-ASA) on
quinone/hydroquinone couple on P(o-ASA) and PANI indicates
PANI/Au (0.1963 cm2 ) and (b) PANI/Au (0.1963 cm2 ). Working solution: that the redox reaction proceeds mainly onto the surface of
H2 SO4 (0.5 M). Scan rate: 50 mV s−1 . P(o-ASA). Otherwise, the redox reaction would be carried out
5298 A.G. Cano Márquez et al. / Electrochimica Acta 52 (2007) 5294–5303

principally on the PANI’s surface and electrocatalytic behavior


would be obtained.

3.2.2. Effect of pH value on the cyclic voltammograms of


P(o-ASA)
To investigate the influence of the pH on the redox behavior
of the P(o-ASA), cyclic voltammograms of electrochemically
synthesized films prepared under the same conditions as above,
were recorded in solutions of different pH values. However, the
P(o-ASA) is soluble in solutions of neutral and alkaline pH; in
consequence, it was impossible to obtain the potentiodynamic
curves of the electrodeposited P(o-ASA). Therefore, in order to
determinate the electrochemical behavior of P(o-ASA), it was
necessary to measure the voltammograms for the polymer in
solution and not as a film. The obtained results are shown in Fig. 5. P(o-ASA) FTIR spectra.
Fig. 4, the cyclic voltammograms are almost equal for both
pH and display a reversible wave at E◦ = 291 mV correspond- are similar to those reported by Li et al. [10] for other sulfonated
ing to electrochemical activity of P(o-ASA). The resemblance polyaniline.
between the curves obtained at pH 4 and 7 demonstrates that
the electroactivity of P(o-ASA) is independent from pH in the 3.3. Spectroscopic analysis of P(o-ASA)
studied interval. For the case of pH 7, the upper scan limit
was extended as a result of an irreversible anodic peak situ- 3.3.1. Fourier transformed infrared spectroscopy
ated in 625 mV associated with film degradation. Additionally, The FTIR spectra of the film measured using KBr pellets
it should be mentioned that the obtained cyclic voltammograms can be seen in Fig. 5. The obtained spectrum is similar to that
reported by Zhang et al. [13] for the P(o-ASA) synthesized in a
mixed solvent of acetonitrile and water. The FTIR spectra give
similar bands to those of a PANI films as well as the broad band
centered at 3437 cm−1 attributed to unresolved NH stretching
modes, the peaks at 1632 and 1483 cm−1 due to the stretching
of aromatic carbons and the C–N stretching band that appears at
1200 cm−1 . In addition to the characteristic bands of the parent
PANI, the spectrum showed distinctive peaks of the sulfonic
group as well as very intense bands at 1060 and 1022 cm−1
assigned to the asymmetric and symmetric O S O stretching
vibrations, whereas another band, seen at 664 cm−1 , is attributed
to the C–S stretching. In summary, the spectra is consistent with
the presence of SO3 − groups attached to the aromatic rings.

3.3.2. Nuclear magnetic resonance spectroscopy


Once the presence of the characteristic peaks due to the sul-
fonic acid group was confirmed, the NMR H+ and C13 spectra
were obtained with the aim of determining the type of coupling
during polymerization.
In the NMR H+ case (Table 1), there is a signal at 2.5 ppm
that belongs to DMSO-d6 , whereas at 6.2 ppm the signal is due
to H2 O and/or to –NH group; the three bands found between 7.3
and 8 ppm belong to aromatic hydrogen atoms. Their coupling
constants (J) are shown in Table 1.
The J values show that H(6) is ortho-coupled with H(5),
which generates a doublet centered at 7.362 ppm. H(5) is also
ortho-coupled with H(6) which produces a doublet. Then, it
couples at meta with H(3). As a result, the signal, centered at
7.73 ppm, doubles again. Finally, H(3) is meta-coupled only with
H(5), which causes a doublet centered at 7.955. According to the
Fig. 4. Cyclic voltammograms of P(o-ASA) (100 mg L−1 ) on Au electrode
(0.1963 cm2 ) in (NH4 )2 SO4 (0.5 M) aqueous solution at: (a) 400, (b) 300, (c) observed signals, the only available carbon atom for polymer-
200, (d) 150, (e) 100, (f) 75, (g) 50, (h) 25, (i) 15, (j) 10, (k) 5 mV s−1 . Solution ization is C(4), which implies that it takes place as a head-to-tail
pH: (A) pH 4, (B) pH 7. mechanism. On the other hand, on comparing these spectra with
A.G. Cano Márquez et al. / Electrochimica Acta 52 (2007) 5294–5303 5299

Table 1
Aromatic hydrogens of P(o-ASA)
Atom J (Hz) Coupling Signal Location (ppm)

H(3) 2.0 meta Doublet 7.95


H(5) 8.25; 2.5, 2.0 ortho; meta, meta Double doublet 7.73
H(6) 8.0 ortho Doublet 7.95

Coupling constants.

those of the parent PANI [19], peaks belonging to H(3), H(5) and was measured for acids with anions of different molar mass
H(6) are seen to shift toward lower wavelengths. This fact is in (Table 3). The cyclic voltammetry response was very similar for
agreement with the presence of the –SO3 H group. Furthermore, all the acids, while the mass decrease obtained for the first peak
it should be pointed out that this group is found as –SO3 − , due oxidation was independent of the molar mass of the anion. This
to the fact that there is no signal corresponding to the –SO3 H result shows that charge compensation is achieved mainly by
group (around 12 ppm). proton ejection. At the same time, the increase in mass associated
In addition, the C13 spectra showed signals (Table 2) located with the second oxidation process is different for each anion size,
between 123 and 130 ppm, which correspond to quaternary car- but no lineal relation between the size of anion and the change
bon atoms, while those found between 136 and 140 ppm belong of mass is observed. This indicates that for the oxidation in this
to methynic or aromatic atom carbons. It should be mentioned region of potential, the mass increase is not only originated by
that each signal was assigned with the aid of tables. the incorporation of anion, which would be reflected in a linear
relationship between the anion size and the mass increase, but
3.4. Charge compensation dynamics in monomer-free also results from the cation departure.
solutions To obtain more quantitative information for the first process,
mass–charge curves were plotted and molar mass equivalents
It is a well-known fact that charge compensation is attained of the exchanged species were calculated from their slopes.
mainly by proton participation in sulfonated polyanilines [11,17, As shown in Fig. 6B, two well-delineated regions are present,
31–33]. In order to confirm this behavior, the ion exchange prop- the first one corresponding to the first oxidation process of the
erties P(o-ASA) were investigated by EQCM. Fig. 6A shows the film and characterized by a negative slope indicating that the
cyclic voltammogram and the corresponding mass change of a charge compensation is mainly effected by ejection of species;
P(o-ASA) film in 0.5 M H2 SO4 . the second region, however, presents a positive slope, which
It can be seen that at the beginning of the scanning no redox means that electroneutralization is made principally by anion
processes occur, and both the current and the mass remain almost insertion.
constant. When the potential becomes more oxidative, the first Taking into account that the charge compensation during
oxidation process takes place accompanied by a mass decrease; the first oxidation process of P(o-ASA) is held exclusively by
this behavior is maintained to attain the second oxidation peak ejection of cations, the molar mass equivalent (MME) of the
(held at 585 mV). These evidences clearly show that electroneu- exchanged species would be −19 g mol−1 , corresponding to
tralization in this potential range is mainly caused by proton a hydronium ion (H3 O+ ). However, as shown in Table 3, the
expulsion, in good agreement with the cationic transport model calculated MMEs differ from this value, thus suggesting the
and in opposition to the behavior of conducting polymers con- participation of anion and/or dissolvent; as the MMEs are very
taining mobile counter-anions. On the other hand, at the end similar (taking into account the experimental error) for all acids,
of the second oxidation process, a mass increment occurs as a the charge compensation is thereby independent of the anion
consequence of anion uptake. It should be mentioned that the size. Consequently, significant participation of anions in the
mass change for full oxidation is different from that reported electroneutralization is discarded. These observations suggest
for the parent PANI [37], where an important mass reduction is that the electroneutralization of P(o-ASA) is achieved mainly
observed. by proton expulsion and probably solvent incorporation. It is
In order to evaluate the extent of anion participation, the mass important to mention that the MME for the first process was
change upon oxidation of P(o-ASA) in 0.5 M acidic solutions estimated also for eight different sweep rates ranging from 5 to
200 mV s−1 in H2 SO4 (0.5 M) and 0.5 M camphorsulfonic acid
Table 2
Carbon atom location obtained for P(o-ASA) Table 3
MME calculated for the first redox process of P(o-ASA)
Atom Type Location (ppm)
Molar mass of anion (g mol−1 ) MME (g eq−1 )
C(l) Methynic 139.52
C(2) Methynic 136.36 HCl 35.5 −2.9
C(3) Quaternary 123.69 HNO3 62.0 −2.8
C(4) Methynic 129.98 H2 SO4 97.1 −6.3
C(5) Quaternary 128.22 HClO4 99.4 −5.4
C(6) Quaternary 125.16 HCS 231.3 −4.0
5300 A.G. Cano Márquez et al. / Electrochimica Acta 52 (2007) 5294–5303

Fig. 6. (A) (a) Potentiodynamic and (b) electromicrogravimetric responses for a P(o-ASA) film deposited on PANI/Au (0.1963 cm2 ) electrode. Working solution:
H2 SO4 (0.5 M). Scan rate: 50 mV s−1 . (B) Mass change–charge plots for P(o-ASA). Data calculated from (A).

(HCS). The obtained values were practically independent of the differences in cation size. Since P(o-ASA) is soluble in neutral
sweep rate and anion size of the acid used. aqueous solution, it was necessary to work in organic media, i.e.
In order to analyze the influence of cation size on the in acetonitrile, in which it is insoluble. The obtained results are
charge compensation mechanism, we have studied current and shown in Fig. 7; in both cases, the potentiodynamic response is
mass responses of P(o-ASA) simultaneously measured during characterized by a considerable decrement in the film electroac-
redox reactions in 0.5 M lithium tetrafluoroborate (TFBLi) and tivity in comparison with that obtained in the aqueous media. In
0.5 M tetraethylammonium tetrafluoroborate (TFBTEA) in ace- fact, during the first scans the cyclic voltammograms of P(o-
tonitrile. These electrolytes were chosen because of the huge ASA) are characterized by a broad anodic peak centered at

Fig. 7. (A) Cyclic voltammetry and EQCM profiles on cycling at 50 mV s−1 of a P(o-ASA) films in: (a) 0.5 M TFBLi and (b) 0.5 M TFBTEA in acetonitrile. (B)
Plots of m/Q for P(o-ASA) films in acetonitrile solutions of (C) TFBLi (0.1 M) and (D) TFBTEA (0.1 M). Data calculated from (A and B).
A.G. Cano Márquez et al. / Electrochimica Acta 52 (2007) 5294–5303 5301

0.670 V and a cathodic wave at −0.060 V. However, at succes- minimum of 30% sulfonation degree has enough dopant ions
sive scans, when the mass response is stabilized, the anodic peak bound for electroneutralization in the first oxidation process.
vanishes as is shown in Fig. 7. It should be mentioned that after This means that the level of cation participation in the charge
the scans were performed in acetonitrile, the films were analyzed compensation mechanism would be equal for both a copoly-
using cyclic voltammetry (CV) in 0.5 M H2 S04 aqueous solu- mer (30% degree of sulfonation or more) and a homopolymer,
tion. The polymer electroactivity was very similar to that shown for the first process. In order to determine whether the cation
in Fig. 6A. Therefore, a different cyclic voltammetry behav- participation in the charge compensation process is limited by
ior shown in acetonitrile is not a consequence of destruction or the quantity of sulfonic acid groups linked to PANI chain or
solubilization of the film. not, the current and mass responses to a potential scan were
On the other hand, the mass change profile is different for the simultaneously measured in 0.5 M H2 SO4 for different copoly-
two cations (Fig. 7). In the case of TFBLi, the obtained curve mers. The electrosynthesis of copolymers was performed using
is complex, showing an augmentation of mass at the beginning different ratios of aniline and o-ASA while keeping the con-
of the scan where negative current (reduction) is present; in the centration constant (0.07 M). The growth of different polymers
region where maximum current is attained a decrease in mass is was obtained by repeated cyclic voltammogram scans between
observed even at the end of the scan. This behavior suggests that −180 and 1050 mV on PANI/Au electrode.
the electroneutralization throughout the entire region of film oxi- To determine whether the sulfonation degree of copolymers
dation is mainly achieved by cation expulsion. When the scan is depends on the composition of electrosynthesis solution, FTIR
reversed, the current is still positive (oxidation), and the decrease spectroscopy was carried out for the films obtained with the
of mass continues, but when the current turns negative (reduc- following proportions of o-ASA:aniline, 50:50 and 30:70. The
tion), a mass increase is observed. As there is an increase in spectra show that the relation between the intensity of absorption
mass during reduction and a decrease in mass during oxida- peak at 1061 and 1200 cm−1 bands, characteristic of sulfonic
tion, the electroneutralization during film oxidation is achieved group, and the bands centered at 1638 and 1480 cm−1 due to
principally through the cation ejection. the stretching of the benzenoid and quinoid structures of PANI,
In the TFBEA case, the curve mass/potential has shown a is greater for the film obtained with the solution containing the
significant mass decrease in almost the entire region of potential highest quantity of o-ASA. This result shows that the number
during film oxidation, whereas a small mass increase is observed of sulfonic groups in the film can be modulated by varying the
only at the end of the scan. This indicates that charge compen- composition of working solution. It should be mentioned that
sation is accomplished principally by expulsion of cation during the intensity of bands at copolymer FTIR spectra indicates the
the oxidation process. It is important to note that the mass change considerable sulfonation degree. This result is relevant in the
in the potential region of the second redox process in acetonitrile, sense that the copolymers reported, obtained on a bare metallic
of both TFBEA and TFBLi, is different from those occurring in electrode, require a 1:10 proportion of aniline and sulfonated
aqueous media, where an important augmentation of mass is monomer to obtain a sulfonation degree of the order of 25% [32].
observed. After the electrochemical synthesis, the films were removed
To clarify the nature of the exchanged species during the from the preparation medium, washed with H2 SO4 (0.5 M) and
oxidation, mass-charge curves were plotted and they are shown studied by EQCM with both an aqueous solution containing
in Fig. 7C and D. Based on these curves, the equivalent mass 0.5 M H2 SO4 and HCS (0.5 M) solution as supporting elec-
exchanges from the beginning of the first oxidation process were trolyte. Cyclic voltammograms and mass change profiles versus
calculated; the values obtained were −2.3 and −111 g mol−1 potential obtained in H2 SO4 are similar to those obtained in
for TFBLi and TFBTEA, respectively. In the case of TFBLi, HCS. The results obtained for H2 SO4 are shown in Fig. 8. The
the value is bigger that the molar mass of cation (−7 g mol−1 ). potentiodynamic response shows that an increase in proportion
This result shows that there is a contribution of anions and/or of o-ASA in the synthesis solution results in a reduction of the
solvent to the electroneutralization process. Meanwhile in the separation between the peaks of the film. This suggests that the
TFBTEA case, the value obtained is rather close to the mass percentage of sulfonated aniline in the polymer increases with
of tetraethylammonium cation (116 gmol−1 ), which indicates the quantity of sulfonated monomer in the synthesis solution. In
that the charge compensation for the first oxidation process is fact, the small separation between the observed peaks when sul-
achieved exclusively through the cation ejection. These results fonic group is incorporated into the film, means that the greater
are in accord with Mello et al. [32], who demonstrated that anion is the amount of –SO3 − groups in the polymer, the smaller is
participation is strongly dependent on cation size. the the gap between the peaks.
Concerning the curves of mass change versus potential, two
3.5. Comparisons of charge compensation process in aspects are of interest: the mass variation observed during the
aqueous media of partially and fully sulfonated polyanilines forward sweep in the potential region between 600 and 900 mV
and the mass change of the first process. In the first case, the mass
It is established that the dopant level (the term commonly is seen to change with the composition of the solution employed
used to denote the number of charge balancing counter ions per in the electrosynthesis. In the case of PANI film and the film
unity of monomer) for substituted PANI ranges from 0.29 to obtained with the 10:90 electrosynthesis solution, a consider-
0.33 for the first oxidation process, and from 0.33 to 0.34 for the able mass decrease occurs in this potential region, associated
second process, respectively [38]. Therefore, PANI film with a with the deprotonization of the film [37]. On the other hand, in
5302 A.G. Cano Márquez et al. / Electrochimica Acta 52 (2007) 5294–5303

Fig. 8. (a) Cyclic voltammetric response and (b) mass change vs. potential curves recorded in H2 SO4 (0.5 M) during potential cycles at 50 mV s−1 , for polymers
obtained from H2 SO4 (0.5 M) solutions and different o-ASA:aniline proportions while keeping the concentration constant (0.07 M), (A) 0:100, (B)10:90, (C) 30:70,
(D) 50:50; (E) 70:30 and (F) 100:00. Mass data were filtered by Microal Origin 6.0 using the 5-point adjacent average method.

the case of films obtained with working solutions at higher ratios tion process: for the majority of copolymers an mass increase
of o-ASA, the mass remains practically constant in this region of is observed, which means anion participation in the electroneu-
potential. Finally, in the case of P(o-ASA), a slight mass increase tralization of the film. Only copolymers with a greater degree of
is seen. This result shows that the charge compensation process sulfonation present a small decrease in mass during the partial
in this potential region is sensitive to the presence of sulfonic oxidation of the film, thus evidencing a greater cation partici-
groups in the films. The greater is the percentage of these groups, pation in charge compensation. However, this decrease is minor
the greater is the extent of neutralization due to the incorporation in comparison with that obtained for the homopolymer. In con-
of species in the polymer matrix. The second aspect of interest clusion, the qualitative analysis of the curves suggests that the
is associated with the variation of mass during the first oxida- degree of film sulfonation influences cation participation in elec-
A.G. Cano Márquez et al. / Electrochimica Acta 52 (2007) 5294–5303 5303

troneutralization during the first oxidation process, because the [7] J. Yue, Z.H. Wang, K.R. Cromack, A.J. Epstein, A.G. MacDiarmid, J. Am.
mass decrease increases with the sulfonation degree. It should Chem. Soc. 113 (1991) 2665.
be mentioned that the electrosynthesis and EQCM test in H2 SO4 [8] M.T. Nguyen, P. Kasai, J.L. Millar, A.F. Diaz, Macromolecules 27 (1994)
3625.
(0.5 M) and HCS (0.5 M) were repeated obtaining similar results. [9] M.T. Nguyen, A.F. Diaz, Macromolecules 27 (1994) 7003.
[10] C. Li, S. Mu, Synth. Met. 149 (2005) 143.
4. Conclusion [11] H. Varela, S.L. de Albuerquerque Maranhão, R.M.Q. Mello, E.A. Ticianelli,
R.M. Torresi, Synth. Met. 122 (2001) 321.
[12] H. Tang, A. Kitani, T. Yamashita, S. Ito, Electrochim. Acta 43 (1998)
This study has demonstrated that the electro-oxidation of o- 2237.
ASA on PANI modified electrodes gives rise to the formation [13] L. Zhang, X. Jiang, L. Niu, S. Dong, Biosens. Bioelectron. 21 (2006) 1107.
of sulfonated homopolymers; these films, unlike those obtained [14] K. Krishnamoorthy, A.Q. Contractor, A. Kumar, Chem. Commun. 3 (2002)
on bare metallic electrodes, are thick and adherent. The quantity 240.
of electrodeposited P(o-ASA) is dependent on PANI thickness. [15] T. Tasuma, T. Ogawa, R. Sato, N. Oyama, J. Electroanal. Chem. 501 (2001)
180.
The result of EQCM shows that electroneutralization during the [16] C. Barbero, M.C. Miras, B. Schnyder, O. Haas, R. Kötz, J. Mater. Chem.
first oxidation process of P(o-ASA) is carried out mainly through 4 (1994) 1775.
the ejection of cations. Additionally, o-ASA and aniline copoly- [17] C. Barbero, R. Kötz, Adv. Mater. 6 (1994) 577.
mers were electrosynthetized, the sulfonation degree of the films [18] C. Thiemann, C.M.A. Brett, Synth. Met. 125 (2002) 445.
obtained was modulated with the proportion of monomers in the [19] J.Y. Lee, C.Q. Cui, X.H. Su, J. Electroanal. Chem. 360 (1993) 177.
[20] J.Y. Lee, X.H. Su, C.Q. Cui, J. Electroanal. Chem. 367 (1994) 71.
electrosynthesis solution. The proportion of charge compensa- [21] J.Y. Lee, C.Q. Cui, J. Electroanal. Chem. 403 (1996) 109.
tion derived from cation transport increases with the percentage [22] R. Mažeikienė, G. Niaura, A. Malinauskas, Synth. Met. 139 (2003) 89.
of sulfonic groups in the film. [23] H. Tang, A. Kitani, T. Yamashita, S. Ito, Synth. Met. 96 (1998) 43.
[24] R. Mažeikienė, A. Malinauskas, Mater. Chem. Phys. 83 (2004) 184.
[25] A. Kitani, K. Satoguchi, H.-Q. Tang, S. Ito, K. Sasaki, Synth. Met. 69
Acknowledgements (1995) 129.
[26] Y. Şahin, K. Pekmez, A. Yildiz, J. Appl. Polym. Sci. 90 (2003) 2163.
The support for this work by the Universidad Autónoma de [27] D.A. Fungaro, Sensors 1 (2001) 206.
San Luis Potosı́ via FAI is gratefully acknowledged. A.G. Cano [28] M. Pasquali, G. Pistoia, R. Rosati, Synth. Met. 58 (1993) 1.
Márquez thanks CONACYT scholarship (180914). [29] S. Cattarin, L. Douvoba, G. Mengoli, G. Zotti, Electrochim. Acta 33 (1988)
1077.
[30] J.C. Lacroix, P. Garcia, J.P. Audière, R. Clément, O. Kahn, Synth. Met. 44
References (1991) 89.
[31] H. Varela, R.M. Torresi, D.A. Buttry, J. Braz. Chem. Soc. 11 (2000) 32.
[1] A.A. Karyakin, A.K. Strakhova, A.K. Yatsimirsky, J. Electroanal. Chem. [32] R.M.Q. Mello, R.M. Torresi, S.I. Córdoba de Torresi, E.A. Ticianelli,
371 (1994) 259. Langmuir 16 (2000) 7835.
[2] C.M.A. Brett, C. Thiemann, J. Electroanal. Chem. 538 (2002) 215. [33] H. Varela, R.M. Torresi, D.A. Buttry, J. Electrochem. Soc. 147 (2000) 4217.
[3] C. Thiemann, C.M.A. Brett, Synth. Met. 123 (2001) 1. [34] G. Zauerbrey, Z. Phys. 155 (1955) 206.
[4] A. Benyoucef, F. Huerta, J.L. Vazquez, E. Morallon, Eur. Polym. J. 41 [35] G.A. Olah, G.K.S. Prakash, M. Arvanaghi, J. Am. Chem. Soc. 102 (1980)
(2005) 843. 6641.
[5] S. Mu, Synth. Met. 143 (2004) 259. [36] A. Malinauskas, Synth. Met. 107 (1999) 75.
[6] X.-L. Wei, Y.Z. Wang, S.M. Long, C. Bobeczko, A.J. Epstein, J. Am. Chem. [37] D. Orata, D.A. Buttry, J. Am. Chem. Soc. 109 (1987) 3574.
Soc. 118 (1996) 2545. [38] E.M. Geniès, J.F. Penneau, M. Lapkowski, New J. Chem. 12 (1988) 764.

You might also like