You are on page 1of 8

Synthetic Metals 161 (2011) 13531360

Contents lists available at ScienceDirect

Synthetic Metals
journal homepage: www.elsevier.com/locate/synmet

Solid-state oxidation of aniline hydrochloride with various oxidants


enkov a, , Elena N. Konyushenko a , Jaroslav Stejskal a , Miroslava Trchov a , Jan Prokes b
Ivana Sed
a
b

Institute of Macromolecular Chemistry, Academy of Sciences of the Czech Republic, Heyrovsky Sq. 2, 162 06 Prague 6, Czech Republic
Faculty of Mathematics and Physics, Charles University Prague, 182 00 Prague 8, Czech Republic

a r t i c l e

i n f o

Article history:
Received 5 January 2011
Received in revised form 21 April 2011
Accepted 29 April 2011
Available online 1 June 2011
Keywords:
Conducting polymer
Polyaniline
Solid-state polymerization
Photoacoustic spectroscopy
Silver
Mechanochemical route

a b s t r a c t
Aniline hydrochloride was oxidized in the solid state with three different oxidants: ammonium peroxydisulfate, iron(III) chloride, and silver nitrate. Polyaniline salt was obtained after a few hours when
ammonium peroxydisulfate was used as an oxidant. The polymerization of aniline hydrochloride with
silver nitrate leads to polyaniline only after several days; in the case of iron(III) chloride, aniline oligomers
were obtained. The conductivity of the polyaniline was 0.21 S cm1 when ammonium peroxydisulfate was
applied; it is comparable with the conductivity of a standard polyaniline. The oxidation with silver nitrate
yields a composite material, polyaniline salt and silver particles, with conductivity 1.5 103 S cm1 . Photoacoustic spectroscopy was employed to study the kinetics of the oxidation reaction. Infrared and UVvis
spectra were efcient tools to characterize the nal products, and to compare their molecular structure
with that of the polyaniline prepared under standard conditions in aqueous medium.
2011 Elsevier B.V. All rights reserved.

1. Introduction
Polyaniline (PANI) is probably the most common conducting
polymer. It is usually prepared by the oxidation of aniline in acidic
aqueous media. Depending on the acidity of the reaction and
the concentrations of reactants, various morphologies have been
obtained [1], such as nanotubes [24], nanobres [5,6], nanogranules, microspheres [2,7] or more complex hierarchical objects
[810].
The oxidation proceeding in the absence of solvents is of special
interests in understanding the chemistry of aniline oxidation and
is an easy single-step method to prepare new material with interesting properties, e.g., conducting material incorporating metal
nanoparticles. Solid-state reactions, in which the solid-phase reactants are used exclusively, including PANI, can be divided into
several groups: (1) the protonation of PANI, when PANI base reacts
with solid acids to give PANI salts [1113], (2) the oxidation of PANI
with various oxidants [12,1416], and (3) the oxidation of aniline
salts with ammonium peroxydisulfate (APS) [1720] or other oxidants [21]. The last type of reactions may involve a solvent, which
is frozen [22,23], i.e., it is a solid, or other solid additives, such as
montmorillonite [24] or heteropolyacids [25]. The mechanochemical preparations of other conducting polymers, such as polypyrrole
[26,27] or polythiophene [26], have also recently been reported.

Corresponding author at: Institute of Macromolecular Chemistry, Academy of


Sciences of the Czech Republic, Heyrovsky Sq. 2, 162 06 Prague 6, Czech Republic.
enkov).
E-mail address: sedenkova@imc.cas.cz (I. Sed
0379-6779/$ see front matter 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.synthmet.2011.04.037

The solid-state redox reactions between aniline salt and an inorganic oxidant are of fundamental importance. In such reaction,
electrons are transferred from the aniline molecules to those of the
oxidant. It has been proposed that such a transfer is much easier
when a conducting polymer is present or generated in the system
[1,28,29]. Transfer of electrons between the molecules thus does
not require their direct contact. For that reason, the oxidative reactions proceed efciently even when the diffusion of molecules is
restricted, such as in the solid state or in frozen media.
The present contribution deals with the solid-state oxidations of
aniline hydrochloride with three oxidants differing in the oxidation
potential, 2.0 V (S2 O8 2 + 2e 2SO4 2 ), 0.77 V (Fe3+ + e Fe2+ )
and 0.80 V (Ag+ + e Ag) relative to the standard hydrogen electrode. The different oxidation potentials have signicant effect on
the structure and conductivity of the obtained product. Further,
oxidation with silver nitrate yields a material with incorporated
silver nanoparticles.
2. Experimental
2.1. Solid-state oxidations
Aniline hydrochloride (20 mmol, 2.59 g) was reacted in the solid
state gradually with three oxidants: APS (25 mmol, 5.71 g), iron(III)
chloride hexahydrate (50 mmol, 13.50 g), or silver nitrate (50 mmol,
8.49 g). Aniline salt and oxidant were mixed with a pestle and
mortar for 5 min and then left at rest in air. The progress of the
reaction was monitored (1) by recording the conductivity of the
pellet made from aniline hydrochloride and oxidant by the com-

1354

I. S edenkov et al. / Synthetic Metals 161 (2011) 13531360

pression at 700 MPa, and (2) in the case of APS, by measuring the
changes of photo acoustic (PA) spectra of the powder. After one day
for APS, one week for iron(III) chloride and silver nitrate, the resulting powders were quickly rinsed with excess of ethanol to remove
non-reacted aniline monomer and soluble aniline oligomers, and
then with water to remove the residual oxidant or its reaction products, such as ammonium sulfate in the case of APS. Parts of the
dried sample were deprotonated to PANI bases by immersion in
1 M NH4 OH.
A reference sample, referred to as standard PANI salt, was
prepared by the oxidation of 0.2 M aniline hydrochloride (Fluka,
Switzerland) with 0.25 M ammonium peroxydisulfate (APS) (LachNer, Czech Republic) [28,30]. The solids were separated by ltration
the next day, rinsed with acetone, dried in air and then over silica
gel at room temperature.

4n

NH2HCl

+
NH
Cl

3. Results and discussion


The oxidation of aniline hydrochloride with APS, iron(III) chloride or silver nitrate in the solid state was studied (Fig. 1). Sulfuric,
hydrochloric or nitric acid is formed as by-product, respectively,
and they, together with the rest of water absorbed in the reactants,
subsequently play an important role in the solvation of the arising
products and their protonation. The sulfuric acid with absorbed
water was even released from the sample compressed into the
pellet as droplets of liquid in the case of oxidation with APS (Fig. 2).

H
N
+
NH
Cl

N
H

5n H2SO4

(FeCl3)

5n (NH4)2SO4

2n HCl

2.2. Characterization
Fourier-transform infrared (FTIR) spectra of the samples dispersed in a potassium bromide matrix were recorded in the
wavenumber range of 4004000 cm1 at 64 scans per spectrum
at 2 cm1 resolution using a fully computerized Thermo Nicolet
NEXUS 870 FTIR Spectrometer with a DTGS TEC detector. Spectra
were corrected for the moisture and carbon dioxide in the optical
path. Spectra measured during the solid-state oxidation of aniline
were recorded with a PA cell, Mtec model 300 using a helium purge.
Carbon black was used as reference. The number of scans was
adjusted to 64 with resolution 8 cm1 in accordance with mirror
velocity 0.16 cm s1 .
UVvisible spectra of PANI base dissolved in Nmethylpyrrolidone were recorded with a spectrometer Lambda 20
(Perkin-Elmer, UK). The conductivity of composites was measured
with a four-point van der Pauw method on pellets 13 mm in
diameter and 1 mm thick compressed at 700 MPa with a manual
hydraulic press using a current source SMU Keithley 237 and a
Multimeter Keithley 2010 voltmeter with a 2000 SCAN 10-channel
scanner card. The scanning electron microscope (SEM) JEOL 6400
and transmission electron microscope (TEM) JEOL JEM 2000 FX
have been used to characterise the morphology of the samples.

+ 5n (NH4)2S2O8

NH2HCl

c
4n

NH2HCl

+
NH
Cl

+ 10n

oligomers

AgNO3

H
N
+
NH
Cl

2n HCl

N
H

10n Ag

10n HNO3

Fig. 1. The simplied scheme of oxidation of aniline hydrochloride with (a) ammonium peroxydisulfate, (b) iron(III) chloride, and (c) silver nitrate.

molecular structure. Finally, FTIR spectroscopy was used to characterize the nal products of the oxidation reaction.
3.1.1. Conductivity
Polyaniline is a conducting material [30,31], although the starting compounds used in its preparation are insulators. Therefore,

3.1. Oxidation of aniline hydrochloride with ammonium


peroxydisulfate
Generally, the oxidation of aniline hydrochloride by the commonly used oxidant, APS, gives PANI, sulfuric or hydrochloric acid,
and ammonium sulfate (Fig. 1a). The solid-state reaction started
immediately when monomer and oxidant were mixed, and the
colour changed from white to light blue and nally to dark green.
APS is one of the strongest oxidants which has oxidation potential 2.0 V and the solid-state oxidation of aniline hydrochloride by
APS was nished in several hours. The kinetics of the oxidation was
illustrated by recording the changes of resistivity with time. Photoacoustic spectroscopy was used to follow the evolution of the

Fig. 2. The pellet made from a mixture of aniline hydrochloride and ammonium
peroxydisulfate 2 h after pressing into the pellet. The mixture of by-product, solution
of ammonium sulfate and ammonium hydrogen sulfate in concentrated sulfuric acid
with absorbed water, is released as a droplet of liquid.

I. S edenkov et al. / Synthetic Metals 161 (2011) 13531360

Fig. 3. Changes of APS/aniline hydrochloride pellet resistance with time.

the monitoring of conductivity changes during the oxidation of the


monomer in the solid state reveals if the conducting PANI salt was
obtained or not, and what time is needed for the completion of this
process.
Aniline hydrochloride was mixed with APS and the pellet was
prepared (Fig. 2). Two Pt electrodes were deposited on the pellet to
follow the resistivity changes (Fig. 3). At the beginning of the measurement the resistivity was high, more then 250 , and it dropped
exponentially down to 50  after 16 min (Fig. 3). Further decrease
in resistivity was signicant and reached a limit of 25  after 2 h.
The relatively low nal conductivity of the pellet is due to the presence of non-conducting products, such as residual reactants and
by-products.
In the case of free powders, the nal product was quickly rinsed
by excess of ethanol to remove unreacted monomer and then by
water to remove soluble by-products and excess of sulfuric acid
unbounded as the dopant in the PANI chains. In this way, the
polymerization of aniline hydrochloride during rinsing has been
avoided. The conductivity of the dry product was 0.21 S cm1 ,
a value close to the conductivity of PANI salt prepared by the
standard procedure, 4.4 S cm1 [30]. After deprotonation of PANI
hydrochloride, the conductivity dropped up to 1.6 109 S cm1 ,
i.e., again the typical behaviour of PANI salt after conversion to PANI
base was observed.
The above experiment illustrates the feasibility and the progress
of solid-state oxidation. As the start of the process is not well
dened, due to the time elapsed between the compression of the
pellet and the start of electric measurement, the qualitative evaluation of the reaction kinetics was not possible. It was also observed
that the polymerization in compressed pellets was faster than that
in the free powder because of more intimate contact between reactants. The nal product had a granular morphology (Fig. 4a) similar
to that of the PANI prepared in solution in a standard way [30,32].
3.1.2. Photoacoustic spectroscopy
The usual technique of FTIR spectroscopy, used to characterize
powdered samples, is based on their dispersion in the potassium
bromide matrix and to form pellets. The matrix can lead to separation of the reactants and, subsequently, the reaction in the KBr
pellet could be slowed down or blocked. The method of attenuated total reection (ATR) also allows the recording of the spectra
of the powders but additional pressure has to be applied to the
reactants to secure a sufcient contact between the sample and
an ATR crystal. From this point of view, the PA spectroscopy gives
us the opportunity to monitor the chemical changes occurring

1355

during the reaction without any additional manipulation with sample.


Photoacoustic spectra measured in situ during the solid-state
oxidation of aniline hydrochloride by APS exhibit a rather low
signal-to-noise ratio due to the small number of scans and the
higher mirror velocity (Fig. 5). This experimental setup was chosen to obtain a spectrum in a reasonably short time interval and to
acquire the spectra reecting the material evolution. Despite lower
spectral quality, the spectral features of both reactants, APS and
aniline hydrochloride, were well observed in the rst spectrum
(time 0:00) (Fig. 5). The most intensive broad band in the region
above 2500 cm1 is the combination of the bands from APS and
aniline hydrochloride spectrum. The band of medium intensity at
about 2000 cm1 is typical of aniline hydrochloride and is probably
an overtone enhanced by the presence of the negatively charged
chloride anions. Two relatively intensive bands of APS at 1420 and
1280 cm1 , assigned to vibration of SO4 group, are superimposed
with several sharp peaks of the aniline spectrum, linked to the ring
vibration [33]. The sharp band at 1050 cm1 conrms the presence
of SO4 moiety in APS [33]. The sharpness of the aniline peaks
reects the fact that aniline hydrochloride is crystalline.
The evolution of the molecular structure is described by the
subsequent spectra (Fig. 5). The main changes observed in the PA
spectra are connected to the gradual depletion of the reactants. The
decrease in the intensity of the main absorption bands is remarkable. Unfortunately, PANI hydrochloride powder has a signicantly
lower PA signal in comparison with aniline monomer or APS. Still,
some changes indicating the conversion to PANI are found. We can
observe an increase in intensity in the region above 2000 cm1 ,
assigned to a polaron band [34], which is associated with free
charge carriers in the conducting form of PANI. The intensity ratio
of two bands of APS present at 1420 and 1280 cm1 in the rst
spectrum changes, the latter becomes broader and its intensity
lower at the end of the reaction. This corresponds to the decomposition of APS into ammonium sulfate or ammonium hydrogen
sulfate, which both have the second absorption band shifted to
lower wavenumbers. The sharp band at 1050 cm1 decreased as
well. The diminishing of the band at about 2000 cm1 reects the
consumption of aniline hydrochloride.
The course of the oxidation was followed by the intensity of
bands connected with aniline hydrochloride at 2000 cm1 and APS
at 1050 cm1 . The integral intensity of both bands reects the fact
that the main part of both reactants was depleted in the rst hour
(Fig. 6). The reaction then signicantly slowed down. The exhaustion of both reactants coincides, and we can assume that the aniline
hydrochloride is continuously oxidized by APS [35].
After one day, the monomer was still present. The PA spectrum
of sample taken after one week, however, does not reveal any signicant spectral features of aniline hydrochloride.
3.1.3. FTIR spectra of the nal product
After one week, the FTIR spectrum of the as-prepared reaction
product was recorded in potassium bromide pellets, in addition to
PA spectra. The spectrum of a reference sample of PANI hydrochloride prepared by standard polymerization in aqueous solution
[30] was recorded as well as the spectra of the reactants and the
expected by-products (Fig. 7).
The spectrum of the sample prepared by solid-state oxidation
of aniline salt with APS (Fig. 7b) reects its complex composition. The most intense broad band at 1153 cm1 with a shoulder
at 1291 cm1 probably comprises the bands of by-products, which
have not become separated from the reaction mixture at this stage.
Ammonium sulfate (Fig. 7c) has an intense band at 1102 cm1
and ammonium hydrogen sulfate (Fig. 7e) at 1200 cm1 with a
shoulder at 1286 cm1 . Both possible by-products have a band in
the region of 14001450 cm1 , the ammonium sulfate spectrum

1356

I. S edenkov et al. / Synthetic Metals 161 (2011) 13531360

Fig. 4. Scanning electron microscopies of PANI prepared with (a) ammonium peroxydisulfate, (b) iron(III) chloride, and (c) silver nitrate.

exhibits a band at 1404 cm1 and ammonium hydrogen sulfate at


1430 cm1 . By their overlap, the broad intense band with maximum
at 1412 cm1 is formed in the spectrum of the nal product.
To detail inspection of the nal product, the spectrum of ammonium sulfate (Fig. 7c), one of the main by-products, was subtracted
from the spectrum of the product of solid-state oxidation (Fig. 7b).
The differential spectrum (Fig. 7d) then reveals the traces of reactants, both aniline hydrochloride and APS. At least three most
intense and sharp peaks from the aniline hydrochloride spectrum (Fig. 7f) can be located in the spectrum of the nal product
observed at 1498, 747 and 688 cm1 . These bands are only slightly
shifted from their position in the spectrum of aniline hydrochloride at 1494, 741 and 686 cm1 . This can be assigned, together with
their broadening, to the interaction with other sample components
and/or to the loss of crystalline structure. A shoulder at 1049 cm1
and a band at 558 cm1 can be linked to the spectrum of APS
(Fig. 7g). Finally, we observe spectral features typically found in the
spectrum of PANI hydrochloride. The intensity of the broad band
at about 1600 cm1 associated with the quinonoid-rings vibration
dominates the intensity of the band at about 1480 cm1 assigned to
the vibration of benzenoid rings [3436]. This may indicate a partial overoxidation of the sample or the presence of over-oxidized
oligomers [37]. The band at 1302 cm1 is connected to -electron
delocalization [34]. The broad band at 1180 cm1 can be, at least
partially, assigned to the vibration of the NH+ group, too, which
band is present in the spectrum of PANI prepared in aqueous
medium at 1145 cm1 [38,39]. It interferes with the band from the
spectrum of ammonium hydrogen sulfate at 1200 cm1 , which is
possibly due to a by-product. The strong band at 875 cm1 reveals
an abundance of trisubstituted aromatic rings, which is responsi-

ble for high amount of irregularities, such as branching of the PANI


chains [35,40]. Finally, spectral features of a second possible byproduct, ammonium hydrogen sulfate (Fig. 7e), are presented in
the differential spectrum 7d as well.
The FTIR spectrum of the insoluble part of the product was
recorded after extraction of the as-prepared sample with ethanol
and water. This procedure removes soluble by-products and residual reactants. The remaining sample exhibits the same spectral
features (Fig. 8a) as PANI prepared by standard polymerization
in aqueous solution [30]. Bands assigned to quinonoid and benzenoid ring vibrations are observed at 1560 and 1475 cm1 [36].
The former band has a shoulder at about 1600 cm1 characteristic of C C ring vibration, which is observed only as the symmetry
is broken by conformational changes induced by protonation [34].
The band at 1298 cm1 is linked to the -electron delocalization.
Two bands characteristic of the conducting emeraldine salt are
located at 1243 cm1 and 1128 cm1 . The former is attributed to
vibrations of the CN+ group, the latter to those of, the NH+
group. The presence of the sulfate counter-ions is proved by the
band at 1039 cm1 . The band due to the out-of-plane vibration of
the adjacent hydrogen atoms on the para-substituted benzenoid
ring is usually found at about 825 cm1 [38,39]. In case of solidstate oxidation, the band is shifted to 796 cm1 , together with two
other weak bands at 924 and 876 cm1 . These differences indicate
the higher rate of ortho- or 1,2,4-substituted aromatic rings in the
material [35,40].
The spectrum of the sample deprotonated by ammonium
hydroxide (Fig. 8b) is consistent with the spectrum of the emeraldine base prepared by the standard procedure (Fig. 8f) [30]. The
main difference is the presence of a local maximum at 1145 cm1 on

I. S edenkov et al. / Synthetic Metals 161 (2011) 13531360

1357

Fig. 7. FTIR spectra of (a) PANI salt prepared by standard oxidation of aniline
hydrochloride with APS in aqueous solution [30]; (b) as-prepared product of
solid-state oxidation of aniline with APS; (c) ammonium sulfate; (d) a product of
solid-state aniline oxidation with APS after subtraction of the spectrum of ammonium sulfate; (e) ammonium hydrogen sulfate; (f) aniline hydrochloride; (g) APS.

typical absorption spectrum of blue PANI base has two distinct


absorption bands located between 315345 nm, which corresponds to a * transition, and 610645 nm, assigned to the
n* transition corresponding to the creation of an intermolecular charge-transfer exciton, typical of the emeraldine base [41],
depending on the preparation and/or processing of PANI [42]. PANI
prepared by the solid-state reaction and in the standard way using
APS as an oxidant [30] had spectra similar to each other. Two main
absorption maxima at 334 and 630 nm were observed and they
correspond to the characteristic blue colour of PANI base (Fig. 9a).

Fig. 5. The progress of aniline hydrochloride oxidation with ammonium peroxydisulfate followed by photoacoustic spectroscopy. Reaction times (in hours) are
given on the individual spectra. The photoacoustic spectra of reactants are shown
for comparison.

the band at 1165 cm1 . The band at 1145 cm1 should be assigned
to vibration of the secondary amine group NH and/or to in-plane
deformation vibration of CH groups on aromatic rings [38].

3.2. Oxidation of aniline hydrochloride with iron(III) chloride


A product with the aky morphology was found after the oxidation of aniline hydrochloride with iron(III) chloride (Fig. 4b). It was
non-conducting, but the precise value of the conductivity could not
be determined because of the poor integrity of the compressed pellets used in electrical measurements. The brown product obtained
in the reaction of monomer with iron(III) chloride had only a single

3.1.4. UVvisible spectra


UVvisible spectra of the nal products deprotonated by ammonia and dissolved in N-methylpyrrolidone are shown in Fig. 9. A

Fig. 6. Changes of integral intensity of photoacoustic signal of bands at 2000 cm1


and 1050 cm1 with time.

Fig. 8. FTIR spectra of products of aniline oxidation with various oxidants in the
solid state (after removal of residual reactants and by-products): (a) aniline oxidized
with APS; (b) deprotonated product of aniline oxidation with APS; (c) deprotonated
product of aniline oxidation with iron(III) chloride; (d) aniline oxidized with silver
nitrate; (e) deprotonated product of aniline oxidation with sliver nitrate; (f) PANI
base prepared by the standard polymerization of aniline chloride in solution.

I. S edenkov et al. / Synthetic Metals 161 (2011) 13531360

1358

a 1.0
solid-state polymerization

Absorbance

standard polymerization

0.5

0.0
400

600

800

substituted benzene ring. The band at 1638 cm1 is ascribed to


stretching vibration of the C O in a quinone [44,45]. The band at
1500 cm1 is connected to the vibration of a quinonoid ring. The
weak band at 1409 cm1 reects the vibrations of phenazine-like
units [35,43]. The presence of primary aromatic amines is proved
by the sharp bands at 3590 cm1 and 3471 cm1 [46]. In the region
of the ring-deformation vibration at about 600 cm1 , the band with
its main maximum at 586 cm1 is formed from overlapping bands
belonging to vibrations of benzene rings substituted in different
positions.
The results of conductivity measurement, FTIR and UVvisible
spectroscopy indicates that the product of the solid-state aniline oxidation with iron(III) chloride is a mixture of aniline
oligomers with a complex molecular structure. We can identify
iminoquinonoid units, cyclic phenazine-like structures, monosubstituted aromatic rings and amines. We assume that reaction of
aniline monomer with iron(III) chloride leads to branched products
of low molecular weight.

Wavelength, nm
3.3. Oxidation of aniline hydrochloride with silver nitrate

Absorbance

b 1.0

Composite materials containing PANI were also prepared with


silver nitrate as the oxidant. A granular PANI structure was formed
(Fig. 4c), accompanied by silver nanoparticles (Fig. 10). The conductivity of the composite material was 1.5 103 S cm1 for the salt
form and 9.2 106 S cm1 for the PANI base. Two absorption maxima were observed in the UVvisible spectrum of PANI prepared
with silver nitrate (Fig. 9b). One is at 380 nm, which corresponds
to * transition of a benzenoid ring, and the second, located at
610 nm, can be assigned to a n* transition. The latter maximum
is less pronounced compared with the PANI prepared with APS.

FeCl3
0.5
AgNO3

0.0
400

600

800

Wavelength, nm
Fig. 9. UVvisible spectra of (a) aniline hydrochloride oxidized with ammonium
peroxydisulfate in water (full line) and in the solid state (dashed line); (b) aniline
hydrochloride oxidized in the solid state with iron(III) chloride (dashed line) or silver
nitrate (full line).

absorption maximum at 358 nm, which corresponds to presence


of over-oxidized oligomers [1] (Fig. 9b) and to the absence of a
polymer component.
3.2.1. FTIR spectra of the nal product
The nal product was washed and dried as described above.
The polymerization with iron(III) chloride or silver nitrate in an
aqueous medium is preceded by a longer induction period than
in the case of the oxidation with APS [9], and the residual reactants are separated before the reaction in the aqueous medium
could start. The spectrum of a sample prepared by the solid-state
oxidation of aniline with iron(III) chloride does not exhibit any
spectral features of PANI (Fig. 8c). The most intense band in the
spectrum is the triplet with local maxima at 1178 cm1 , 1119 cm1 ,
and 1079 cm1 . The band at 1178 cm1 corresponds to the inplane bending vibration of the CH group on an aromatic ring
or the vibration of the N Q N structure in the oxidized species.
The deformation vibrations of the CH group are responsible for
the band at 1119 cm1 as well, and both peaks were observed in
the spectra of polyphenosafranine [43]. The last band from the
triplet at 1097 cm1 is associated with the vibration of a mono-

3.3.1. FTIR spectra of the nal product


The bands typical of quinonoid and benzenoid-ring vibrations in
PANI chains can be found at 1575 and 1494 cm1 in the spectrum
(Fig. 8d). Their shift to higher wavenumbers by comparison with the
spectrum 8a indicates the lower protonation level of the material
[36]. A strong sharp peak of the nitrate anion observed at 1384 cm1
in the spectrum reects the protonation of PANI by nitric acid,
which is a reaction by-product. The band assigned to -electron
delocalization in the conducting PANI form is present at 1297 cm1 .
The band linked to the stretching vibration of CN+ is observed at
1243 cm1 . The lower intensity of the band at 1137 cm1 , which
is connected to the vibration of the NH+ group, conrms the
assumption that the protonation level of the reaction product is
lower [34,39].
The spectrum of the deprotonated sample is devoid of the features connected with counter-ions, and it reveals more about the
molecular structure of PANI (Fig. 8e). The spectrum of the deprotonated sample differs from the spectrum of standard PANI base
prepared with APS oxidant [30] (Fig. 8f). The rst band atypical
of emeraldine base is found at 1638 cm1 and can be assigned to
vibration of the carbonyl group in quinones. The band attributed
to vibration of the quinonoid ring is observed at 1568 cm1 . The
band assigned to the vibration of benzenoid rings in structure is
situated at 1492 cm1 . The band at 1444 cm1 , which is usually
found in the spectra of PANI oligomers [2,35], is well observed.
Assignment of the last band is still under debate and it may be connected with the presence of the C N stretching or N N stretching
vibration in azobenzene [33,46]. The band typical of emeraldine
base, attributed to CN stretching vibration in neighborhood of a
quinonoid ring, usually found at 1375 cm1 , is shifted to 1357 cm1 .
The broad band present in the spectrum of standard PANI base at
1300 cm1 is reduced to a weak sharp peak at 1288 cm1 . Similar
features were observed in the spectra of samples prepared in alka-

I. S edenkov et al. / Synthetic Metals 161 (2011) 13531360

1359

Fig. 10. Transmission electron microscopy of PANI prepared with (a) ammonium peroxydisulfate and (b) silver nitrate.

line media [2]. An additional weak band observed at 1175 cm1


belongs to vibration of group N Q N and to the bending vibration of the CH group on an aromatic ring [37]. The band at around
830 cm1 , associated with the deformation vibration of the CH
group on a 1,4-disubstituted aromatic ring, is signicantly suppressed. At the same time, the band at 740 cm1 connected with
the vibration of the CH group on a mono or 1,2-disubstituted ring
is enhanced [40].
The FTIR spectrum of the sample prepared by the solid-state
oxidation of aniline hydrochloride with silver nitrate after deprotonation clearly indicates that this reaction does not lead to regular
PANI chains. It is probable that shorter PANI chains and oligomers,
in which aniline constitutional unites are coupled not only in the
para-position but also in the ortho-position or with azo links, are
produced. The use of the oxidant with lower oxidation potential
than that of APS, silver nitrate, leads to a higher fraction of aniline
oligomers. Despite this fact, the spectrum of the protonated product had spectral features typical of the conducting emeraldine salt.
Concentrated nitric acid, present as a by-product of the reaction,
was able to protonate both oligomers and conducting structures as
they were formed.

4. Conclusions
Aniline hydrochloride was oxidized in the solid state by three
oxidants with different oxidation potentials: ammonium peroxydisulfate, iron(III) chloride, and silver nitrate. Solid-state reaction
with APS leads to a conducting product with conductivity of
0.21 S cm1 , and non-conducting material was obtained after reaction with iron(III) chloride. The oxidation with silver nitrate leads
to a product with conductivity 1.5 103 S cm1 . The conducting powder had granular morphology; while that prepared with
iron(III) chloride resembled akes. It was shown that, for the oxidant with the highest oxidation potential, APS, the polymerization
proceeds well even in the solid state and the product conductivity and molecular structure are comparable to those of the PANI
prepared by the standard procedure in aqueous solution. Photoacoustic spectroscopy was applied to study the kinetics of the
reaction with APS. The reaction was completed within one day in
pellets, and in a few days in the free powder. Iron(III) chloride has
the lowest oxidation potential and the result of FTIR and UVvisible
spectroscopy support the hypotheses that the reaction stops with
the formation of non-conducting oligomers. When silver nitrate

was used as an oxidant, the composite material with silver was


obtained, in which both PANI chains and aniline oligomers are
present.
Acknowledgements
The authors thank the Grant Agency of the Academy of Sciences
of the Czech Republic (IAA 100500902 and IAA 400500905) and the
Czech Grant Agency (203/08/0686) for nancial support.
References
[1] J. Stejskal, I. Sapurina, M. Trchov, Prog. Polym. Sci. 35 (2010) 14201481.
[2] J. Stejskal, I. Sapurina, M. Trchov, E.N. Konyushenko, Macromolecules 41
(2008) 35303536.
[3] C. Laslau, Z. Zujovic, J. Travas-Sejdic, Prog. Polym. Sci. 35 (2010) 14031419.
[4] Z.F. Ding, D. Yang, R.P. Currier, S.J. Obrey, Macromol. Chem. Phys. 211 (2010)
627634.
[5] C.-H. Yang, Y.-K. Chin, H.-E. Cheng, C.-H. Chen, Polymer 46 (2005) 1068810698.
[6] B.-H. Kim, D.H. Park, J. Jee, S.-G. Yu, S.-H. Lee, Synth. Met. 150 (2005) 279284.
[7] Y. Zhu, G. Ren, M. Wan, L. Jiang, Macromol. Chem. Phys. 210 (2009) 20462051.
[8] Z. Zhang, S. Sui, L. Zhang, M. Wan, Y. Wei, L. Yu, Adv. Mater. 17 (2005)
28542857.
c-Marjanovi

Polymer
[9] N.V. Blinova, J. Stejskal, M. Trchov, I. Sapurina, G. Ciri
c,
50 (2009) 5056.
[10] X.C. Wang, Y. Li, Y. Zhao, J. Liu, S. Tang, W. Feng, Synth. Met. 160 (2010),
20082014.
[11] J. Stejskal, I. Sapurina, M. Trchov, J. Prokes, I. Krivka, E. Tobolkov, Macromolecules 31 (1998) 22182222.
[12] M. Trchov, J. Stejskal, J. Prokes, Synth. Met. 101 (1999) 840841.
[13] W. Gu, Z.-Q. Li, S.-M. Zhu, D. Zhang, Acta Chim. Sinica 66 (2008) 10971101.

[14] I. Sedenkov,
M. Trchov, J. Stejskal, J. Prokes, ACS Appl. Mater. Interfaces 1
(2009) 19061912.
[15] K. Gupta, G. Chakraborty, S. Chatak, P.C. Jana, A.K. Meikap, J. Appl. Polym. Sci.
115 (2009) 29112917.
[16] H. Ding, Y. Long, J. Shen, M. Wan, J. Phys. Chem. B 114 (2010) 115119.
[17] J.X. Huang, J.A. Moore, J.H. Acquaye, R.B. Kaner, Macromolecules 38 (2005)
317321.
[18] X.-S. Du, C.-F. Zhou, G.-T. Wang, Y.-W. Mai, Chem. Mater. 20 (2008) 38063808.
[19] S. Bhadra, N.H. Kim, K.Y. Rhee, J.H. Lee, Polym. Int. 58 (2009) 11731180.
[20] O.Yu. Posudievsky, O.A. Goncharuk, R. Barille, V.D. Pokhodenko, Synth. Met. 160
(2010) 462467.
[21] C.-F. Zhou, X.-S. Du, Z. Liu, S.P. Ringer, Y.-W. Mai, Synth. Met. 159 (2009)
13021307.

A. Riede, M. Helmstedt, P. Mokreva, J. Prokes, Polymer


[22] J. Stejskal, M. Sprkov,
40 (1999) 24872492.
[23] E.N. Konyushenko, J. Stejskal, M. Trchov, N.V. Blinova, P. Holler, Synth. Met.
158 (2008) 927933.
[24] I. Bekri-Abbes, E. Drasta, React. Funct. Polym. 70 (2009) 1118.
[25] J. Gong, X.-J. Cui, Z.-W. Xie, S.-G. Wang, L.-Y. Qu, Synth. Met. 129 (2002)
187192.
[26] O.Yu. Posudievsky, O.A. Goncharuk, V.D. Pokhodenko, Synth. Met. 160 (2010)
4751.

1360

I. S edenkov et al. / Synthetic Metals 161 (2011) 13531360

[27] S.P. Palaniappan, P. Manisankar, Mater. Chem. Phys. 122 (2010) 1517.
[28] E.N. Konyushenko, M. Trchov, J. Stejskal, I. Sapurina, Chem. Pap. 64 (2009)
5664.
[29] I. Sapurina, J. Stejskal, Polym. Int. 57 (2008) 12951325.
[30] J. Stejskal, R.G. Gilbert, Pure Appl. Chem. 74 (2002) 857867.
[31] J. Stejskal, A. Riede, D. Hlavat, J. Prokes, M. Helmstedt, P. Holler, Synth. Met.
96 (1998) 5561.
enkov, M. Trchov, I. Sapurina, M. Cieslar,
[32] E.N. Konyushenko, J. Stejskal, I. Sed
J. Prokes, Polym. Int. 55 (2006) 3139.
[33] L.J. Bellamy, The Infra-red Spectra of Complex Molecules, third ed., Chapman
and Hall, London, 1975.
[34] Z. Ping, J. Chem. Soc. Faraday Trans. 92 (1996) 30633067.
enkov, E.N. Konyushenko, J. Stejskal, P. Holler, G. Ciri
c
[35] M. Trchov, I. Sed
J. Phys. Chem. B 110 (2006) 94619468.
Marjanovic,
[36] Y. Furukawa, F. Ueda, Y. Hyodo, I. Harada, T. Nakajima, T. Kawagoe, Macromolecules 21 (1988) 12971305.
[37] M. Hasik, C. Paluszkiewitz, E. Wenda, Vib. Spectrosc. 29 (2001) 191
195.

[38] E.T. Kang, K.G. Neoh, K.L. Tan, Prog. Polym. Sci. 23 (1998) 277324.
[39] M.I. Boyer, S. Quillard, E. Rebourt, G. Louarn, J.P. Buisson, A. Monkman, S. Lefrant,
J. Phys. Chem. B 102 (1998) 73827392.
[40] M. Hasik, A. Drelinkiewicz, E. Wenda, C. Paluszkiewicz, S. Quillard, J. Mol. Struct.
596 (2001) 8999.
[41] G.G. Wallace, G.M. Spinks, L.A.P. Kane-Maguire, P. Teasdale, Conductive Electroactive Polymers, Intelligent Materials Systems, second ed., CRC Press, Boca
Raton, 2003.
[42] J. Stejskal, P. Kratochvl, N. Radhakrishnan, Synth. Met. 61 (1993) 225231.
c-Marjanovi

N.V. Blinova, M. Trchov, J. Stejskal, J. Phys. Chem. B 111


[43] G. Ciri
c,
(2007) 21882199.
[44] J. Krz, L. Sarovoytova, M. Trchov, E.N. Konyushenko, J. Stejskal, J. Phys. Chem.
B 113 (2009) 66666673.
[45] Z.D. Zujovic, L. Zhang, G.A. Bowmaker, B.A. Kilmartin, J. Travas-Sejdic, Macromolecules 41 (2008) 31253135.
[46] G. Socrates, Infrared and Raman Characteristic Group Frequencies, third ed.,
Wiley, New York, 2001.

You might also like