You are on page 1of 9

Electrochimica Acta 367 (2021) 137566

Contents lists available at ScienceDirect

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

Coulometric response of solid-contact anion-sensitive electrodes


Tingting Han, Ulriika Mattinen, Zekra Mousavi, Johan Bobacka∗
Åbo Akademi University, Johan Gadolin Process Chemistry Centre, Laboratory of Molecular Science and Engineering, Biskopsgatan 8, Åbo FI-20500, Finland

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

Article history: In this work, the coulometric response of nitrate, perchlorate, and sulfate solid-contact anion-sensitive
Received 8 September 2020 electrodes was investigated. The coulometric transduction method was originally introduced and so far
Revised 1 November 2020
mainly studied for solid-contact cation-selective sensors using poly(3,4-ethylenedioxythiophene) (PEDOT)
Accepted 24 November 2020
doped with poly(styrene sulfonate) as transducer. Here we provide additional proof-of-concept for the
Available online 27 November 2020
coulometric transduction method by focusing on the electrochemical characteristics of anion sensors. PE-
Keywords: DOT was electrodeposited in the presence of small anions, including chloride, nitrate, sulfate, and per-
Solid-contact anion-sensitive electrodes chlorate. The counterion was found to influence the yield of electroactive PEDOT, which is an important
Anion sensors parameter for the coulometric response. Anion-sensitive electrodes were prepared by coating the PEDOT
Coulometric responses solid contact with plasticized PVC-based anion-sensitive membranes by drop-casting or spin-coating. The
Impedance influence of the thickness of the PEDOT film and the anion-sensitive membrane on the coulometric re-
PEDOT
sponse was studied in detail. The solid-contact anion sensors showed fast charge transfer and ion trans-
port properties, making them suitable for coulometric sensing. Also for anion-sensitive electrodes, the
analytical signal was amplified by increasing the thickness of the PEDOT solid contact, which is fully
consistent with earlier works on cation-selective electrodes. This work shows that the coulometric trans-
duction principle is feasible and robust for various types of solid-contact ISEs.
© 2020 The Author(s). Published by Elsevier Ltd.
This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/)

1. Introduction ride ions [11]. PEDOT has been used less frequently as ion-to-
electron transducer in anion sensors [17–21], compared to cation-
Ion-selective electrodes (ISEs) are well-known electrochemical selective SCISEs. Actually, oxidized (p-doped) conducting polymers,
sensors that are used in a broad range of applications, such as such as PEDOT, should be well suited as transducers in anion-
clinical diagnostics, environmental analysis, and industrial pro- sensitive SCISEs. Therefore, the lower number of anion-selective
cess monitoring [1–5]. Conventional ISEs with internal filling so- SCISEs can be related mainly to the lack of ionophores that are
lution have limitations as they have to be used in the verti- highly selective to anions.
cal position, need regular refilling, and are not easily minia- ISEs are generally used in the potentiometric transduction
turized. This has resulted in extensive research on solid-contact regime where the potential is related to the ion activity of pri-
ion-selective electrodes (SCISEs) over the last three decades [1]. mary ion in solution, according to the Nernst equation [1]. In ad-
Among many other materials, the conducting polymer poly(3,4- dition to potentiometry, several other electrochemical approaches
ethylenedioxythiophene) (PEDOT) has often been used as a solid have been explored for ISEs. Various electrochemical interroga-
contact (ion-to-electron transducer) in SCISEs [1,6–10]. The redox tion methods for ISEs have been introduced and developed by
capacitance related to the reversible oxidation of PEDOT has been Bakker’s group [9,22,23]. Amemiya et.al [24] was able to determine
utilized to improve the potential stability [6,11,12] and to control nanomolar concentrations of perchlorate in drinking water by us-
the standard potential [13,14] of SCISEs. The morphology and ca- ing ion-transfer stripping voltammetry. Recently we introduced a
pacitance of PEDOT can, however, be influenced by the doping coulometric transduction method for SCISEs, which is the topic of
counterion during electropolymerization [7,11,15,16]. For example, this work [25]. In this method, the potential between the SCISE
PEDOT films doped with negatively charged MWCNTs were found and the reference electrode is held constant and the current be-
to exhibit higher redox capacitance than PEDOT doped with chlo- tween the SCISE and a counter electrode is measured. A change
of the ion activity in solution gives rise to oxidation/reduction of
∗ the solid contact (PEDOT) and a transient current is measured and
Corresponding author.
E-mail address: johan.bobacka@abo.fi (J. Bobacka). integrated to obtain the charge which is the analytical measured

https://doi.org/10.1016/j.electacta.2020.137566
0013-4686/© 2020 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/)
T. Han, U. Mattinen, Z. Mousavi et al. Electrochimica Acta 367 (2021) 137566

signal [25]. The cumulated charge can be amplified by increasing 0.014 mA (0.2 mA/cm2 ) for 71 s and 710 s corresponding to a total
the thickness (capacitance) of the solid contact [26, 27]. In order polymerization charge of 1 and 10 mC, respectively. Additionally,
to decrease the response time, the resistance of the ion-selective PEDOT(Cl) and PEDOT(NO3 ) were electrodeposited by applying the
membrane (ISM) was lowered by enlarging the geometric surface same current for 355 s to obtain 5 mC polymerization charge.
area [28] and by using spin-coated thin-layer ISMs [29]. Theoretical The resulting GC/PEDOT(Cl), GC/PEDOT(NO3 ), GC/PEDOT(SO4 )
modeling of the coulometric response based on chloride-sensitive and GC/PEDOT(ClO4 ) electrodes were dried overnight in air at
SCISEs and K+ -selective SCISEs was presented [29, 30]. Further- room temperature. After drying, GC/PEDOT(Cl) and GC/PEDOT(NO3 )
more, coulometric SCISEs were investigated in detecting small pH electrodes were covered with a drop-cast thick-layer or a spin-
changes in seawater [26] and K+ in serum [26,30]. The results ob- coated thin-layer ion-selective membrane (ISM). Three replicates
tained so far indicate that the coulometric transduction method of each anion-sensitive SCISE were prepared for all the experi-
can significantly improve the sensitivity of SCISEs. The coulomet- ments. The spin-coated thin-layer ISM was applied by adding 3
ric response mode was recently shown to work also with SCISEs drops of the membrane cocktail onto the rotating GC electrode
for the detection of the divalent cations Ca2+ and Pb2+ [31]. surface with a rotation speed of 1500 rpm. The volume of each
In this work, the general feasibility of the new coulometric drop was ca 15 μl but most of the volume was swept away from
transduction method is explored for anion sensing. PEDOT doped the rotating electrode surface leaving only a thin-layer of ISM. The
with chloride, nitrate, sulfate and perchlorate are studied with anion-sensitive SCISEs with drop-cast thick membranes were pre-
cyclic voltammetry (CV). Solid-contact anion-sensitive electrodes pared by adding the corresponding cocktail in a 50 μl aliquot cov-
(anion-sensitive SCISEs) are prepared utilizing either PEDOT(Cl) ering the whole electrode disk surface (GC electrode and the PVC
or PEDOT(NO3 ) as solid contact that are covered with drop- body). The composition of the cocktail in %(w/w) was 15% TDMACl,
cast (thick) or spin-coated (thin) PVC-based ion-selective mem- 51% o-NPOE and 34% PVC, dissolved in THF (dry mass = 15%).
branes. Nitrate, perchlorate and sulfate-sensitive SCISEs with vari- The selectivity of this type of anion-sensitive membrane follows
able thickness of PEDOT(Cl) are investigated by chronoamperom- the Hofmeister series [32-34]. The anion-sensitive SCISEs covered
etry and coulometry. Additionally, nitrate-sensitive SCISEs with with drop-cast thick or spin-coated thin ISMs were dried overnight
PEDOT(NO3 ) as solid contact are studied, while sulfate-sensitive followed by an overnight conditioning in 0.1 M NaNO3 , KClO4 or
SCISEs are also characterized by electrochemical impedance spec- Na2 SO4 solution before the measurements. All the electrodes were
troscopy (EIS). conditioned in their corresponding solutions between different ex-
periments.
2. Experimental part
2.3. Potentiometric and chronoamperometric measurements
2.1. Materials
Prior to chronoamperometric measurement, all the SCISEs were
Tridodecylmethylammonium chloride (TDMACl), 3,4- checked using the potentiometric method to verify the proper
ethylenedioxythiophene (EDOT, 97%), sodium sulfate, sodium functioning of the electrodes. Both potentiometric and chronoam-
nitrate, potassium perchlorate, sodium chloride and potassium perometric measurements were done following the protocol de-
chloride were purchased from Sigma-Aldrich. 2-nitrophenyl octyl scribed in our previous work [28]. The nitrate and sulfate-
ether (o-NPOE), poly(vinyl chloride) (PVC) of high molecular sensitive SCISEs were studied in NaNO3 and Na2 SO4 solutions, re-
weight and tetrahydrofuran (THF, >99.5%) were purchased from spectively, without any background electrolyte. The perchlorate-
Fluka. De-ionized water (ELGA Purelab Ultra, resistivity 18.2 sensitive SCISE was studied in KClO4 solutions containing 0.1 M
MΩcm) was used for preparing all the aqueous solutions. NaCl as constant ionic background. The 0.1 M NaNO3 , 0.01 M
Na2 SO4 , and 0.1 M KClO4 starting solutions were sequentially
2.2. Electrode preparation (stepwise) diluted either with deionized water (in the case of ni-
trate and sulfate-sensitive SCISEs) or with 0.1 M NaCl (in the case
Glassy carbon (GC) disk electrodes (SIGRADUR R
G, Germany) of perchlorate-sensitive SCISEs) resulting in log aNO− or log
3
with 3 mm diameter (made by inserting a GC rod in a PVC cylin- aClO− or log aSO2− = 0.18 decades/dilution step. A Metrohm dou-
4 4
der) were polished by using abrasive papers of different coarse- ble junction Ag/AgCl/3 M KCl//1 M KCl was used as reference elec-
ness, diamond pastes with particle diameters of 15 μm, 9 μm, trode for nitrate and perchlorate-sensitive SCISEs and Ag/AgCl/3 M
3 μm and 1 μm, respectively, and finally with 0.3 μm Al2 O3 paste. KCl//0.1 M Na2 SO4 was used as reference electrode for sulfate-
Cleaning of the GC electrode surface was completed by immersing sensitive SCISEs in all potentiometric and chronoamperometric
them in 1.0 M HNO3 for 5 min, rinsing with deionized water, ultra- measurements.
sonication in ethanol for 5 min and finally ultrasonication in deion-
ized water for 5 min. The solution for polymerization of PEDOT 2.4. Chronoamperometry using applied potential steps
solid contacts (PEDOT(Cl), PEDOT(NO3 ), PEDOT(SO4 ), PEDOT(ClO4 ))
contained 0.01 M EDOT and 0.1 M of KCl, NaNO3 , Na2 SO4 and For comparison purposes, the chronoamperometric response of
KClO4 , respectively. The polymerization solution was kept under sulfate-sensitive SCISEs with 5 mC PEDOT(Cl) as solid contact and
stirring overnight to ensure complete dissolution of the monomer. spin-coated thin-layer membrane were also studied by applying
Before polymerization, the solution was purged with N2 gas for potential steps (10 mV/step) in 0.1 M Na2 SO4 using a conventional
10 min to remove oxygen from the solution. PEDOT was electro- three-electrode cell, which was connected to the Autolab instru-
chemically deposited on the electrode surface using an Autolab ment described earlier. The applied potential for each step was
general purpose electrochemical system (PGSTAT20, FRA2, AUTO- 10 mV and ranged from 0.1 V to 0.21 V (oxidation of PEDOT) and
LAB, Eco Chemie, B.V., The Netherlands), connected to a conven- in the reverse direction from 0.21 V to 0.1 V (reduction of PEDOT).
tional three electrode cell. Each GC electrode was connected as The time interval between each potential step was 60 s.
working electrode, and a GC rod was used as counter electrode.
A Metrohm double junction Ag/AgCl/3 M KCl//1 M KCl was used 2.5. Electrochemical impedance spectroscopy
as the reference electrode. PEDOT doped with various counterions,
i.e. chloride, nitrate, sulfate and perchlorate ions, were electrode- Electrochemical impedance spectroscopy measurements were
posited galvanostatically on GC by applying a constant current of performed for sulfate-sensitive SCISEs in 0.1 M Na2 SO4 at the open

2
T. Han, U. Mattinen, Z. Mousavi et al. Electrochimica Acta 367 (2021) 137566

Fig. 1. Cyclic voltammograms of 1 mC and 10 mC of GC/PEDOT(Cl) (a, e), GC/PEDOT(SO4 ) (b, f), GC/PEDOT(NO3 ) (c, g) and GC/PEDOT(ClO4 ) (d, h) films measured in 0.1 M
NaNO3 (black), 0.1 M KClO4 (red), 0.1 M Na2 SO4 (blue), 0.1 M KCl (green) with 50 mV/s scan rate and a potential window from −0.2 V to 0.5 V. (For interpretation of the
references to colour in this figure legend, the reader is referred to the web version of this article.)

circuit potential (OCP) using the same Autolab (with FRA) instru- For example, a decrease of the nitrate activity in the solution
ment described above and a conventional three electrode cell. A will induce a positive potential change at the ISM | solution inter-
Metrohm single junction Ag/AgCl/3 M KCl was used as reference face, which in turn gives a negative current transient that shifts the
electrode and a glassy carbon rod as the counter electrode. The fre- equilibrium in Eq. (1) to the right. As a consequence of the poten-
quency range was 100 kHz – 10 mHz and the excitation amplitude tiometric response (Nernst equation) of SCISEs, the obtained cu-
was 10 mV (RMS) around OCP. mulated charge of anion-sensitive and cation-selective SCISEs has
opposite signs, but otherwise the operation principle is analogous.
2.6. Cyclic voltammetry
3.1. Cyclic voltammetry of PEDOT

Cyclic voltammetry (CV) for GC electrodes covered with 1


The cyclic voltammograms of GC/PEDOT(Cl), GC/PEDOT(NO3 ),
mC and 10 mC of PEDOT(Cl), PEDOT(NO3 ), PEDOT(SO4 ) and
GC/PEDOT(SO4 ), and GC/PEDOT(ClO4 ) in various electrolyte solu-
PEDOT(ClO4 ) solid contact were carried out in 0.1 M KCl, NaNO3 ,
tions are shown in Fig. 1. In the potential range of −0.2 V to
Na2 SO4 and KClO4 solutions. In these measurements, the Autolab
0.5 V, the cyclic voltammograms of PEDOT doped with various
general purpose electrochemical system connected to a conven-
counterions show no apparent oxidation or reduction peaks but
tional one-compartment three-electrode electrochemical cell was
a capacitive-like behavior typical for PEDOT-based films in aque-
used. A Metrohm single junction Ag/AgCl/3 M KCl was used as ref-
ous electrolytes [7]. In the coulometric measurement mode, the
erence electrode and a glassy carbon rod as the counter electrode.
capacitive-like current of the PEDOT solid contact is of importance
The cyclic voltammograms were recorded in the potential range of
since the analytical signal is the cumulated charge deriving from
−0.2 V to +0.5 V with a scan rate of 50 mV/s.
the redox reaction of the PEDOT solid contact.
The voltammetric response of 1 mC GC/PEDOT electrodes cy-
3. Results and discussions cled in various electrolyte solutions are compared in Fig. 1a-d.
The cyclic voltammograms of GC/PEDOT(Cl), GC/PEDOT(SO4 ) and
In the coulometric transduction method for SCISEs, the poten- GC/PEDOT(NO3 ) were relatively independent of the supporting
tial between the SCISE and the reference electrode is held con- electrolyte (KCl, Na2 SO4 , NaNO3 , KClO4 ) used during CV (Fig. 1a-
stant and the current between the SCISE and the counter elec- c). However, the cyclic voltammograms of GC/PEDOT(ClO4 ) showed
trode is measured [25-28]. A change of the ion activity in solu- slightly higher currents and also some dependence on the support-
tion gives rise to oxidation/reduction of the solid contact (PEDOT) ing electrolyte used during CV (Fig. 1d).
and a transient current is measured and integrated to obtain the As the loading of the PEDOT film was increased to 10 mC
charge which is the analytical signal. The working principle of the (Fig. 1e-h), the differences in CV became more pronounced for
constant potential coulometric method has been clearly and thor- the studied electrolytes. As shown in Fig. 1e-h, the capacitive-like
oughly described in our previous works for cation-selective SCISEs current response of 10 mC GC/PEDOT electrodes in various elec-
[25–28]. trolyte solutions is clearly dependent on the anion present during
For an anion-sensitive SCISE, the oxidation/reduction of the electropolymerization. The capacitive currents of 10 mC GC/PEDOT
PEDOT-based solid contact will ideally follow the reaction shown electrodes tends to increase in the sequence: GC/PEDOT(Cl) ≈
in Eq. (1) where a nitrate-sensitive SCISE with a PEDOT(NO3 ) solid GC/PEDOT(SO4 ) < GC/PEDOT(NO3 ) < GC/PEDOT(ClO4 ). The differ-
contact is used as an example. ences in current can be related to different amounts of PEDOT or
different electroactivity of PEDOT at a given polymerization charge,
PEDOTn+ (NO3 − )n + xe− ↔ PEDOT(n−x)+ (NO3 − )n−x + xNO3 − depending on the doping anions used in the electropolymerization.
(1) This implies that for a given polymerization charge, a higher yield

3
T. Han, U. Mattinen, Z. Mousavi et al. Electrochimica Acta 367 (2021) 137566

Fig. 2. Electrochemical impedance spectra recorded for sulfate-sensitive SCISEs Fig. 3. Low-frequency capacitance (CL ) values for the sulfate-sensitive SCISEs (taken
with 1, 5, and 10 mC PEDOT(Cl) solid contact covered with drop-cast thick-layer ISM from Table 1) vs polymerization charge of the PEDOT(Cl) solid contact for spin-
(a) and spin-coated thin-layer ISM (b). The supporting electrolyte is 0.1 M Na2 SO4 . coated and drop-cast ISM. For 1 and 5 mC PEDOT(Cl) the CL values for spin-coated
The measurement is done at open circuit potential with 10 mV (RMS) perturbation and drop-cast ISM overlap.
amplitude and 100 kHz – 10 mHz frequency range. (For interpretation of the refer-
ences to colour in this figure legend, the reader is referred to the web version of
this article.)
drop-cast thick-layer and spin-coated thin-layer ISM are roughly
Table 1 the same, which indicates that the oxidation/reduction (dop-
Low-frequency capacitance (CL ) values for sulfate-sensitive SCISEs with ing/undoping) process of PEDOT is not limited by ion transport
PEDOT(Cl) solid contact calculated from the impedance spectra in Fig. 2. through the ISM. As clearly shown in Fig. 3, the redox capacitance
Polymerization charge of PEDOT(Cl) / mC ISM CL / μF (CL ) of PEDOT in the SCISEs increases linearly proportional to the
polymerization charge (thickness) of the PEDOT(Cl) solid contact in
1 Spin-coated 16
1 Drop-cast 15 a predictable manner.
5 Spin-coated 114
5 Drop-cast 116
10 Spin-coated 223 3.3. Coulometric response of nitrate-sensitive SCISEs
10 Drop-cast 245

The chronoamperometric and coulometric response recorded


for the nitrate-sensitive SCISEs with 1, 5, and 10 mC PEDOT(NO3 )
of electroactive PEDOT(ClO4 ) and PEDOT(NO3 ) is obtained, in com- as solid contact covered with drop-cast and spin-coated ISM is
parison to PEDOT(Cl) and PEDOT(SO4 ). shown in Fig. 4. The starting solution is 0.1 M NaNO3 that was
The results in Fig. 1e-h show also that the supporting elec- stepwise diluted with water. As shown in Fig. 4, the electrode with
trolyte used during CV plays a role on the resulting current re- larger polymerization charge of PEDOT gives a larger cumulated
sponse, especially for GC/PEDOT(ClO4 ) and GC/PEDOT(NO3 ). For the charge (Q) and requires longer time to reach equilibrium after each
electropolymerization of PEDOT, the doping counterions can re- change in the activity of the primary anion (NO3 - ) in the solu-
markably affect the structural feature, morphology and capacitance tion. The signal amplification is proportional to the polymerization
of the resulting polymer film [7,11,15]. This phenomenon may be charge of the PEDOT-based solid contact, in analogy with cation-
related to anion size and hydrophobicity (Hofmeister series). Inter- selective SCISEs [27].
estingly, the voltammograms of GC/PEDOT(Cl) are relatively inde- Nitrate-sensitive SCISEs with PEDOT(NO3 ) as solid contact and
pendent of the supporting electrolyte used during potential cycling spin-coated membrane give higher current peaks due to a lower
(CV). Therefore, GC/PEDOT(Cl) was considered to be a suitable solid resistance of the thin ISM (Fig. 4c), compared to the thick drop-
contact when comparing various anion-sensitive SCISEs. cast ISM (Fig. 4a). These results are in good qualitative agreement
with the results from our previous work using K+ -SCISEs [26].
3.2. Electrochemical impedance spectroscopy The coulometric response of nitrate-sensitive SCISEs with
PEDOT(NO3 ) solid contact and thin spin-coated ISM is very fast
Electrochemical impedance spectroscopy (EIS) measurements and even for the 10 mC PEDOT(NO3 ) solid contact the cumulated
were performed for sulfate-sensitive SCISEs with 1, 5, and 10 mC charge (Q) stabilizes within 2 min (Fig 4c, d). On the contrary,
PEDOT(Cl) as solid contact covered with drop-cast thick-layer or the coulometric response of the nitrate-sensitive SCISE with 10 mC
spin-coated thin-layer membranes. The EIS measurements were PEDOT(NO3 ) solid contact and drop-cast (thick) ISM does not quite
performed at the OCP of the respective SCISE. As shown in Fig. 2, reach equilibrium within 5 min (Fig. 4a, b)
the diameter of the high-frequency semicircle decreases when the Amplification of the cumulated charge Q with increasing thick-
thickness of the anion-sensitive membrane (ISM) is decreased. This ness of the conducting polymer is also confirmed by nitrate-
shows that the high-frequency semicircle represents the bulk re- sensitive SCISEs with PEDOT(Cl) solid contact and spin-coated thin-
sistance (in parallel with the geometric capacitance) of the ISM. layer ISM, as shown in Fig. 5. The cumulated charge Q is linearly
The thickness of the PEDOT solid contact does not influence the proportional to log aNO− in the studied range of −1 to −2.45.
3
impedance at high frequencies, as the resistance of the PEDOT As it was shown in Fig. 1, electrolyte solutions used in the
solid contact film is expected to be negligible compared to that electropolymerization of PEDOT-based solid contact films and in
of the ISM [6]. the cyclic voltammetry measurements affect the current responses.
The low frequency (almost vertical) part of the impedance In Fig. 6, a comparison of the cumulated charge Q vs time of
spectra (10 mHz) is related to the bulk redox capacitance of the nitrate-sensitive SCISEs with 1 mC PEDOT(Cl) or 1 mC PEDOT(NO3 )
PEDOT(Cl) solid contact [7]. The results in Table 1 show that solid contact and drop-cast membranes upon stepwise dilutions is
the redox capacitance (CL ) values of sulfate-sensitive SCISEs with shown. The two replicates of electrodes with PEDOT(Cl) solid con-

4
T. Han, U. Mattinen, Z. Mousavi et al. Electrochimica Acta 367 (2021) 137566

Fig. 4. Chronoamperograms for the first three dilutions (a, c) and the corresponding chronocoulograms (b, d) of nitrate-sensitive SCISEs with 1, 5, and 10 mC PEDOT(NO3 )
covered with drop-cast thick-layer and spin-coated thin-layer ISM. The starting solution is 0.1 M NaNO3 diluted with water to have log aNO−3 = 0.18 decades/dilution step.
(For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Fig. 5. Chronocoulograms (a) and the cumulated charge Q vs log aNO−3 (b) of nitrate-sensitive SCISEs with 1, 5, and 10 mC PEDOT(Cl) covered with spin-coated thin-layer
ISM. The Q values in Fig. 5b correspond to the saturated signals in Fig. 5a. (For interpretation of the references to colour in this figure legend, the reader is referred to the
web version of this article.)

tact show good electrode-to-electrode reproducibility in the coulo- the difference between the coulometric response of the SCISEs
metric measurement mode. with 1 mC PEDOT(NO3 ) and 1 mC PEDOT(Cl) as solid contacts
Despite the equivalent charge of the 1 mC PEDOT solid con- (Fig. 6) is even larger than expected from the CVs recorded in
tacts, the cumulated charge Q vs time curve of the electrode with aqueous solutions (Fig. 1a, c). This suggests that the capacitance of
PEDOT(NO3 ) solid contact is significantly larger than that of the PEDOT depends also on the contacting medium, i.e. aqueous elec-
SCISEs with PEDOT(Cl). A larger cumulated charge Q corresponds trolyte vs ISM.
to a larger redox capacitance of the PEDOT solid contact. However,

5
T. Han, U. Mattinen, Z. Mousavi et al. Electrochimica Acta 367 (2021) 137566

tact and is linearly proportional to log aSO2− in the studied range


4
of −2 to −2.9.
When comparing the coulometric response of the sulfate-
sensitive SCISEs (Fig. 8) and the nitrate-sensitive SCISEs (Fig. 5),
it can be seen that the cumulated charge Q of the SCISEs that
are sensitive to the divalent anion (sulfate) is half of the cumu-
lated charge Q for the monovalent anion (nitrate), as expected from
Nernst equation.
Fig. 9 shows the chronoamperograms recorded for the sulfate-
sensitive SCISEs, based on 5 mC PEDOT(Cl) solid contact with spin-
coated ISM, when 10 mV potential steps are applied in both di-
rections. The current response of the SCISEs upon oxidation with
an increasing potential step (No. 1) and reduction with a decreas-
ing potential step (No. 2) are shown in Fig. 9a, while the corre-
Fig. 6. Chronocoulograms for nitrate-sensitive SCISEs based on 1 mC PEDOT(Cl) and
sponding charge is shown in Fig. 9b. The 10 mV potential step
PEDOT(NO3 ) solid contact covered with drop-cast ISM during sequential dilution
steps. The starting solution (0.1 M NaNO3 ) is diluted with water (log aNO−3 = 0.18 results in ca 1 μC cumulated charge Q as shown in Fig. 9b. This
decades/step). (For interpretation of the references to colour in this figure legend, is approximately twice the ca 0.6 μC charge observed for one di-
the reader is referred to the web version of this article.) lution step (log aSO2− = 0.18 decades/step) that corresponds to
4
ca 5 mV change at the ISM | sample interface. The change of the
applied potential by the potentiostat and by varying the phase
boundary potential between solution and ISM by changing the ac-
3.4. Coulometric response of sulfate-sensitive SCISEs tivity of the primary ion are thus in good agreement. Furthermore,
the results shown in Fig. 9 prove the reversibility of the oxida-
In Fig. 7, the chronoamperometric and coulometric responses of tion/reduction reaction of the PEDOT solid contact. When consid-
sulfate-sensitive SCISEs with spin-coated thin-layer and drop-cast ering that the results shown in Fig. 9 correspond to two individual
thick-layer membranes are compared. The polymerization charge SCISEs (No.1 and No. 2), it can be concluded that the electrode-
of thePEDOT(Cl) solid contact for these electrodes was 1 and 5 to-electrode reproducibility is good. The differences observed be-
mC. Again, the sulfate-sensitive SCISE with spin-coated thin-layer tween No. 1 and No. 2 in Fig 9 include any possible irreproducibil-
ISM results in higher current peaks and faster responses (shorter ity in the electrode preparation, including electropolymerization of
equilibration time) than the drop-cast ISM for both thicknesses of PEDOT(Cl) and spin-coating of the ISM.
the solid contact. These obtained results are in good agreement
with the results described above (Fig. 2, Fig. 4) and results from
our earlier works [26,29]. The chronocoulograms obtained by inte- 3.5. Coulometric response of perchlorate-sensitive SCISEs
gration of the current response is shown as dashed lines in Fig. 7.
For sulfate-sensitive SCISEs with spin-coated membranes and 1 mC A comparison of the chronocoulograms of perchlorate, ni-
PEDOT(Cl) as solid contact, the chronocoulograms stabilize within trate, and sulfate-sensitive SCISEs with 10 mC PEDOT(Cl) solid
less than 1 min after each dilution step (Fig. 7a). contact covered with spin-coated thin-layer ISM during sequen-
The chronoamperograms recorded for sulfate-sensitive SCISEs tial dilution steps is shown in Fig. 10. The starting solutions
with 1, 5, and 10 mC PEDOT(Cl) as solid contact covered with spin- are 0.01 M Na2 SO4 for sulfate-sensitive SCISE and 0.1 M NaNO3
coated thin-layer membrane are shown in Fig. 8. The starting so- for nitrate-sensitive SCISE that were diluted with water. In the
lution (0.01 M Na2 SO4 ) is stepwise diluted with water. It can be case of perchlorate-sensitive SCISEs, the starting solution is 0.1 M
seen from Fig. 8a that the current peaks after each dilution are KClO4 + 0.1 M NaCl as background electrolyte that was diluted
rather reproducible. Also here, thicker PEDOT(Cl) solid contacts re- with 0.1 M NaCl. The coulometric response of perchlorate-sensitive
quire longer equilibration time after each dilution step (Fig. 8a). SCISEs was not influenced by the presence of the interfering ion
The chronocoulograms and the cumulated charge Q vs log SO24− are (Cl− ) in the solution. The perchlorate-SCISE was also observed
presented in Fig. 8b. The cumulated charge Q increases with in- to give a Nernstian potentiometric response to perchlorate, in
creasing polymerization charge (thickness) of PEDOT(Cl) solid con- presence of 0.1 M NaCl, within the concentration range 10−1 to

Fig. 7. Chronoamperograms for 1st and 2nd dilution steps (solid line) and the corresponding chronocoulograms (dash line) for sulfate-sensitive SCISEs with 1 mC (a) and 5
mC (b) PEDOT(Cl) solid contact covered with drop-cast thick-layer and spin-coated thin-layer ISM. The starting solution is 0.01 M Na2 SO4 diluted with water. (For interpre-
tation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

6
T. Han, U. Mattinen, Z. Mousavi et al. Electrochimica Acta 367 (2021) 137566

Fig. 8. Chronoamperograms (a) and chronocoulograms (b) for sulfate-sensitive SCISEs with 1, 5, and 10 mC PEDOT(Cl) solid contact covered with spin-coated thin-layer ISM.
Cumulated charge Q vs log aSO2− is shown in the inset of Fig. 8b. The starting solution 0.01 M Na2 SO4 was diluted with water. (For interpretation of the references to colour
4
in this figure legend, the reader is referred to the web version of this article.)

Fig. 9. Chronoamperograms (a) and chronocoulograms (b) for sulfate-sensitive SCISE with 5 mC PEDOT(Cl) solid contact covered with spin-coated thin-layer ISM, measured
in 0.1 M Na2 SO4 when applying 10 mV potential steps using the potentiostat. The time interval between potential steps is 60 s. (For interpretation of the references to colour
in this figure legend, the reader is referred to the web version of this article.)

− −
series: ClO4 > SCN− > I− > NO− − −
3 > Br > Cl ≈ HCO3 > H2 PO4 >
2−
2−
SO4 [32-34]. The selectivity coefficient (log KPot ClO4,Cl
) for the
perchlorate-sensitive SCISE used here is expected to be in the
range −3 to −4 [4].
The results presented in Fig. 10 for nitrate, perchlorate, and
sulfate-sensitive SCISEs are logical and follow the Nernst equa-
tion as the coulometric response for the monovalent perchlorate-
sensitive SCISE is roughly similar to nitrate-sensitive SCISE and ap-
proximately double of that of the divalent sulfate-sensitive SCISE.

Conclusions

In this work, the electrochemical characteristics of solid-contact


anion-sensitive electrodes based on PEDOT as solid contact and
a plasticized PVC-based anion sensitive membrane were investi-
Fig. 10. Chronocoulograms for perchlorate, nitrate and sulfate-sensitive SCISEs with gated. Results obtained by cyclic voltammetry of GC/PEDOT elec-
10 mC PEDOT(Cl) solid contact covered with spin-coated ISM during sequential
trodes showed that the yield of electroactive PEDOT depended on
dilution steps. The starting solutions for sulfate and nitrate-sensitive SCISEs are
0.01 M Na2 SO4 and 0.1 M NaNO3 , respectively (diluted with water). The starting so- the supporting electrolyte used during electrodeposition as well
lution for the perchlorate-sensitive SCISE is 0.1 M KClO4 + 0.1 M NaCl background as the electrolyte used during the subsequent potential cycling of
electrolyte diluted with 0.1 M NaCl. (For interpretation of the references to colour the GC/PEDOT electrodes. The different yield of PEDOT for a given
in this figure legend, the reader is referred to the web version of this article.)
polymerization charge was reflected in the coulometric response
of the solid-contact anion-sensitive electrodes where PEDOT was
10−3.4 . This can be explained by the fact that the selectivity of used as the solid contact (GC/PEDOT/ISM). The coulometric re-
the TDMACl ion-exchanger-based ISM towards different anions cor- sponse of GC/PEDOT/ISM was shown to be faster for spin-coated
relates with their relative lipophilicity, following the Hofmeister (thin) ISM than for drop-cast (thick) ISM. Furthermore, the am-

7
T. Han, U. Mattinen, Z. Mousavi et al. Electrochimica Acta 367 (2021) 137566

plification of the coulometric signal was proportional to the poly- [8] J. Bobacka, A. Ivaska, A. Lewenstam, Potentiometric ion sensors based on con-
merization charge (thickness) of PEDOT-based solid contact and ducting polymers, Electroanalysis 15 (2003) 366–374, https://doi.org/10.1002/
elan.20 0390 042.
inversely proportional to the charge of the anion. When apply- [9] G.A. Crespo, M. Cuartero, E. Bakker, Thin layer ionophore-based membrane for
ing a potential step to the GC/PEDOT/ISM electrode, the cumulated multianalyte ion activity detection, Anal. Chem. 87 (2015) 7729–7737, https:
charge was reversible and of similar magnitude to the charge ob- //doi.org/10.1021/acs.analchem.5b01459.
[10] E. Bakker, Electroanalysis with membrane electrodes and liquid-liquid inter-
tained for an activity change in solution causing a similar change faces, Anal. Chem. 88 (2016) 395–413, https://doi.org/10.1021/acs.analchem.
of potential. Results obtained by electrochemical impedance spec- 5b04034.
troscopy (EIS) for GC/PEDOT/ISM clearly showed that the high- [11] Z. Mousavi, J. Bobacka, A. Lewenstam, A. Ivaska, Poly(3,4-
ethylenedioxythiophene) (PEDOT) doped with carbon nanotubes as
frequency semicircle originated from the ISM and was independent
ion-to-electron transducer in polymer membrane-based potassium
of the PEDOT thickness. The thickness of the ISM also did not influ- ion-selective electrodes, J. Electroanal. Chem. 633 (2009) 246–252,
ence the low-frequency capacitive part of the impedance spectra, https://doi.org/10.1016/j.jelechem.20 09.06.0 05.
[12] J. Bobacka, Conducting polymer-based solid-state ion-selective electrodes, Elec-
which was related to the redox capacitance of the PEDOT film. The
troanalysis 18 (2006) 7–18, https://doi.org/10.10 02/elan.20 0503384.
low-frequency capacitance value increased linearly with increas- [13] U. Vanamo, J. Bobacka, Electrochemical control of the standard potential of
ing thickness of PEDOT solid contact. The coulometric results were solid-contact ion-selective electrodes having a conducting polymer as ion-to-
thus in line with experimental results obtained by EIS and followed electron transducer, Electrochim. Acta 122 (2014) 316–321, https://doi.org/10.
1016/j.electacta.2013.10.134.
the theoretical predictions of the Nernst equation. Altogether, the [14] U. Vanamo, J. Bobacka, Instrument-free control of the standard potential of po-
results presented in this work showed that the coulometric trans- tentiometric solid-contact ion-selective electrodes by short-circuiting with a
duction method is appropriate and applicable for anion-sensitive conventional reference electrode, Anal. Chem. 86 (2014) 10540–10545, https:
//doi.org/10.1021/ac501464s.
solid-contact electrodes. [15] X. Wang, P. Sjöberg-Eerola, J.E. Eriksson, J. Bobacka, M. Bergelin, The effect
of counter ions and substrate material on the growth and morphology of
poly(3,4-ethylenedioxythiophene) films: towards the application of enzyme
Credit author statement electrode construction in biofuel cells, Synth. Met. 160 (2010) 1373–1381,
https://doi.org/10.1016/j.synthmet.2010.01.033.
Tingting Han: Methodology, Investigation, Data analysis, Writ- [16] S.A. Spanninga, D.C. Martin, Z. Chen, X-ray photoelectron spectroscopy study
of counterion incorporation in poly(3,4-ethylenedioxythiophene) (PEDOT) 2:
ing - Original Draft. Ulriika Mattinen: Data analysis, Writing - Re- polyanion effect, toluenesulfonate, and small anions, J. Phys. Chem. C. 114
view & Editing, Supervision. Zekra Mousavi: Data analysis, Writing (2010) 14992–14997, https://doi.org/10.1021/jp104591d.
- Review & Editing, Supervision. Johan Bobacka: Conceptualiza- [17] P. Sjöberg-Eerola, J. Bobacka, A. Lewenstam, A. Ivaska, All-solid-state chlo-
ride sensors based on electronically conducting, semiconducting and insulat-
tion, Methodology, Writing - Review & Editing, Supervision, Project
ing polymer membranes, Sens. Actuat. B Chem. 127 (2007) 545–553, https:
administration, Funding acquisition. //doi.org/10.1016/j.snb.20 07.05.0 04.
[18] T.A. Bendikov, T.C. Harmon, Long-lived solid state perchlorate ion selective
sensor based on doped poly(3,4-ethylenedioxythiophene) (PEDOT) films, Anal.
Declaration of Competing Interest Chim. Acta. 551 (2005) 30–36, https://doi.org/10.1016/j.aca.2005.07.004.
[19] Z. Jarolímová, G.A. Crespo, X. Xie, M. Ghahraman Afshar, M. Pawlak, E. Bakker,
The authors declare that they have no known competing finan- Chronopotentiometric carbonate detection with all-solid-state ionophore-
based electrodes, Anal. Chem. 86 (2014) 6307–6314, https://doi.org/10.1021/
cial interests or personal relationships that could have appeared to ac5004163.
influence the work reported in this paper. [20] K. Martin, S.A. Kadam, U. Mattinen, J. Bobacka, I. Leito, Solid-contact acetate-
selective electrode based on a 1,3-bis(carbazolyl)urea-ionophore, Electroanaly-
sis 31 (2019) 1061–1066, https://doi.org/10.10 02/elan.20180 0790.
Acknowledgements [21] A. Michalska, A. Galuszkiewicz, M. Ogonowska, M. Ocypa, K. Maksymiuk,
PEDOT films: multifunctional membranes for electrochemical ion sens-
ing, J. Solid State Electrochem. 8 (2004) 381–389, https://doi.org/10.1007/
Financial support by the Academy of Finland (Project No. s10 0 08-0 03-0459-8.
317829) is gratefully acknowledged. This work is part of the activ- [22] M. Cuartero, G.A. Crespo, E. Bakker, Ionophore-based voltammetric ion activity
ities of the Johan Gadolin Process Chemistry Centre (PCC) at Åbo sensing with thin layer membranes, Anal. Chem. 88 (2016) 1654–1660, https:
//doi.org/10.1021/acs.analchem.5b03611.
Akademi University. [23] S. Jansod, M. Cuartero, T. Cherubini, E. Bakker, Colorimetric readout for poten-
tiometric sensors with closed bipolar electrodes, Anal. Chem. 90 (2018) 6376–
6379, https://doi.org/10.1021/acs.analchem.8b01585.
Supplementary materials
[24] Y. Kim, S. Amemiya, Stripping analysis of nanomolar perchlorate in drink-
ing water with a voltammetric ion-selective electrode based on thin-layer
Supplementary material associated with this article can be liquid membrane, Anal. Chem. 80 (2008) 6056–6065, https://doi.org/10.1021/
ac8008687.
found, in the online version, at doi:10.1016/j.electacta.2020.137566.
[25] E. Hupa, U. Vanamo, J. Bobacka, Novel ion-to-electron transduction principle
for solid-contact ISEs, Electroanalysis 27 (2015) 591–594, https://doi.org/10.
References 10 02/elan.20140 0596.
[26] T. Han, U. Mattinen, J. Bobacka, Improving the sensitivity of solid-contact
[1] J. Bobacka, A. Ivaska, A. Lewenstam, Potentiometric ion sensors, Chem. Rev. 108 ion-selective electrodes by using coulometric signal transduction, ACS Sens. 4
(2008) 329–351, https://doi.org/10.1021/cr068100w. (2019) 900–906, https://doi.org/10.1021/acssensors.8b01649.
[2] A. Lewenstam, Routines and challenges in clinical application of electrochem- [27] U. Vanamo, E. Hupa, V. Yrjänä, J. Bobacka, New signal readout principle for
ical ion-sensors, Electroanalysis 26 (2014) 1171–1181, https://doi.org/10.1002/ solid-contact ion-selective electrodes, Anal. Chem. 88 (2016) 4369–4374, https:
elan.20140 0 061. //doi.org/10.1021/acs.analchem.5b04800.
[3] E. Bakker, P. Bühlmann, E. Pretsch, Carrier-based ion-selective electrodes and [28] T. Han, U. Vanamo, J. Bobacka, Influence of electrode geometry on the re-
bulk optodes. 1. General characteristics, Chem. Rev. 97 (1997) 3083–3132, sponse of solid-contact ion-selective electrodes when utilizing a new coulo-
https://doi.org/10.1021/cr940394a. metric signal readout method, ChemElectroChem 3 (2016) 2071–2077, https:
[4] P. Bühlmann, E. Pretsch, E. Bakker, Carrier-based ion-selective electrodes and //doi.org/10.10 02/celc.20160 0575.
bulk optodes. 2. Ionophores for potentiometric and optical sensors, Chem. Rev. [29] Z. Jarolímová, T. Han, U. Mattinen, J. Bobacka, E. Bakker, Capacitive model for
29 (1998) 1593–1688, https://doi.org/10.1021/cr970113+. coulometric readout of ion-selective electrodes, Anal. Chem. 90 (2018) 8700–
[5] C. Zuliani, D. Diamond, Opportunities and challenges of using ion-selective 8707, https://doi.org/10.1021/acs.analchem.8b02145.
electrodes in environmental monitoring and wearable sensors, Electrochim. [30] Y.O. Kondratyeva, E.G. Tolstopjatova, D.O. Kirsanov, K.N. Mikhelson, Chronoam-
Acta 84 (2012) 29–34, https://doi.org/10.1016/j.electacta.2012.04.147. perometric and coulometric analysis with ionophore-based ion-selective elec-
[6] J. Bobacka, Potential stability of all-solid-state ion-selective electrodes using trodes: a modified theory and the potassium ion assay in serum samples, Sens.
conducting polymers as ion-to-electron transducers, Anal. Chem. 71 (1999) Actuat. B Chem. 310 (2020) 127894, https://doi.org/10.1016/j.snb.2020.127894.
4932–4937, https://doi.org/10.1021/ac990497z. [31] T. Han, Z. Mousavi, U. Mattinen, J. Bobacka, Coulometric response char-
[7] J. Bobacka, A. Lewenstam, A. Ivaska, Electrochemical impedance spectroscopy acteristics of solid contact ion-selective electrodes for divalent cations,
of oxidized poly(3,4-ethylenedioxythiophene) film electrodes in aqueous J. Solid State Electrochem. 24 (2020) 2975–2983, https://doi.org/10.1007/
solutions, J. Electroanal. Chem. 489 (20 0 0) 17–27, https://doi.org/10.1016/ s10 0 08- 020- 04718- 8.
S0 022-0728(0 0)0 0206-0.

8
T. Han, U. Mattinen, Z. Mousavi et al. Electrochimica Acta 367 (2021) 137566

[32] W.E. Morf, The Principles of Ion-Selective Electrodes and of Membrane Trans- [34] K.N. Mikhelson, Ion-Selective Electrodes, Springer, Heidelberg, 2013 https://doi.
port, Akademiai Kiado, Budapest, 1981. org/10.1007/978- 3- 642- 36886- 8_2.
[33] A.L. Grekovich, K.N. Mikhelson, An anomalous behavior of anion-exchange
membranes with low concentration of quaternary ammonium sites: an ap-
parent selectivity to bicarbonate and phosphate, and its true nature, Electro-
analysis 14 (2002) 1391–1396, https://doi.org/10.1002/1521-4109(200211)14:
19/201391::AID-ELAN13913.0.CO;2-S.

You might also like