Вы находитесь на странице: 1из 9

Sensors and Actuators B 177 (2013) 659667

Contents lists available at SciVerse ScienceDirect

Sensors and Actuators B: Chemical


journal homepage: www.elsevier.com/locate/snb

Review

Anti-fouling PEDOT:PSS modication on glassy carbon electrodes for continuous


monitoring of tricresyl phosphate
Xiaoyun Yang a , Jeffrey Kirsch a , Eric V. Olsen b , Jeffrey W. Fergus a , Aleksandr L. Simonian a,
a
b

Materials Research and Education Center, Auburn University, Auburn, AL 36849, United States
Clinical Research Facility, 81st Medical Group, Keesler AFB, MS 39534, United States

a r t i c l e

i n f o

Article history:
Received 27 June 2012
Received in revised form
15 November 2012
Accepted 17 November 2012
Available online xxx
Keywords:
PEDOT:PSS
Electrode fouling
Anti-fouling
Tricresyl phosphate

a b s t r a c t
In this paper the application of poly(3,4-ethlenedioxythiophene)-poly(styrene sulphonate) (PEDOT:PSS)
as an anti-fouling modication on a glassy carbon electrode has been investigated for use in an electrochemical sensing system for continuous monitoring of gaseous tricresyl phosphate (TCP). PEDOT is a type
of conductive polymer with high stability in aqueous solution. The amphiphilic nature of poly(sodium-4styrenesulfonate) (NaPSS) helps repel the oxidation products of cresol and reduces electrode fouling. The
composite modication has high reproducibility which enables the quantitative determination of gaseous
TCP. Although the linear range of detection (50300 ppb) is narrower than that previously reported, this
modied electrode enables continuous monitoring of TCP without the need for electrode polishing that
may limit practical application of the sensor in the aircraft cabin.
2012 Elsevier B.V. All rights reserved.

Contents
1.
2.

3.

4.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Materials and methods. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.
Reagents, solutions, and instruments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.
Electrodes and electrochemical sensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Electropolymerization of PEDOT:PSS on glassy carbon electrode . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.
2.4.
Automatic sampling system for hydrolysis of TCP samples and TCP in engine oils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Diffusion controlled cresol oxidation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1.
3.2.
Microscopy images of electrode fouling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3.
Controlled PEDOT:PSS modication on glassy carbon electrode . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Detection of cresol with modied electrodes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.4.
3.5.
Detection of TCP with modied electrodes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.6.
Detection of hydrolysates from engine oils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Biographies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1. Introduction
Increasing reports have been made concerning the illness of
ight passengers and crew members due to suspicious exposure

Corresponding author. Tel.: +1 334 844 4485; fax: +1 334 844 3400.
E-mail addresses: als@eng.auburn.edu, simonal@auburn.edu (A.L. Simonian).
0925-4005/$ see front matter 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.snb.2012.11.057

659
660
660
660
660
661
661
661
663
663
664
664
665
666
666
666
667

to tricresyl phosphate (TCP), one of common additives in jet


engine oils [1]. As an organophosphate, TCP consists of phosphate
and cresol. It is a known neurotoxin affecting human health by
inhibiting vital enzymes, such as acetylcholine esterase [2] and
carboxyesterases [3], and causing the organophosphorous induced
delayed neuropathy (OPIDN) [46]. Despite the toxicity of TCP, it is
used in many industrial applications, such as ame retardants and
plasticizers [7,8]. Organophosphate containing hydraulic uids

660

X. Yang et al. / Sensors and Actuators B 177 (2013) 659667

have been applied since the 1970s by the US Air Force [9]. Over
several decades, 3 wt% TCP has been added into commercial jet
engine oils, such as Mobil Jet Oil [10]. TCP could be mixed with the
breathable air entering the airliner cabin during engine oil leakage
[11]. Although the most toxic ortho-isomer of TCP is intentionally
excluded from these uids [12], alternatively used meta- and
para-isomers can have a potentially severe effect on human health
mainly through inhalation of aerosols and dermal adsorption.
The most commonly used methods for detecting TCP include:
gas chromatography and mass spectrometry GCMS [13], high
performance liquid chromatography (HPLC) [14] or thin layer
chromatography (TLC) [15]. Gas chromatography measurement can be based on ame photometric detector (GC/FPD),
nitrogenphosphorus sensitive detector (GC/NPD) [16]. Although
they are able to detect TCP below ppm level and with high selectivity, these devices have several drawbacks, such as large size, high
cost, complexity, and a need for highly trained operators, which
make their use on aircraft impractical. Alternatively, electrochemical sensors have the advantages of low price, high sensitivity, small
size, ease of operation, and rapid response. A portable real-time
electrochemical sensor with linear response for ppb level of TCP
in gas has been recently reported by our group [17,18]. Since TCP
itself is electrochemically inactive, it was converted to electroactive cresol for detection. However, the sensor system suffered
from electrode fouling which limited its repetitive usage in long
term applications. A long lasting electrode is needed for on-site
continuous monitoring, so electrode surfaces must be modied to
minimize or eliminate fouling.
Electrode fouling is a common problem during electrochemical analysis of phenolic compounds, such as cresol, and causes
decay in signal response during repetitive measurements [1927].
The fouling is caused by the formation of a passive polymeric lm
on the electrode surface [23,25,28,29]. Upon anodic oxidation of
cresol, phenoxy radicals form, which then couple to form dimers
or oligomers, and nally, a polymeric lm deposits on electrode
surface [30]. This polymeric lm is tough, thermally stable, and
so chemically inert that little oxidation or hydrolysis will occur in
either acidic or basic media [26]. The lms are characterized by
low permeability [31] and strong adhesion to the electrode, which
blocks the surface, and yields electrode fouling [1922,24]. Various
surface treatments and modications have been used to reduce or
even avoid electrode fouling by preventing polymeric substrates
from absorbing onto the surface of electrode [32,33]. For instance,
a special compound called sodium 3,5-dibromo-4-nitroso benzene
sulfonate (DBNBS) was used as an anti-fouling agent against the formation of the cresol polymeric lm [33]. The DBNBS molecule reacts
with the oxidized radical of cresol to form a compound which does
not adhere to the surface of the electrode and consequently prevents the fouling effect. However, this approach is difcult to realize
in practical applications. Overall, the DBNBS should be present in
the buffer solution in stoichiometric concentration with the analyte of interest, which makes its application for real world samples
problematic.
Having advantages of high electronic conductivity and porosity
[34], conducting polymers have attracted considerable interests in
recent years. Modication of conductive polymer for preventing
electrode fouling has been reported by many researchers [3538].
Poly(3,4-ethylenedioxythiophene) (PEDOT) is one of the widely
used conductive polymers. Patra and Munichandraiah [34], Heras
et al. [39], and Lupu et al. [40] applied PEDOT on electrodes
through electropolymerization for detection of phenolic compounds. The surfactant poly(sodium-4-styrenesulfonate) (NaPSS)
was additionally used during PEDOT electropolymerization to
avoid the problems of [41]: (1) the low solubility of thiophene
structures in water, (2) the oxidation potentials higher than that of
water, and (3) and the water-catalyzed formation of thienyl cation

radicals which can activate concomitant reactions and prevent the


formation of the main polymer [42]. Moreover, the hydrophobic
hydrocarbon residues of PSS exhibit strong afnity for PEDOT, while
the hydrophilic sulphonic groups are oriented toward or even protrude into the solution and may hence induce poor adhesion of
the fouling polymeric lm [39,43]. NaPSS also induces the formation of a permeable and less compact polymer network, and yields
high (ionic) conductivity and permeability to the fouling polymeric
lm. This makes the permeation of phenolic compounds through
the lm and oxidation inside possible, ensuring electroneutrality
[39,43]. In the present work, PEDOT:PSS has been used to modify the commonly used glassy carbon electrode [44,45] to develop
a sensor for continuous monitoring of TCP in gas and has been
investigated.
2. Materials and methods
2.1. Reagents, solutions, and instruments
All aqueous solutions were prepared using de-ionized
Milli-Q water (18 m cm). 3,4-Ethylenedioxythiophene (EDOT)
(SigmaAldrich, MO), poly(sodium-4-styrenesulfonate) (NaPSS)
(SigmaAldrich, MO, MW 70,000), and lithium perchlorate
(LiClO4 ) (SigmaAldrich, MO) were used for electro-synthesis of
PEDOT:PSS [39]. The stock 20 mM tri-p-cresyl phosphate (p-TCP)
(SigmaAldrich, MO) solution was prepared in methanol (BDHVWR) and diluted to obtain TCP samples with concentrations of
interest. All TCP samples were converted to cresol with the aid of
alkaline catalyst. Alkaline catalyst was made from NaOH (Fisher,
NJ) and neutral aluminum oxide (Al2 O3 ) (SigmaAldrich, MO) and
packed in Pipet lter tips (USA Scientic, Inc., 20 L leveled lter
tips). p-Cresol (Acros Organics, NJ, 99+%) was dissolved in 0.2 M
Na2 HPO4 /0.2 M KH2 PO4 /10 mM NaCl buffer (0.4 M phosphate
buffer, pH = 6.67). Na2 HPO4 , KH2 PO4 , and NaCl were obtained from
SigmaAldrich, MO.
2.2. Electrodes and electrochemical sensor
All amperometric experiments were performed with CH 93
Instruments (CH1910B) Bi-Potentiostat. A desktop computer was
used to collect the data. Flow injection analysis (FIA) was carried
out using a unicell electrode set (BASi, IN) and a switch injection unit (Valco Instruments Co. Inc.) with a 50 L sample loading
loop. The ow rate was maintained at 20 mL/min by using a single
syringe pump (KD Scientic, MA). The unicell electrode set consists of one glassy carbon working electrode cell (2 mm ), one
electrode set including the stainless steel counter electrode and
Ag/AgCl reference electrode, and one circular gasket (BASi, IN). A
coating solution was applied on the Ag/AgCl reference electrode
surface before detection as per instruction from the company. The
working electrode was polished with alumina powder (BUEHLER,
IL, 1, 0.3, and 0.05 m in order). 10 mM NaCl was contained in all
solutions to maintain the potential of reference electrode. 0.64 V
vs. Ag/AgCl/10 mM NaCl was applied in amperometry. In batch
mode, a glassy carbon electrode (3 mm ), Pt counter electrode,
and Ag/AgCl/3 M KCl reference electrode (BASi, IN) were used.
2.3. Electropolymerization of PEDOT:PSS on glassy carbon
electrode
The electropolymerization of monomer of EDOT was performed
under amperometric conditions in aqueous solution. The solution
contained 5 mM EDOT (0.7108 g/L), 0.1 M NaPSS (20.62 g/L), and
0.1 M LiClO4 (10.64 g/L). The unicell glassy carbon electrode was
polished with alumina as mentioned in Section 2.2, rinsed and sonicated with de-ionized water in an ultrasonic bath for 5 min, ushed

X. Yang et al. / Sensors and Actuators B 177 (2013) 659667

661

Fig. 1. Schematic representative of TCP sampling system. TCP was heated to 230 C in oil bath and gasied along with N2 bubbling, and entered the alkaline catalyst column
set inside an automatic TCP conversion box. In the electronically controlled TCP conversion box, as shown in the dotted square in above gure, TCP was hydrolyzed for later
detection. The control of the automatic TCP conversion box is described in Ref. [17].

with ethanol and water, and dried with N2 gas. 0.5 mC of charge
was applied on the glassy carbon electrode with 2 mm diameter
in order to obtain the thickness of interest of electropolymerization layer, unless otherwise stated. The potential was maintained at
0.95 V vs. Ag/AgCl/3 M KCl. The color of electrode surface turned yellow after modication. For comparison, different amount of charges
have been applied including 0.3, 1, 1.3, 2, 4, and 20 mC on the same
electrode.
2.4. Automatic sampling system for hydrolysis of TCP samples
and TCP in engine oils
Since p-TCP has a very low saturated vapor pressure at room
temperature (1 104 mmHg 0.01 Pa at 20 C, 100 Pa at 220 C,
1000 Pa at 260 C, [4446]), the TCP samples in gas phase are not
readily available. A system was built in our lab for TCP sampling
and is shown schematically in Fig. 1. 0.5 mL of TCP methanol solutions with concentrations of 10, 20, 30, 50, 100, 200, and 300 M
were prepared from the 20 mM TCP stock solution. N2 was bubbled
at a ow rate of 1.1 L/min through the container with TCP for
5 min at room temperature and then 5 min at 230 C (temperature
controlled with an oil bath) to make sure all TCP was evaporated.
The total N2 volume in 10 min was 11 L 0.5 mol (considered at
room temperature), therefore, TCP samples with 10, 20, 30, 50,
100, 200, and 300 ppb in gas phase were realized. Each sample was
own along with N2 through the alkaline catalyst column where

it was hydrolyzed, following the dotted (red) arrow path in Fig. 1.


The alkaline catalyst column was prepared by packing a 100 mg of
mixture of NaOH and Al2 O3 (1:10 wt) inside. The preparation of the
catalyst is described in more detail in Ref. [17]. In the electronically
controlled TCP conversion box (dotted square in Fig. 1), valves
turned after 10 min of gas ow and the pump worked to push 3 mL
0.4 M phosphate buffer solution to wash out TCP hydrolyzate for
detection at the sensor, following the solid (blue) arrow path in
Fig. 1 (its method of control was mentioned in Ref. [17]).
Engine oil samples were prepared by dissolving the samples into
methanol in 2 L/0.5 mL ratio and with the same procedure as TCP
samples. Oil BP274 does not include TCP, Mobil Jet Oil II includes
13%, and BP2380 includes 15% TCP [47]. BP274 oil samples spiked
with TCP in different concentrations were also prepared, sampled,
and measured.
3. Results and discussion
3.1. Diffusion controlled cresol oxidation
The oxidation process of cresol on a bare glassy carbon working
electrode was studied with cyclic voltammetry (CV). In the CV
curve (Fig. 2), current increased quickly at the beginning in each
potential scan due to the charging current, followed by a at line.
Starting from around 0.4 V, current increased again which indicates
the onset of oxidation of cresol. The oxidation was restricted to

Fig. 2. Cyclic voltammetry results of cresol oxidation on bare glassy carbon electrode. (a) First cycle of cyclic voltammetry of 100 M cresol in 0.4 M phosphate buffer solution
at scan rates varying from 10 to 200 mV/s. scan rates increase along the arrow direction. Range of potential: 0800 mV vs. Ag/AgCl/3 M KCl. (b) Calibration of the relation
between the peak current density and square root of the scan rate. The coefcient of determination R2 indicates the variability from linear t. Glassy carbon working electrode
(3 mm ) was used for all of detections and polished between uses.

662

X. Yang et al. / Sensors and Actuators B 177 (2013) 659667

Fig. 3. Observations of fouling on glassy carbon electrode from SEM and EDS. (a) SEM image of Au sputtering coated bare electrode, (b) SEM image of fouled electrode, and
(c) EDS results of the bare and fouled electrodes.

the vicinity of electrode surface and reduced the concentration


of oxidizable cresol in this area. Once the concentration gradient
existed between the vicinity of electrode surface and the bulk
solution, diffusion occurred which yielded a mass ow of cresol
from bulk to electrode surface. In this batch mode, migration was
reduced to negligible levels by addition of a high concentration of
phosphate ions and convection was avoided by preventing stirring
and vibrations, which made diffusion the only path for mass
transfer. We dene a diffusion layer where linear concentration
gradient exists. The thickness of the diffusion layer increased with
time which would atten the concentration prole. In contrast,
the potential continued to scan positively and decreased the

concentration of oxidizable cresol in the vicinity of electrode


surface, which steepen the concentration prole. The slope of concentration difference was intensied; when the effect of potential
scan dominated rst, so did the diffusion, and thus the current
increased which yields the left shoulder of CV peak. Potential scan
kept going positively and the concentration of oxidizable cresol
in the vicinity decreased approaching zero. By the point when
vincinity concentration dropped relatively slower, since it already
approached zero, the diffusion layer thicken effect dominated
and thus the slope of concentration difference reduced. The
current started to decrease, meanwhile, an IV peak was created
[48].

Fig. 4. PEDOT:PSS modication of glassy carbon electrodes. (a) Amperometric results of 4 electrodes applied with 0.5 mC of charge in EDOT:PSS solution, (be) optical images
of these 4 electrodes after modication with 0.5 mC of charge, (fi) optical images of 4 electrodes after modication with 1, 1.3, 2, and 4 mC of charge, (jl) SEM images of
electrode surface applied with 0.3, 0.5, and 2 mC of charge in EDOT/PSS solution.

X. Yang et al. / Sensors and Actuators B 177 (2013) 659667

663

Fig. 5. Amperometric results of 10 M cresol on bare and modied electrodes. (a) Initial amperometric peaks obtained from successive injections of 10 M cresol in 0.4 M
phosphate buffer. Two experiments have been carried out on bare electrode and ve on modied electrodes, two examples of which are shown here. (b) Calibration of peak
current vs. peaks. Note that the rst peak in each case has been normalized to 100%. Unicell glassy carbon electrodes with 2 mm diameter were used.

The results of CV experiments demonstrating the change of the


IV at different scan rates are shown in Fig. 2. 100 M p-cresol
in 0.4 M phosphate buffer was used as the analyte solution, and
the potential was scanned in the range of 00.8 V vs. Ag/AgCl/3 M
KCl with scan rates of 10, 20, 50, 80, 100, 120, 150, and 200 mV/s.
Following the higher scan rate, the whole IV cycle expanded due
to the increased charging current. The peak current also increased
(Fig. 2a). Fig. 2b shows the calibration of current density vs square
root of scan rate. The linear relation indicates the oxidation process of cresol is limited by the diffusion to electrode surface, while
the electron transfer at electrode/solution interface is relatively
quick [41,49]. The diffusion controlled property of cresol oxidation
ensures that current signal is proportional to the bulk concentration
in CV as well as amperometry mode.
Looking at the reversal scan curve in Fig. 2a, we note the absence
of reduction peak which indicates the process was not reversible.
Cresol was oxidized to radicals, followed by the radical coupling
and the formation of inert dimers or oligomers. A stable polymeric
layer formed, nally, and brought the electrode fouling [30]. Indeed,
while this fouling layer could be removed by physical polishing, it
is inert to chemicals and cannot be oxidized until a potential of up
to 3 V is applied [50].

the reproducibility. To show the ability to control modication, 4 unicell glassy carbon electrodes (2 mm ) were immersed
in EDOT/NaPSS solution and applied 0.5 mC of charge (potential = 0.95 V in single-potential amperometry mode). Fig. 4ae
shows the results of these modications. The amperometry proles were almost the same for these 4 electrodes as shown in Fig. 4a,
and all of these 4 electrodes had the same color after modication

3.2. Microscopy images of electrode fouling


The formation of fouling layer was studied by SEM, EDS, and
AFM. For SEM imaging, the electrodes were sputter coated prior to
analysis. One electrode was left polished, while the other one was
fouled by 10 injections of 10 M cresol in 0.4 M PB buffer in amperometry mode. The polished electrode and fouled electrode were
compared in SEM images (Fig. 3a and b). Only sputter coated Au was
seen on polished electrode, but additional structures were observed
from the surface of fouled electrode. EDS shows the existence of
oxygen on the fouled electrode, which could come from the fouling
products of cresol (Fig. 3c). AFM images also conrm the existence
of fouling layer on electrode surface (Results not shown). The surface prole of fouled electrodes varied in the range of 030 nm
while the variation in the bare electrode was less than 15 nm.
3.3. Controlled PEDOT:PSS modication on glassy carbon
electrode
To reduce the fouling from oxidation of cresol, PEDOT:PSS
modication was applied on glassy carbon electrode via electropolymerization. One important parameter for modication is

Fig. 6. Amperometric results of injections of 0.210 M cresol in 0.4 M phosphate


buffer. (a) Two representative sets of continuously amperometric results. Different
concentrations of cresol samples were injected in random order to avoid potential system error and by this way electrode fouling could be seen more clearly.
(b) Calibration curve of detection of cresol with different concentrations on modied electrode. The coefcient of determination R2 indicates the variation of linear
t. Error bars are marked as bars above and below current symbols. Unicell glassy
carbon electrodes with 2 mm diameter were used.

664

X. Yang et al. / Sensors and Actuators B 177 (2013) 659667

which indicates the uniform and controllable conguration. Additionally, 4 more electrodes were coated with 1, 1.3, 2, and 4 mC
of charge. The color of modied electrode surface changed with
different amounts of charge (Fig. 4fi).
SEM was also carried out to compare the electrode surfaces
modied with 0.3, 0.5, and 2 mC of charge. Images from left to
right correspond to increasingly more charge (Fig. 4jl). During SEM
imaging, the conguration was seen to be uniform over the whole
electrode surface and so representative SEM images in small areas
can be produced. Fig. 4jl illustrate the change in morphology and
particle size of the PEDOT layer. Cracks formed on the PEDOT:PSS
layer prepared by applying 2 mC of charge

3.4. Detection of cresol with modied electrodes


The controlled PEDOT:PSS modication described above was
used to reduce fouling effect in the amperometric detection of
cresol. 10 M cresol in 0.4 M phosphate buffer was injected successively on both modied (left side of Fig. 5a) and un-modied
electrodes (right side of Fig. 5a). The peak currents decayed in
both cases after repeated exposure to cresol, but magnitude of
the decrease was much lower in the modied electrode. For comparison, the peak height was normalized to the rst peak in each
case being normalized to 100%. The successive peaks were seriously decayed on bare electrode but much less fouling occurred

on modied electrode (Fig. 5b). The comparison indicates that for


continuous on-site monitoring of TCP, the bare electrode is not a
good candidate. Although thoroughly polishing the electrode could
remove the fouling layer, this would require trained personnel and
adds complexity to sensor operation. Modied electrodes, however, overcome this limitation.
With modied electrodes, a series of cresol samples with concentrations of 0.2, 0.5, 1, 2, 5, 7.5, and 10 M were detected in
amperometry mode 3 times at each concentration). Two representative examples of initial amperometric results are shown in Fig. 6a.
All samples were injected in random order to avoid system error
and to show fouling clearer since signal from a sample would decay
more if a higher concentration was previously injected. The calibration curve is shown in Fig. 6b. A linear relationship between the
peak current and the concentration of cresol was obtained, conrmed by the R2 being very close to unity. The error bars which
were contributed mainly from electrode fouling were too small to
be distinguished from the data symbols. Therefore, the detection of
0.210 M cresol was reliable with the modied electrode.

3.5. Detection of TCP with modied electrodes


To show the ability of detecting TCP samples, several TCP
samples with 10, 20, 30, 50, 100, 200, and 300 ppb were gasied, hydrolyzed, and detected with PEDOT:PSS modied electrode.

Fig. 7. Detection of TCP on bare/modied electrode. (a, left) Calibration of detection of TCP samples with concentrations of 10, 20, 30, 50, 100, 200, and 300 ppb on PEDOT:PSS
modied electrode, (a, right) Calibration of detection of 30, 50, 100, and 300 ppb TCP on bare electrode. 10 and 20 ppb were not detected and compared with modied
electrode, (b) responses from modied electrode were normalized to 100% and the error bars and signals from bare electrode were relatively calibrated. RSD was calculated
from the ratio of Stdev (standard deviation) to Avg (average). Unicell glassy carbon electrodes with 2 mm diameter were used. Note that the bare electrode was polished
after measurement of each set of 4 samples, while modied electrode was not polished through three sets of measurements.

X. Yang et al. / Sensors and Actuators B 177 (2013) 659667

Experiments have been carried out three times. The calibration of


all three experiments is shown in the left frame of Fig. 7a. Error bars
indicate the variation between all three experiments. A linear relationship was observed between current signal and concentration
of sample in the range of 50300 ppb. For comparison, the corresponding results from bare electrodes are shown in the right side
of Fig. 7a, excluding 10 and 20 ppb TCP. Although a linear relation
was also observed, the error bars were much bigger due to electrode
fouling. Note that the bare electrode was polished after measurement of each set of 4 samples. The fouling effect would be much
more serious if the electrode were left unpolished through the
whole measurement, which was done for modied electrode. To
observe the comparison clearly, the responses from modied electrodes were normalized to 100% and the signals from bare electrode
and error bars were relatively calibrated (Fig. 7b). Comparing the
error bars and RSD values from modied electrode with those from
bare electrode, modied electrode was seen to have much less fouling. Indicated by these results, the sensing system with PEDOT:PSS
modied electrode is able to continuously monitor TCP in gas phase

665

in the concentration range of 50300 ppb. It should be mentioned


that the lower point of the linear range for the PEDOT:PSS modied electrode (50 ppb) is comparable with others work. De Nola
reported the LOD of 1.4 pg/20 nL (70 ppb, TCP in isohexane) using
gas chromatography [13].
Previously, our group reported the detection of TCP with bare
glassy carbon electrodes with a linear range of 5300 ppb [51].
However, the bare electrodes required thorough polishing before
each measurement, which would make impossible its usage for
continuously monitoring TCP on the aircraft.
3.6. Detection of hydrolysates from engine oils
Finally, to test the ability of this system to determine TCP in
air from real samples, such as samples from jet engine oils, 2 L
of commercially available BP Turbo Oil 274, BP Turbo Oil 2380, and
Mobile Jet Oil II were mixed in 0.5 mL of methanol and were used to
prepare the hydrolysate samples with the same procedure as TCP
hydrolysates described above. The results of oil hydrolysates with

Fig. 8. Detection of commercial engine oil and oil spiked hydrolysate samples. All samples were measured randomly 3 times. (a) Results of hydrolysate samples from engine
oils BP 274, Mobil, and BP 2380 with modied electrode; (b) with unmodied electrode. (c) Calibration of the concentrations of TCP in the oil samples, and the results from
modied electrode and unmodied electrode were compared with the claimed values by manufactures. (d) Results of hydrolysate samples from oil BP 274 spiked with
0200 ppb TCP with modied electrode; (e) with unmodied electrode. (f) Comparison of the results between the modied and unmodied electrode. RSD (relative standard
deviation) was calculated the same as Fig. 7.

666

X. Yang et al. / Sensors and Actuators B 177 (2013) 659667

modied and unmodied electrodes are compared in Fig. 8ac. For


each set of amperometric measurements, several oil samples were
injected in random order. This procedure was repeated three times.
Fig. 8df show the results of measurements in BP274 oil spiked with
0, 50, 100, and 200 ppb TCP. Error bars indicate the variation among
three sets of experiments that is mainly brought on by the fouling.
Comparison between Fig. 8a and b, d and e shows that the fouling
effect has been reduced with modication. Moreover, the measurements with unmodied electrode required polishing before each
measurement, while the modied one did not.
Concentrations of TCP in these oils are summarized in Fig. 8c
and f. The concentration of TCP was calculated from peak current
and converted to % unit. The electrochemical responses from the
oil hydrolysates had a similar trend compared to the TCP concentrations in the oils as expected from the manufacturers data. The
deviations from the measurements of different oils with the modied electrode were much less than unmodied electrode (Fig. 8c),
the same as the RSD from BP 274 spiking TCP samples (Fig. 8f).
Thus, the modication appears to have reduced the effect of fouling
when detecting TCP in real engine oils and allows the continuous
monitoring in practical application on aircraft.
4. Conclusions
An electrochemical sensing system for continuous monitoring
of TCP in gas phase has been developed. The electrode fouling
occurring upon oxidation of cresol brought a critical problem
for continuous monitoring in long term applications. PEDOT:PSS
modication was controllable and was able to greatly reduce the
electrode fouling. By applying PEDOT:PSS modication, this sensing
system obtained reliable results from detecting TCP in the concentration range of 50300 ppb. The response to commercial jet engine
oils was similar to that of TCP methanol samples. This modication could be used in a wider range of applications. In the future,
we will focus on the detailed mechanisms of electrode fouling and
anti-fouling effect.
Acknowledgements
The authors would like to thank for the nancial supports from
US Federal Aviation Administration of the Ofce of Aerospace
Medicine and AUDFS Center, grant USDA-20053439415674A. This
material was based on work which supported ALS by the National
Science Foundation, while working at the Foundation. The views
expressed in this article are those of the authors, and do not reect
the ofcial policy or position of the United States Air Force, Department of Defense, NSF, or the U.S. Government. The help with AFM
imaging is also appreciated from Prof. Maria Auad in the Polymer
and Fiber Engineering Department at Auburn University. Also the
authors would like to thank L.C. Mathison and Dr. Clyde Wikle in
our department for the help on electronics and mechanical work
during sensor fabrication.
References
[1] AFL-CIO, Aircraft Air Quality: Whats Wrong with it and What Needs to be done,
The Aviation Subcommittee of The Transportation & Infrastructure Committee
U.S. House of Representatives, Washington, DC, 2003, pp. 140.
[2] H. Kinugasa, et al., Evaluation of toxicity in the aquatic environment based on
the inhibition of acetylcholinesterase, Nippon Suisan Gakkaishi 67 (4) (2001)
696702.
[3] J.E. Casida, et al., Mechanisms of selective action of pyrethroid insecticides, Annual Review of Pharmacology and Toxicology 23 (1) (1983)
413438.
[4] M.B. Abou-Donia, D.M. Lapadula, Mechanisms of organophosphorus esterinduced delayed neurotoxicity: type I and type II, Annual Review of
Pharmacology and Toxicology 30 (1) (1990) 405440.

[5] P.H. Craig, M.L. Barth, Evaluation of the hazards of industrial exposure to tricresyl phosphate: a review and interpretation of the literature, Journal of
Toxicology and Environmental Health, Part B: Critical Reviews 2 (4) (1999)
281300.
[6] L.et al. Rosenstock, Chronic central nervous system effects of acute organophosphate pesticide intoxication. The Pesticide Health Effects Study Group, Lancet
338 (8761) (1991) 223.
[7] A. Bacaloni, et al., Occurrence of organophosphorus ame retardant and plasticizers in three volcanic lakes of central Italy, Environmental Science &
Technology 42 (6) (2008) 18981903.
[8] N. Van den Eede, et al., Analytical developments and preliminary assessment
of human exposure to organophosphate ame retardants from indoor dust,
Environment International 37 (2) (2011) 454461.
[9] M.D. David, J.N. Seiber, Analysis of organophosphate hydraulic uids in U.S. Air
Force Base soils, Archives of Environmental Contamination and Toxicology 36
(1999) 235241.
[10] C. Winder, J.-C. Balouet, The toxicity of commercial jet oils, Environmental
Research 89 (2002) 146164.
[11] Federal-Register, Aryl Phosphates, in: Response to the Interagency Testing
Committee, 1983, p. 5745257460.
[12] M.D. David, J.N. Seiber, Accelerated hydrolysis of industrial organophosphates
in water and soil using sodium perborate, Environmental Pollution 105 (1)
(1999) 121128.
[13] G. De Nola, J. Kibby, W. Mazurek, Determination of ortho-cresyl phosphate
isomers of tricresyl phosphate used in aircraft turbine engine oils by gas chromatography and mass spectrometry, Journal of Chromatography A 1200 (2)
(2008) 211216.
[14] M. Mutsuga, et al., Determination method of tricresyl phosphate in polyvinyl
chloride, Shokuhin Eiseigaku Zasshi: Food Hygiene and Safety Science 44 (1)
(2003) 2631.
[15] J. Bhattacharyya, et al., Detection and estimation of tricresyl phosphate in mustard oil, Forensic Science 3 (1974) 263270.
[16] A.E. Habboush, S.M. Farroha, H.I. Khalaf, Extraction-gas chromatographic
method for the determination of organophosphorus compounds as lubricating
oil additives, Journal of Chromatography A 696 (1995) 257263.
[17] X. Yang, et al., Portable and remote electrochemical sensing system for detection of tricresyl-phosphate in gas phase, Sensors and Actuators B: Chemical
(2011).
[18] X. Yang, et al., Portable electrochemical sensor for detection of tricresylphosphate, ECS Transactions 35 (35) (2011) 3543.
[19] Z. Ezerskis, Z. Jusys, Electropolymerization of chlorinated phenols on a Pt electrode in alkaline solution. Part I: a cyclic voltammetry study, Journal of Applied
Electrochemistry 31 (10) (2001) 11171124.
[20] Z. Ezerskis, Z. Jusys, Electropolymerization of chlorinated phenols on a Pt electrode in alkaline solution. Part III: a Fourier transformed infrared spectroscopy
study, Journal of Applied Electrochemistry 32 (7) (2002) 755762.
[21] Z. Ezerskis, Z. Jusys, Electropolymerization of chlorinated phenols on a Pt electrode in alkaline solution. Part IV: a gas chromatography mass spectrometry
study, Journal of Applied Electrochemistry 32 (5) (2002) 543550.
[22] Z. Ezerskis, G. Stalnionis, Z. Jusys, Electropolymerization of chlorinated phenols
on a Pt electrode in alkaline solution. Part II: an electrochemical quartz crystal
microbalance study, Journal of Applied Electrochemistry 32 (1) (2002) 4955.
[23] M. Ferreira, et al., Electrode passivation caused by polymerization of different
phenolic compounds, Electrochimica Acta 52 (2) (2006) 434442.
[24] M. Fleischmann, et al., A Raman spectroscopic investigation of the electropolymerization of phenol on silver electrodes, Electrochimica Acta 28 (11) (1983)
15451553.
[25] M. Gattrell, D.W. Kirk, A study of the oxidation of phenol at platinum and preoxidized platinum surfaces, Journal of the Electrochemical Society 140 (6) (1993)
15341540.
[26] M. Giuliano, M.M. Marco, An overview of phenol electropolymerization for
metal protection, Journal of the Electrochemical Society 134 (12) (1987)
643C652C.
[27] J.D. Rodgers, W. Jedral, N.J. Bunce, Electrochemical oxidation of chlorinated
phenols, Environmental Science & Technology 33 (9) (1999) 14531457.
[28] F. Fichter, Die elektrolytische Oxydation des Toluols, Zeitschrift fr Elektrochemie und Angewandte Physikalische Chemie 19 (20) (1913) 781784.
[29] A. Nishiyama, et al., Anodic oxidation of 4-hydroxycinnamic acids and related
phenols, Chemical and Pharmaceutical Bulletin 31 (8) (1983) 28452852.
[30] R. Epur, Electrochemical Sensors for the Detection of Tricresyl Phosphate and
Determination of Acid Content in Engine Oils, Auburn University, 2009.
[31] S.H. Glarum, J.H. Marshall, Polarization relaxation in electrodeposited
polyphenylene-oxide lms, Journal of the Electrochemical Society 132 (12)
(1985) 29392944.
[32] S.K. Wheeler, L.A. Coury Jr., W.R. Heineman, Fouling-resistant, polymermodied graphite electrodes, Analytica Chimica Acta 237 (1990) 141148.
[33] V.A. Pedrosa, et al., Copper nanoparticles and carbon nanotubes-based electrochemical sensing system for fast identication of tricresyl-phosphate in
aqueous samples and air, Sensors and Actuators B: Chemical 140 (1) (2009)
9297.
[34] S. Patra, N. Munichandraiah, Electro-oxidation of phenol on polyethylenedioxythiophene conductive-polymer-deposited stainless steel substrate,
Journal of the Electrochemical Society 155 (3) (2008) F23F30.
[35] S. Lupu, et al., Polythiophene derivative conducting polymer modied electrodes and microelectrodes for determination of ascorbic acid. Effect of possible
interferents, Electroanalysis 14 (78) (2002) 519525.

X. Yang et al. / Sensors and Actuators B 177 (2013) 659667


[36] A. Malinauskas, Electrocatalysis at conducting polymers, Synthetic Metals 107
(2) (1999) 7583.
[37] H.B. Mark, et al., The electrochemistry of neurotransmitters at conducting
organic polymer electrodes: electrocatalysis and analytical applications, Bioelectrochemistry and Bioenergetics 38 (2) (1995) 229245.
[38] S.S. Rosatto, L.T. Kubota, G. de Oliveira Neto, Biosensor for phenol based on the
direct electron transfer blocking of peroxidase immobilising on silicatitanium,
Analytica Chimica Acta 390 (13) (1999) 6572.
[39] M.A. Heras, et al., A poly(3,4-ethylenedioxythiophene)-poly(styrene
sulphonate) composite electrode coating in the electrooxidation of phenol,
Electrochimica Acta 50 (78) (2005) 16851691.
[40] S. Lupu, et al., Electrochemical sensors based on platinum electrodes modied
with hybrid inorganicorganic coatings for determination of 4-nitrophenol and
dopamine, Electrochimica Acta 54 (7) (2009) 19321938.
[41] E. Tamburri, et al., Growth mechanisms, morphology, and electroactivity of
PEDOT layers produced by electrochemical routes in aqueous medium, Synthetic Metals 159 (56) (2009) 406414.
[42] N. Sakmeche, et al., Improvement of the electrosynthesis and physicochemical properties of poly(3,4-ethylenedioxythiophene) using a sodium
dodecyl sulfate micellar aqueous medium, Langmuir 15 (7) (1999)
25662574.
[43] L. Pigani, et al., Electro-oxidation of chlorophenols on poly(3,4-ethylenedioxythiophene)-poly(styrene sulphonate) composite electrode, Electrochimica Acta 52 (5) (2007) 19101918.
[44] WHO, et al., Tricresyl Phosphate, World Health Organization, 1990.
[45] D.R. Lide, CRC Handbook of Chemistry and Physics: A Ready-Reference Book of
Chemical and Physical Data, CRC Pr I Llc, 2004.
[46] CAMEO-Chemicals, Tricresyl Phosphate, 1999.
[47] C. van Netten, V. Leung, Comparison of the Constituents of Two Jet Engine
Lubricating Oils and Their Volatile Pyrolytic Degradation Products, Applied
Occupational and Environmental Hygiene 15 (2000) 277283.
[48] A.J. Bard, L.R. Faulkner, Electrochemical Methods: Fundamentals and Applications, 2nd ed., Wiley, New York, 2001, xxi, 833 p.
[49] X. Du, Z. Wang, Effects of polymerization potential on the properties
of electrosynthesized PEDOT lms, Electrochimica Acta 48 (12) (2003)
17131717.
[50] R.C. Koile, D.C. Johnson, Electrochemical removal of phenolic lms from a platinum anode, Analytical Chemistry 51 (6) (1979) 741744.

667

[51] X. Yang, et al., Portable and remote electrochemical sensing system for detection of tricresyl-phosphate in gas phase, Sensors and Actuators B: Chemical 161
(1) (2012) 564569.

Biographies
Xiaoyun Yang received the B.S. degree in Chemistry from Tongji University,
Shanghai, China in 2006, followed by 2-years studying for Masters degree in Physics
department. Currently, he is pursuing his PhD degree in Materials Engineering at
Auburn University.
Jeffrey Kirsch received his B.S. in Materials Engineering at Auburn University,
Auburn, Alabama in 2011. Currently, he is pursuing his Masters Degree in Materials
Engineering at Auburn University.
Dr. Eric V. Olsen is the Director of the Clinical Research at Keesler AFB Mississippi. He
received his Ph.D. in Biological Sciences and M.S. in Microbiology from Auburn University. His research interests include PCR assay development, piezoelectric-based
biosensor systems and bio-preservation techniques. He is a Lt. Colonel in the United
States Air Force with over 20 years of service.
Dr. Jeffrey W. Fergus received his B.S. degree in Metallurgical Engineering from the
University of Illinois in 1985 and his Ph.D. degree in Materials Science and Engineering from the University of Pennsylvania in 1990. He was a post-doctoral research
associate in the Center for Sensor Materials at the University of Notre Dame and, in
1992, joined the Materials Engineering program at Auburn University, where he is
currently a professor. His research interests are generally in high-temperature and
solid-state chemistry of materials, including electrochemical devices (e.g. chemical
sensors and fuel cells) and the chemical stability of materials (e.g. high temperature
oxidation).
Dr. Alex L. Simonian is a Professor of Materials Engineering at Auburn University and
a Biosensing Program Director at NSF. He received his M.S. in Physics from the Yerevan State University (Armenia, USSR), a Ph.D. in Biophysics and a Doctor of Science
degree in Bioengineering from the USSR Academy of Science. His current research
interests are primarily in the areas of bioanalytical sensors, nano-biomaterials and
functional interfaces.

Вам также может понравиться