Sunteți pe pagina 1din 8

Talanta 77 (2008) 710717

Contents lists available at ScienceDirect

Talanta
journal homepage: www.elsevier.com/locate/talanta

Fabrication of an ammonia gas sensor using inkjet-printed


polyaniline nanoparticles
Karl Crowley a , Aoife Morrin a , Aaron Hernandez a , Eimer OMalley a , Philip G. Whitten b ,
Gordon G. Wallace b , Malcolm R. Smyth a , Anthony J. Killard a,
a
b

School of Chemical Sciences, National Centre for Sensor Research, Dublin City University, Glasnevin, Dublin 9, Ireland
ARC Centre of Excellence for Electromaterials Science, Intelligent Polymer Research Institute, University of Wollongong, Northelds Avenue, Wollongong, NSW 2522, Australia

a r t i c l e

i n f o

Article history:
Received 14 March 2008
Received in revised form 26 June 2008
Accepted 4 July 2008
Available online 23 July 2008
Keywords:
Inkjet printing
Polyaniline nanoparticles
Gas phase ammonia analysis
Conductimetric
Interdigitated electrode array
Thermal analysis
Temperature control

a b s t r a c t
This work details the fabrication and performance of a sensor for ammonia gas analysis which has been
constructed via the inkjet-printed deposition of polyaniline nanoparticle lms. The conducting lms
were assembled on interdigitated electrode arrays and characterised with respect to their layer thickness
and thermal properties. The sensor was further combined with heater foils for operation at a range of
temperatures. When operated in a conductimetric mode, the sensor was shown to exhibit temperaturedependent analytical performance to ammonia detection. At room temperature, the sensor responded
rapidly to ammonia (t50 = 15 s). Sensor recovery time, response linearity and sensitivity were all signicantly improved by operating the sensor at temperatures up to 80 C. The sensor was found to have a stable
logarithmic response to ammonia in the range of interest (1100 ppm). The sensor was also insensitive to
moisture in the range from 35 to 98% relative humidity. The response of the sensor to a range of common
potential interferents was also studied.
2008 Elsevier B.V. All rights reserved.

1. Introduction
Ammonia is a highly toxic, polluting gas. With global production in excess of 100 million tonnes per annum, it has a wide
variety of uses including the production of nitrogenous fertilizers and other nitrogenous chemicals, as well as an industrial
refrigerant. Monitoring of leakage in a wide range of industrial
applications is thus deeply desired. A range of devices is available for monitoring ammonia, the most established of which are
electrochemical devices based on electrolytic cells. An overview
of the various ammonia sensing technologies under development
was detailed by Timmer et al. [1]. In brief, other ammonia sensing
devices include those based on solid state sensors [2], spectroscopic techniques [3] and conducting polymers [4]. Electrolytic
devices have been around for decades but generally suffer from
lower detection limits and limited accuracy. A variety of solid state
devices have shown promise with detection limits down to 1 ppm,
though selectivity can be an issue [1]. Spectroscopic sensors range
from simple, non-selective pH indicator-based sensors to complex
spectrometer-based systems capable of measuring down to 1 ppb

NH3 [3]. Though powerful, these systems are generally large and
costly, suited to the laboratory rather than low cost sensors. Conducting polymer-based sensors have been increasingly employed
over the last few years with detection limits in the ppm region being
demonstrated. However, these sensors can suffer from irreversible
reactions leading to a reduction in response [1].
Polyaniline (PANI) is a highly versatile conducting polymer.
Sensors composed of PANI and other conducting polymers have
routinely been applied to the analysis of ammonia and other
gases, such as a chip-based solution-cast PANI sensor described
by Kukla et al. [5], an ammonia sensor based on a PANIdodecylbenzenesulfonate emulsion (in chloroform) described by
Wu et al. [6] and more recently, the use of a PANI-coated lter paper
for a number of applications including the colorimetric detection
of gaseous and aqueous ammonia [7]. PANI is an excellent material for ammonia sensing as it deprotonates the amine groups in
the emeraldine salt converting it to the emeraldine base form with
a corresponding drop in conductivity of several orders of magnitude. A simplied version of this reaction is given in the following
equation:
PANI H+ + NH3  PANI + NH4 +

Corresponding author. Tel.: +353 17007871; fax: +353 17007873.


E-mail address: tony.killard@dcu.ie (A.J. Killard).
0039-9140/$ see front matter 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.talanta.2008.07.022

(1)

The very high afnity of PANI for ammonia does have its drawbacks, particularly with regards to sensor regeneration, where long
timescales of minutes or hours are required to return the sensor to

K. Crowley et al. / Talanta 77 (2008) 710717

its initial state. The chip-based PANI sensor described by Kukla et


al. incorporated heaters that were used to run a thermal regeneration step once the electrode displayed deviation [5]. This method
was used to reclaim sensor performance after exposures to lower
ambient ammonia conditions. Unfortunately, even with the thermal regeneration, irreversible changes to the sensor were observed
if it was exposed to high concentrations for an extended period.
A major issue for a sensor composed of PANI is the difculty in
handling or processing the polymers. Traditionally, aniline could be
polymerised by either chemical or electrochemical means, both of
which gave little control over the type of lm formed and its morphology at both the macro- and microscales. The advent of aqueous
formulations of conducting polymers and aqueous dispersions of
conducting polymer nanoparticles and other nanostructured materials has opened up a new range of fabrication options [8], while
also demonstrating superior functional properties for sensing and
other applications [9,10].
In the past few years, inkjet printing has rapidly emerged as a
means to print a wide variety of materials. Conducting polymers
[1113], metallic layers [14,15], and bio-materials [16,17] are examples of just some of the materials printed. Inkjet printing is also a
means to fabricate nely patterned, thin lms of conducting polymers. Previously, techniques such as spin-coating or drop-casting
have been used to apply conducting polymers though these methods do not readily allow any patterning and have various levels of
deposition control. A signicant advantage of inkjet printing is that
it is non-contact. Thus, multiple layers of varying composition can
be built up without the concerns of cross-contamination or physical
damage to previously printed layers. It also allows ner detailing
than screen printing with feature sizes of 10 m or less compared with the 50100 m possible with screen printingwhich
allows the fabrication of smaller, more complex devices. Wang et al.
achieved even greater resolution by modifying the substrate surface
energy to allow the possibility of sub-micron sized features with
inkjet-printed PEDOT/PSS [18]. In addition, devices fabricated from
inkjet-printed conducting polymer layers are more amenable to
low cost mass production [19], compared with more conventional
means of manufacture such as chemical or electrochemical polymerisation. Conducting polymers have previously been used in a
variety of devices, including electrochromic displays [20], batteries
and supercapacitors [21,22], fuel cells [23], antistatic layers [24] and
basic electrical components have also been fabricated using these
materials. Recently, an Al/PANI Schottky diode was constructed via
inkjet printing and assessed for its electronic and thermoelectric
properties [25]. Inkjet printing allows many of these devices to
be fabricated for a fraction of the cost of equivalent silicon-based
devices, allowing for much wider application in low cost/high volume applications such as RFID [19], smart packaging, disposable
biomedical devices, etc.
Previously, an aqueous polyanilineDBSA nanoparticle suspension was characterised and optimised for inkjet printing [26].
Optimal synthesis conditions resulted in polyaniline nanoparticle
suspensions with an average particle diameter of 82 nm. UVvis
analysis showed that the polyaniline was in the emeraldine salt
form with a high DBS doping level. In addition, a rheological study
was performed on the polyaniline nanosuspension; comparing
properties such as viscosity and surface tension with commercial
inkjet ink. In this paper, we describe the fabrication and application
of a fully printed sensor for ammonia fabricated from an inkjetprinted lm of the polyanilineDBSA nanoparticles deposited onto
screen-printed silver interdigitated electrodes. Temperature control of the sensor was provided by commercially available heater
foils. The sensor was optimised for gas phase analysis and applied
to the analysis of ammonia in air. Methods of enhancing recovery
times and tuning the sensor response were also investigated and

711

optimised. The effects of temperature, humidity, long term stability


and effect of potential interferents were also investigated.
2. Experimental
The synthesis of the polyaniline nanoparticle suspension has
been detailed previously [26,27]. In brief, 0.6 ml of freshly distilled
aniline (SigmaAldrich), 3.4 g dodecylbenzenesulfonic acid (DBSA,
TCI) and 0.36 g ammonium persulfate (APS, Sigma) were added to
40 ml deionised water and stirred for 2.5 h until a dark green viscous
solution was formed. The solution was centrifuged at 5000 rpm
for 30 min and then dialysed against 0.05 M sodium dodecyl sulfate (SDS) for 2 days to remove excess DBSA and APS. Electrical
measurements were performed on either a CHI 1000 potentiostat (CH Instruments, Inc.) or a Palmsens (Palm Instruments BV),
both controlled via a PC. Sensors were connected to the potentiostat in a two-electrode conguration (working electrode and
common reference and auxiliary electrode). Current measurements
were performed either using xed potential at typically +0.1 V or
by applying potential cycling (saw-tooth) waveform to the sensor
between two equivalent but opposite potentials, e.g. +0.1 to 0.1 V.
Headspace gas measurements were performed either in a 1.1-L or
a 3-L chamber, both with sealed access points for cables. Ammonia was introduced through an additional sealed opening in both
chambers. The 3-L chamber contained a fan for rapid mixing of the
introduced gas and an Impulse XP Gas Detector (Zellweger Analytics) was employed within the 1.1-L chamber as a control method
for ammonia analysis.
2.1. Fabrication of sensors
NanoPANI-modied interdigitated electrode arrays (nanoPANIIDAs) were fabricated using screen printing and inkjet printing
methods. The electrode patterns were designed using AutoCAD
(Autodesk) and silver and carbon IDA patterns were prepared using
screen printing. Screen printing was performed using a DEK model
247 and the IDA patterns were printed on 175 m polyethylene
terephthalate (PET 505, HiFi lms Ltd., Dublin) substrate and cured
at 150 C for 30 min.
Inkjet printing was performed using unmodied Epson C46/C48
piezoelectric printers. Both the black and color printer cartridges
were opened and the ink removed. The cartridges were rinsed
thoroughly with deionised water prior to being relled with the
nanoPANI suspension. Print designs were drawn using standard
Windows software (e.g. MS Word) and printing was performed
through the supplied software in the standard way. The printer
was rst primed with standard cartridges and once good quality
prints were obtained with the commercial ink, the cartridges were
replaced with the nanoPANI cartridges. Priming the printer with
the nanoPANI solution involved the use of cleaning cycles and ink
purging features until a good quality print was achieved. These
cleaning/purging cycles were controlled using the SSC Service Utility (available from http://www.ssclg.com/epsone.shtml). Following
printing, the sensors were gently rinsed in deionised water before
being heat cured at 75 C for 30 min.
Sensors were also constructed by attaching polyimide-based
exible heating foils (Minco, USA) to the back of the PET substrate.
Temperature was monitored through a Thermal Tab thermistor
(S665PDZ40B, Minco, USA). Heater control and temperature monitoring was performed with a CT325 Temperature Controller (Minco,
USA).
2.2. Electrical measurement
The electrode design chosen as suitable for gas phase analysis
was the interdigitated array (IDA). IDA electrodes have been widely

712

K. Crowley et al. / Talanta 77 (2008) 710717

used in conjunction with thin lms, typically with conductimetric


measurements for both characterisation and sensor applications. Zaretsky et al. presented a detailed theoretical treatment
of interdigitated electrodes and developed an impedance-based
continuum model that allowed the estimation of a number of
parameters, including lm thickness, lm permittivity and conductivity [28]. Sheppard et al. utilised this model and found that
cell constants calculated from an electromagnetic eld model were
in strong agreement with their experimental results [29]. In the
case of the nanoPANI sensor, a silver IDA substrate was employed
as a means of measuring the change in measured current (and
hence lm conductivity) as the lm interacts with gaseous ammonia. Current measurements were performed using either potential
step or potential cycling techniques. For unheated sensors, a small
baseline drift was observed when a xed potential (+0.1 V) was
employed. Cycling the potential between two equal but opposite
potentials (e.g. +0.1 and 0.1 V at 0.2 V s1 ) was found to minimise this drift. The result was an Ohmic plot. Processing this
data involved sampling the current at a predened potential. The
sampled current data was then plotted against time, yielding a
quasi-chronoamperometric plot. In heated sensors the short-term
drift was not observed so a xed potential was employed.
2.3. Thermal characterisation of nanoPANI lms
Differential scanning calorimetry (DSC) was performed in a
nitrogen atmosphere using a TA Instruments (USA) Q200 differential scanning calorimeter. The samples used for DSC consisted of
dropcast nanoPANI dispersions (10 mg sample size). DSC heating
scans were performed at a rate of 10 C min1 , and cooling scans
at a rate of 5 C min1 . Thermogravimetric analysis (TGA) was performed with a TA instruments Q500 thermogravimetric analyzer in
an air atmosphere at a heating rate of 5 C min1 .
2.4. Characterisation of the effects of temperature and humidity
Heat curing and preliminary heating experiments were performed using a Mammert 400 oven. The effects of temperature and
humidity were assessed by placing the electrodes in a MTCL 350
environmental chamber (Tas Ltd., UK) over a range of temperature
and humidity settings, the connections made through a sealable
cable port in the environmental chamber.
3. Results and discussion
3.1. Sensor fabrication
Fig. 1(a) and (b) shows the two sensors without and with the
heater foil, respectively. Even with the heating foil, sensor thickness
was below 1 mm. The numbers visible at the base of the electrodes
(e.g. 200 1500 in Fig. 1(a)) refer to the digit width and digit spacing in micrometers. Varying the spacing and digit number would
affect the measured current but was not found to have any effect on
normalised ammonia response, i.e. the measured current response
divided by the initial, baseline current (I/I0 ). Fig. 1(c) shows an
exploded schematic of the foil-heated sensor.
3.2. Thermal characterisation
Most chemical sensors are affected by thermal conditions, with
extremes in temperature having an effect on response or lifetime.
Conducting polymer-based sensors are no exception. Many thermal analyses have been performed on polyaniline compositions as
differences in these materials such as dopant, synthesis method,
morphology, etc., can all have an effect on the behaviour of the

Fig. 1. Polyaniline interdigitated electrodes (nanoPANI-IDAs) shown alone (a) and


with a thermofoil heater (b). Exploded schematic diagram of the nanoPANI-IDA
electrode showing the different layers of the sensor (c).

material and make comparisons difcult. Previously, Jurczyk et al.


noted that a variety of polyaniline nanocomposites were gravimetrically stable below 200 C, with small mass losses at lower
temperatures attributable to loss of moisture [30]. Neoh et al. also
observed relative stability below 200 C for PANIDBSA with the
onset of weight loss occurring between 200 and 225 C [31]. Chen
observed a gradual decrease in mass above 100 C for a PANIDBSA
powder, attributable to moisture loss, with an accelerated drop in
mass over 250 C which was explained as evaporation or degradation of DBSA [32].
As noted earlier, temperatures will affect the equilibrium of Eq.
(1) in many ways. Therefore, to determine the effect of heat on the
nanoPANI sensors, TGA and DSC were performed on cured (75 C
for 30 min) and uncured samples of the nanoPANI particle suspension. From the TGA plot in Fig. 2(a) it can be seen that the cured
sample displays a gradual weight loss between 50 and 150 C, over
which the polymer loses approximately 8% of its initial mass. This
initial mass loss is most likely due to the loss of residual moisture
in the lm and is similar that observed by Chen [32] and Tsotcheva
et al. [33] for PANIDBSA powder. At 150 C the rate of loss could
be seen to accelerate and then remain essentially constant over the
temperature range of 150 500 C as the sample weight dropped
to approximately 30%. In contrast, the uncured sample displays
a rapid mass drop of approximately 25% as the temperature is
increased to 100 C, indicating moisture removed from the lm. At
temperatures above 100 C, the uncured sample follows the same
trend as the cured sample. This is clearer in the inset of Fig. 2(a)
where the rate of % mass change per unit time is plotted against
temperature where both samples display similar variations at temperatures greater than 100 C. At the end of the experiment, a white
residue approximately 10% of the original mass remained. In addition, the rate of weight loss was constant which implies weight loss
at these temperatures was due to drying, as a change in slope and
hence a change in the rst differential is expected if the mechanism
changes. The accelerated weight drop above 150 C occurred at
noticeably lower temperatures than observed in other studies, and

K. Crowley et al. / Talanta 77 (2008) 710717

713

Fig. 2. Thermogravimetric analysis (a) and differential scanning calorimetry (b) of nanoPANI. TGA: sample temperature increased at a rate of 5 C min1 , inset shows rate of
change of % mass against temperature. DSC: sample temperature increased at 10 C min1 and decreased at 5 C min1 . Endothermic is shown as negative heat ow. Cured
sample (solid line) and uncured (dotted line).

is due to degradation of PANIDBSA. The DSC data in Fig. 2(b) for the
cured sample shows no major features over the range investigated
(40 to 100 C) whereas the uncured sample shows distinctive
peaks: at 10 C (exothermic) and 38 C (endothermic), shifting to
30 C on subsequent scans. These peaks are due to crystallisation
and melting of hydrated DBSA complexes [33]. Therefore, it is
apparent that once a PANI lm is cured, it displays no observable
structural changes over the temperature range studied while
uncured lms show changes associated with the water present.
3.3. Effect of operating temperature on sensor behaviour
The impact of elevated temperature on the sensor performance
was investigated with the assumption that an increase in temperature would have an impact on the reaction kinetics on the
interaction of PANI with ammonia due to changes in the association
and dissociation rate constants of the binding interaction as well
as the partition coefcient of the ammonia between the solid and

gaseous phases. To this end and for its potential application across
a range of environmental conditions, it was vital to understand the
effects of heat on the behaviour of the sensor; specically the effect
on sensor conductivity, ammonia response and lifetime. Numerous models have been proposed to explain conduction mechanisms
within polyaniline and the variations in conductivity observed at
different temperatures. Kulkarni et al. observed only very slight
increases in conductivity for a variety of different inorganic acids
over a 25125 C range (perchloric acid being a particular exception,
showing a drop over this range) [34]. Yakuphanoglu and Senkal
found three different conductive regions for polyaniline synthesised in ionic liquid [25].
The effect of temperature on the sensor background current is
given in Fig. 3. In Fig. 3(a), a nanoPANI-IDA sensor was progressively heated from room temperature to 160 C. Initially, only a
slight increase in current was observed, though at temperatures
above 60 C, this increase was much more pronounced. At temperatures between 120 and 140 C, the rate of increase began to slow

Fig. 3. Temperature effects on nanoPANI-IDA sensors under ambient atmospheric conditions. (a) Effect on background current of increasing sensor temperature from 20 to
160 C. (b) Temperature cycling between 40 and 100 C. Electrodes were cycled from +1 to 1 V at 2 V s1 , current sampled at +1 V. Rapid degradation indicated by .

714

K. Crowley et al. / Talanta 77 (2008) 710717

Fig. 4. Response (a) and recovery (b) of the nanoPANI-IDA on exposure to ammonia vapour at 60 ppm. Electrodes were cycled from +0.1 to 0.1 V at 0.2 V s1 , current sampled
at +0.1 V. 1 layer (solid), 10 layers (dashed), 25 layers (medium dashed) and 40 layers (short dashed).

before a rapid, irreversible degradation was observed above 140 C


at which point a permanent drop in measured current occurred.
This rapid degradation correlates with the onset of thermal degradation implied by the accelerated rate of weight loss above these
temperatures in the TGA data in Fig. 2(a).
Fig. 3(b) shows the effect of cycling the temperature between
40 and 100 C on the background current. In the rst cycle,
a similar curve to that in Fig. 3(a) was obtained. However, in
the subsequent two scans (corresponding to a decrease and
increase in temperature, respectively) an apparently linear relationship formed over the full range. This difference in conductivity
behaviour observed between the rst and subsequent temperature
cycles matches the heat ow changes observed for the DSC data in
which also stabilised following the rst cycle. The reason for this
process is not clear but may relate to the removal of moisture from
the lm in the initial cycle as it has been found that, when left under
ambient conditions, the initial thermal prole is re-established.
This demonstrates that the linearity of the conductivity change of
cured lms between 40 and 100 C correlates with the lack of
features observed in the DSC data for the cured lm in Fig. 2(b).
3.4. Effect of print thickness on ammonia response at room
temperature
An important consideration in the fabrication of the sensor
was the effect that the number of nanoPANI layers would have
on the sensor behaviour. Thicker lms would presumably result
in better conductance and therefore greater measured currents.
However, this may impact negatively on the response time of the
sensor. In terms of production, thinner lms mean less materials
and faster manufacture. However, care must be taken to ensure
viable and reproducible lms are obtained. To this end, the relationship between nanoPANI layer thickness and electrode response
to ammonia was investigated. The thickness of nanoPANI layers
has been found to be approximately 170 nm per layer (for 10 layers
and less) [35]. To assess the effect of increasing layer thickness on
current under atmospheric conditions, a range of nanoPANI lms
were inkjet-printed onto the IDAs using different numbers of prints
and therefore different layer thicknesses. After each print, the fresh
nanoPANI layer was allowed to dry after which the current response
(A) was measured through each electrode. From 1 to 10 prints a
linear response (y = 60.4x 67.2) was observed as the layers were

built up. Over 10 prints, the measured current began to plateau, levelling off completely over 20 prints which appeared to indicate that
the nanoPANI lm was moving from conductance that was limited
by thickness to bulk conductance behaviour.
Film thickness may be expected to have a signicant effect on
the analytical response to ammonia. Response and recovery times
of sensors are vitally important parameters. The Instrument Association of America (ISA) species that ammonia detectors should
reach a minimum of 50% of response within 90 s (i.e. t50 < 90 s) on
exposure to a xed concentration of ammonia gas [36]. Likewise for
recovery, a t50 of 90 s is specied when the detector is exposed to
clean air. For the purposes of testing the sensors, t50 was taken as
the point at which the sensor reached 50% of its nal steady state
value. A range of sensors of varying numbers of prints were exposed
to a quantity of ammonia to ascertain the effect of lm thickness.
Fig. 4 shows the response (a) and recovery (b) of the nanoPANI
sensors on exposure to ammonia gas (60 ppm) in which the measured current has been normalised with respect to the initial and
nal readings and where time zero is taken as the moment of exposure to the ammonia vapour. From Fig. 4(a) it can be observed that
all sensors displayed a rapid response to the ammonia. However, the
single layer print displayed the fastest response with a t50 below
15 s. The thicker lms (10, 25 and 40 layers) took marginally longer
but still yielded t50 values in the region between 15 and 20 swell
below the 90 s specied in the ISA standard [36].
Fig. 4(b) shows the recovery of the sensors in air once the ammonia had been removed (time zero). In contrast to the rapid response
times, the recovery was very slow in all cases with little difference
observed due to lm thickness. The rapid response of the electrodes
coupled with slow recovery implied that under ambient conditions,
the equilibrium of Eq. (1) was heavily shifted to the right, demonstrating the high afnity of nanoPANI (emeraldine salt) to ammonia.
As no major variation in response time was found for the different thicknesses, in addition to the fact that larger currents were
obtained for thicker lms, further sensors were fabricated using 10
inkjet-printed layers, corresponding to approximately 1.7 m layer
thickness.
3.5. Quantitative analysis of ammonia at room temperature
Initial calibration testing involved ascertaining the response of
the sensors to ammonia under ambient conditions. Two replicate

K. Crowley et al. / Talanta 77 (2008) 710717

Fig. 5. Effect of heat on sensor response and recovery to the repeated application
and removal of 50 ppm ammonia at 60, 70, 80 C. +0.1 V potential applied to sensor.
Currents normalised with respect to initial current (clean atmosphere).

calibrations were performed on the same electrode, 35 days apart.


For a sensor at room temperature, a drop in measured current by
about a factor of ve was observed when the sensor was exposed
to ammonia at sub-10 ppm concentrations. This sharp initial drop
implies that the sensor can potentially monitor ammonia concentrations well into the sub ppm region, although this region is
generally outside that required for environmental health and safety
applications, which was the subject of this study. Plotting normalised current, (I I0 )/I0 , against the log of concentration yielded
a linear t of y = 0.129x + 0.631 on the rst day and y = 0.127x + 0.628,
35 days later displaying good stability over the 5-week period
tested.
In Fig. 4(b) the very slow recovery of the sensors after ammonia
exposure was noted. To test the effect of temperature on the analytical response, sensors with heater foils were used (see Fig. 1(b) and
(c)). These heater-sensors allowed a rapid appraisal of temperature
effects on the sensor behaviour. The thin build of the sensors also
allowed rapid temperature equilibration and reduced power consumption, which are also important concerns for sensors. As the
application of heat shifts the equilibrium of Eq. (1) to the left, it is
expected that sensitivity to ammonia will also decrease. Earlier, it
was noted that the sharp initial drop in measured current between 0
and 10 ppm (440 nM L1 ) implied that the sensors possessed higher
sensitivity than that required for environmental analytical applications (1100 ppm, 44 nM L1 to 4.4 M L1 ). Therefore there was
scope to trade off sensitivity for improved sensor recovery times
by employing operation at elevated temperatures.
3.6. Analytical characterisation at elevated temperatures
Preliminary studies had shown that the most promising temperature region for sensor recovery was between 60 and 80 C as
temperatures below this did not result in noticeably better recovery
than that obtained at room temperature conditions. Fig. 5 shows
the operation of the sensor at temperatures of 60, 70 and 80 C.
The heater foil was powered on after 150 s and stabilised at its
desired temperature within 60 s. An increase in the signal noise was
observed once the heater was activated. This was due to the sen-

715

Fig. 6. Effect of operating temperature on analytical response to ammonia (I/I0 vs.


[NH3 ]). Inset shows (I0 I)/I0 vs. log[NH3 ] at 60 C, 70 C and 80 C. +0.1 V potential
applied to sensor.

sor temperature drifting above and below the desired temperature


due to the controller being a simple on/off device and could be easily eliminated with improved control electronics. At 600 s, 50 ppm
ammonia was introduced into the chamber and then vented 300 s
later, this procedure being repeated twice more. As sensor temperature was increased, the recovery time was seen to decrease
noticeably. At 60 C, the recovery was still signicant, with the sensor showing partial recovery after 5 min and only approaching the
baseline after 15 min, following the third ammonia injection. The
sensor at 70 C displayed a noticeably faster recovery reaching the
baseline at 5 min, while the sensor heated to 80 C displayed the
fastest recovery, with full recovery in approximately 2.5 min.
Fig. 6 compares the responses obtained for ammonia for heated
sensors at a range of temperatures. The main plot shows the current drops measured at each temperature on a normalised scale
to allow comparison, which also takes into account the increase in
background current due to the elevation in temperature. Compared
with the response of the unheated sensor, the sensors heated to 60,
70 and 80 C show a much decreased initial drop due to the reduction in the signal-to-background ratio. The inset shows the same
data (for 60, 70 and 80 C) with a linear positive slope ((I0 I)/I0 vs.
log[NH3 ]), where the initial value (no NH3 present) was equal to
zero.
Once again, the greater drop at lower temperatures for similar
concentrations of ammonia implies that ammonia concentrations
of 1 ppm should be easily detectable. Kemp et al. demonstrated
that this temperature-dependent increase in response to ammonia
extends well below 0 C for polypyrrole [37]. Therefore, varying the
temperature of the sensor may be a means of tuning the response
to the analytical region of interest. In this case, the desired range
was 1100 ppm, and log plots could be obtained by heating the
sensor element to 70 C. Heating the sensors had the twin benets
of considerably improving the regeneration time while yielding an
improved response in the region of interest. In terms of stability,
heating the sensors was found to cause a gradual reduction in the
sensor response over time. Three weeks of continuous heating at
70 C resulted in a steady drop to approximately 85% of the initial

716

K. Crowley et al. / Talanta 77 (2008) 710717

Fig. 7. Response and recovery of nanoPANI sensors (at 70 C) to 100 ppm of different species. (a) Trimethylamine (TMA) and triethylamine (TEA). (b) Hydrogen sulde,
nanoPANI-silver IDA, hydrogen sulde, nanoPANI carbon IDA and NOx . In both graphs, the response to 100 ppm ammonia is given. +0.1 V potential applied to sensor.

background current. Reducing the continuously applied temperature to 50 C resulted in stable baselines over a 2-month period
(data not shown).
3.7. Effect of humidity and interferents
To ascertain the effect of humidity on the sensor response,
nanoPANI-IDAs were enclosed within an environmental chamber
and subjected to changes in atmospheric conditions. The effect of
water vapour on the ammonia response of a sensor is difcult to
determine precisely as the moisture will react with ammonia to
produce ammonium hydroxide. However, the effect of humidity
on the sensor background can be readily established. A number of
polyaniline sensors have been found to display small but observable
effects due to water vapour [38,39]. The nanoPANI sensor showed
only a slight change in background response from 35 to 98% relative humidity with a mean value of 12.06 0.42 A over the range
tested.
To assess the effect of potential interferents, the nanoPANI sensors were exposed to a variety of gaseous species. For the nanoPANI
sensors heated to 70 C, a wide variety of volatile organics (such
as acetone, methanol, dichloromethane, etc.) were tested at 100,
1000 ppm and above, with no measurable response being obtained
for any of these species at these concentrations. As the mechanism
for detection in polyaniline sensors is generally based on the protonation/deprotonation of the polymer, these sensors can be prone to
interference from other acidic or basic species. Carbon dioxide, carbon monoxide and acetic acid vapour were found to have little effect
on the nanoPANI sensor at concentrations of 100 and 1000 ppm.
However, a number of gaseous species were found to interfere with
the sensor. Fig. 7 shows the nanoPANI sensor response to a variety
of acidic and basic gaseous species and selectivity data of the sensor
to these species, relative to ammonia, is provided in Table 1.
Fig. 7(a) compares the response of two volatile amines with
that of ammonia at 100 ppm concentrations. Aliquots of these were
injected at 300 s and the sensors allowed stabilise over 30 min, after
which the chamber was vented with air to assess sensor recovery.
In comparison with the fast ammonia response, the two volatile
amines were seen to take considerably longer to reach steady state
and were still not fully stable after 30 min. This could be due to
the fact that the larger molecules require more time to penetrate
the nanoPANI lm. From Table 1 it can be seen that the sensors

displayed a similar selectivity for ammonia and trimethylamine,


with triethylamine response about 70% of this after 30 min exposure. Relatively rapid recovery can be observed for ammonia and
trimethylamine, though triethylamine only displayed a very weak
recovery over this time. Fig. 7(b) shows the results obtained for
100 ppm hydrogen sulde and NOx . When exposed to hydrogen
sulde, the nanoPANI sensors (on silver IDAs) displayed an unexpected drop in measured current; approximately three times that
observed for ammonia. When the same test was performed for
nanoPANI sensors on carbon IDAs, the measured current shows a
steady increase of approximately 25%, equivalent but opposite to
the ammonia response, implying that the nanoPANI is being further
protonated. In a study on hydrogen sulde sensors based on modied polyanilne, Virji et al. noted a weak response was obtained
for the unmodied polyaniline in the presence of hydrogen sulde, though this was for a 10-ppm exposure [40]. The hydrogen
sulde response observed for the nanoPANI-silver IDA electrode is
clearly due the reaction between hydrogen sulde and the silver
electrode. It should be noted that use of carbon or silver IDA electrodes had little effect on the normalised response obtained for
ammonia, though carbon IDA based sensors displayed an intrinsic
conductivity of about two and a half orders of magnitude below
nanoPANI-silver IDA sensors. The injection of 100 ppm NOx had a
very pronounced effect on measured current, with a drop of approximately 3.5 orders of magnitude. Once the system was purged with
air, a very slow recovery was observed. This was found to level off
at approximately 33% the initial measured current over the following 12 h. Identical results were obtained whether silver or carbon
interdigitated electrodes substrates were employed and point to
an effect of the gas on the nanoPANI. This effect was consistent
with the effect observed for nitrogen dioxide on the emeraldine
Table 1
Response and selectivity data of the nanoPANI sensors to a variety of gaseous species
(100 ppm)
Species

I/I0

(I0 I)/I0

Selectivity

NH3
H2 S (carbon IDA)
H2 S (silver IDA)
NOx
Trimethylamine
Triethylamine

0.76
1.25
0.28
3.84E03
0.75
0.83

0.24
0.25
0.72
1.00
0.25
0.17

1
1.04
2.96
4.13
1.02
0.72

K. Crowley et al. / Talanta 77 (2008) 710717

salt form of PANI, suggesting that the NO2 component of the NOx
mixture is affecting this change. Elizalde-Torres et al. found that
the emeraldine salt form was oxidised rst to emeraldine base and
then to pernigraniline (the non-conductive, fully oxidised form of
PANI) by NO2 [41]. They also noted that the desorption of NO2 from
PANI was only slowly reversible in an ambient, humid atmosphere.
Current and future work involves improving the sensor selectivity to discriminate against these interfering species, possibly using
selective membranes or modications to the nanoPANI layer.
4. Conclusion
The development and optimisation of a sensor composed of
an inkjet-printed polyaniline nanoparticles was demonstrated. An
aqueous dispersion of the nanoPANI was deposited over a silver
interdigitated array using a piezoelectric inkjet printing technique.
The nanoPANI lms were found to have stable thermal properties
which made them amenable to operation at elevated temperatures.
The nanoPANI sensors were found to be highly responsive to
gaseous ammonia with calibration plots obtained within the analytically important 1100 ppm region. Heating the sensors was
found to have the double benet of improving recovery times and
allow control of the analytical prole while the use of heater foils
was established as a low cost and readily implementable means to
control sensor temperature. Though the sensors displayed no interference from changes in humidity or from a range of volatile organic
compounds, some cross-sensitivity to other acidic and basic species
was observed.
Acknowledgement
The authors would like to thank Enterprise Ireland for funding
under the Proof of Concept fund PC/06/113.
References
[1] B. Timmer, W. Olthuis, A. van den Berg, Sens. Actuators B: Chem. 107 (2005)
666.
[2] A. Dubbe, Sens. Actuators B: Chem. 88 (2003) 138.
[3] G.H. Mount, B. Rumburg, J. Havig, B. Lamb, H. Westberg, D. Yonge, K. Johnson,
R. Kincaid, Atmos. Environ. 36 (2002) 1799.

[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]

717

D. Nicolas-Debarnot, F. Poncin-Epaillard, Anal. Chim. Acta 475 (2003) 1.


A.L. Kukla, Y.M. Shirshov, S.A. Piletsky, Sens. Actuators B: Chem. B37 (1996) 135.
S.Z. Wu, F. Zeng, F.X. Li, Y.L. Zhu, Eur. Polym. J. 36 (2000) 679.
D. Dutta, T.K. Sarma, D. Chowdhury, A. Chattopadhyay, J. Colloid Interface Sci.
283 (2005/3/1) 153.
A. Morrin, F. Wilbeer, O. Ngamna, S.E. Moulton, A.J. Killard, G.G. Wallace, M.R.
Smyth, Electrochem. Commun. 7 (2005) 317.
H.D. Tran, R.B. Kaner, Chem. Commun. (2006) 3915.
F. Masdarolomoor, P.C. Innis, S. Ashraf, R.B. Kaner, G.G. Wallace, Macromol. Rapid
Commun. 27 (2006) 1995.
J. Bharathan, Y. Yang, Appl. Phys. Lett. 72 (1998) 2660.
B. Ballarin, A. Fraleoni-Morgera, D. Frascaro, S. Marazzita, C. Piana, L. Setti, Synth.
Met. 146 (2004) 201.
Y. Yoshioka, G.E. Jabbour, Synth. Met. 156 (2006) 779.
H.H. Lee, K.S. Chou, K.C. Huang, Nanotechnology 16 (2005) 2436.
K.J. Lee, B.H. Jun, T.H. Kim, J. Joung, Nanotechnology 17 (2006) 2424.
T. Okamoto, T. Suzuki, N. Yamamoto, Nat. Biotechnol. 18 (2000) 438.
T. Xu, J. Jin, C. Gregory, J.J. Hickman, T. Boland, Biomaterials 26 (2005) 93.
J.Z. Wang, J. Gu, F. Zenhausem, H. Sirringhaus, Appl. Phys. Lett. 88 (2006) 133502.
V. Subramanian, P.C. Chang, J.B. Lee, S.E. Molesa, S.K. Volkman, IEEE Trans.
Compon. Pack. Technol. 28 (2005) 742.
R.J. Mortimer, A.L. Dyer, J.R. Reynolds, Displays 27 (2006) 2.
G. Wegner, Polym. Adv. Technol. 17 (2006) 705.
K.S. Ryu, S.K. Jeong, J. Joo, K.M. Kim, J. Phys. Chem. B 111 (2007) 731.
V. Neburchilov, J. Martin, H.J. Wang, J.J. Zhang, J. Power Sources 169 (2007) 221.
G. Deeuw, R. Samijn, I. Hoogmartens, D. Vanderzande, J. Gelan, Synth. Met. 57
(1993) 3702.
F. Yakuphanoglu, B.F. Senkal, J. Phys. Chem. C 111 (2007) 1840.
O. Ngamna, A. Morrin, A.J. Killard, S.E. Moulton, M.R. Smyth, G.G. Wallace, Langmuir 23 (2007) 8569.
S.E. Moulton, P.C. Innis, L.A.P. Kane-Maguire, O. Ngamna, G.G. Wallace, Curr.
Appl. Phys. 4 (2004) 402.
M.C. Zaretsky, L. Mouayad, J.R. Melcher, IEEE Trans. Electr. Insul. 23 (1988) 897.
N.F. Sheppard, R.C. Tucker, C. Wu, Anal. Chem. 65 (1993) 1199.
M.U. Jurczyk, A. Kumar, S. Srinivasan, E. Stefanakos, Int. J. Hydrogen Energy 32
(2007) 1010.
K.G. Neoh, M.Y. Pun, E.T. Kang, K.L. Tan, Synth. Met. 73 (1995) 209.
C.H. Chen, J. Polym. Res.-Taiwan 9 (2002) 195.
D. Tsotcheva, T. Tsanov, L. Terlemezyan, S. Vassilev, J. Therm. Anal. Calorim. 63
(2001) 133141.
V.G. Kulkarni, L.D. Campbell, W.R. Mathew, Synth. Met. 30 (1989) 321.
A. Morrin, O. Ngamna, E. OMalley, N. Kent, S.E. Moulton, G.G. Wallace, M.R.
Smyth, A.J. Killard, Electrochim. Acta 53 (2008) 5092.
ISA-92.03.01-1998Performance Requirements for Ammonia Detection Instruments (25500 ppm), Instrument Society of America, 1998, p. 22.
N.T. Kemp, A.B. Kaiser, H.J. Trodahl, B. Chapman, R.G. Buckley, A.C. Partridge,
P.J.S. Foot, J. Polym. Sci. Part B: Polym. Phys. 44 (2006) 1331.
G.E. Collins, L.J. Buckley, Synth. Met. 78 (1996) 93.
L. Grigore, M.C. Petty, J. Mater. Sci.: Mater. Electron. 14 (2003) 389.
S. Virji, J.D. Fowler, C.O. Baker, J. Huang, R.B. Kaner, B.H. Weiller, Small 6 (2005)
624.
J. Elizalde-Torres, H. Hu, A. Garcia-Valenzuela, Sens. Actuators B 98 (2004) 218.

S-ar putea să vă placă și