Você está na página 1de 8

Talanta 78 (2009) 498505

Contents lists available at ScienceDirect

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

Lead(II) ion-selective electrode based on polyaminoanthraquinone particles


with intrinsic conductivity
Xin-Gui Li a,b, , Xiao-Li Ma a , Mei-Rong Huang a,b,
a Institute of Materials Chemistry, Key Laboratory of Advanced Civil Engineering Materials of the Ministry of Education, College of Materials Science and Engineering,
Tongji University, 1239 Si-Ping Road, Shanghai 200092, China
b Key Laboratory of Molecular Engineering of Polymers of the Ministry of Education, Fudan University, Shanghai 200433, China

a r t i c l e

i n f o

Article history:
Received 2 August 2008
Received in revised form
27 November 2008
Accepted 29 November 2008
Available online 6 December 2008
Keywords:
Polyaminoanthraquinone
LeadII) ion-selective electrode
Conducting polymer
Lead(II) ionophore
Performance optimization
Membrane electrode

a b s t r a c t
A new polyvinylchloride membrane electrode was facilely prepared by using polyaminoanthraquinone
(PAAQ) microparticles with an intrinsically electrical conductivity as a lead(II) ionophore. It is found
that the electrode performance will signicantly be improved with adding 1 wt% PAAQ microparticles and decreasing the membrane thickness. A 90 m-thick membrane electrode consisting of
PAAQ(salt):polyvinyl chloride:dioctylphthalate:sodium tetraphenylborate of 1:33:66:1 (wt) but without
any traditional lead(II) ionophore achieved the optimal performance and exhibited a good Nernstian
response for Pb(II) ions over a wide concentration range from 2.5 106 to 0.1 M with a slope of
28.9 mV/decade and a detection limit down to 776 nM. A reasonably short response time of 12 s was
revealed together with a long lifetime over a period of around 4 months in a wide pH range between 2.8
and 5.2. A xed interference method indicated that the electrode has an excellent selectivity for lead(II)
ion over alkali, alkaline earth and other heavy metal ions. The proposed electrode has been also found to
be a powerful indicator electrode for potentiometric titration of Pb(II) ions with EDTA. The electrode can
be used to accurately monitor the Pb(II) pollution in environmental waters.
2009 Published by Elsevier B.V.

1. Introduction
Lead(II)ISE is of great signicance because the lead is ubiquitous in the environment and extremely hazardous to human health.
A large number of lead(II)ISEs reported are mainly based on polymeric membrane containing ionophores [16]. The key part of these
electrodes is a highly sensing ionophore having strong afnity for
a particular metal ion but giving low or even no response to others.
Enormous efforts have been made to design and synthesize suitable materials that are highly selective to lead(II) ions. Macrocyclic
crown ethers [24], calixarenes [57], and Schiff base [8,9], have
been widely investigated for this purpose. Other ligands such as
porphyrin [10,11], pyridinecarboximide [12], piroxicam [13], tetrabenzyl pyrophosphate [14], capric acid [15], phenyl disulde [16],
dithiodibenzoic acid [17], and quinaldic acid derivatives [18] are
also served for the fabrication of the lead(II)ISEs. In spite of availability of various lead(II)ISEs, the narrow working concentration

Corresponding authors at: Institute of Materials Chemistry, Key Laboratory of


Advanced Civil Engineering Materials of the Ministry of Education, College of Materials Science and Engineering, Tongji University, 1239 Si-Ping Road, Shanghai 200092,
China. Fax: +86 21 65980524.
E-mail addresses: adamxgli@yahoo.com (X.-G. Li), meironghuang@ymail.com
(M.-R. Huang).
0039-9140/$ see front matter 2009 Published by Elsevier B.V.
doi:10.1016/j.talanta.2008.11.045

range, high detection limit, slow response rate and especially poor
selectivity over interfering ions have restricted their widespread
application. In particular, the ionophores studied are almost all the
lipophilic compounds that easily leak out from matrix. Only one
report on the polymer as ionophores with good anti-leaking ability and then long lifetime was found [19]. However, the response
time is long because of its intrinsically high membrane resistance.
Therefore, it is of great challenge and signicance to further search a
new ionophore that possesses an intrinsically low membrane resistance and therefore high sensitivity and selectivity to lead(II) ions
so as to develop lead(II)ISEs with excellent detection performance
including quick responsibility and long duration.
In recent years, it is found that the conducting polymers especially some polymers from aromatic diamines have a unique ability
to form stable complexes with some heavy metal ions such as Pb(II),
Hg(II) and Ag(I) ions [2026]. Polyphenylenediamine synthesized
in our laboratory was found to possess a strong capability to adsorb
lead ions through complexation between Pb2+ ions and NH/ = N
groups in the macromolecular chains [23]. Concerning these properties, attempts have been made to employ these polymers to
extract and sense heavy metal ions. Carbon paste electrode modied with poly(1,8-diaminonaphthalene) was successfully used
to determine lead(II) ions in a concentration range from 0.2 to
10 M [21]. The detection process was complicated due to an indispensable combination of ion pre-enrichment and anodic-stripping

X.-G. Li et al. / Talanta 78 (2009) 498505

voltammogram, accompanying with a quite narrow work concentration range. Nevertheless, these results have indicated the
possibility of some aromatic conducting polymers with strong
afnity to lead(II) ions to act as ionophores in lead(II)ISEs. Unfortunately, no relevant reports have been found up to now.
A novel multifunctional polyaminoanthraquinone (PAAQ) was
successfully synthesized through a chemically oxidative polymerization [26]. PAAQ exhibits a strong adsorbability to lead ions owing
to its strong complexibility with lead ion through N and O coordination centers in the polymer chains. It seems that the PAAQ could be a
potential ionophore for construction of a new membrane sensor for
lead ions. The ingredient and thickness of the membrane were optimized to provide unique Pb2+ ISE that could result in reproducible,
noiseless and stable potentials. The excellent performance of the
electrode containing PAAQ as the sensing ionophore to lead(II) ions
in the determination of lead(II) ions has been elaborated for the rst
time.

499

10 m. A circular membrane of 16 mm diameter was carefully cut


out and glued to one end of plastic tube that would be lled with
0.10 M Pb(NO3 )2 solution as internal reference solution. The prepared electrodes were conditioned in a 0.01 M Pb(NO3 )2 solution
for 12 h and nally washed by distilled deionized water until stable
potentials were reached before using.
2.4. Potential measurement
All potentiometric measurements were performed by a PHS3C digital pH meter (Shanghai Kangyi Instruments Factory, China).
A double-junction saturated calomel electrode (SCE) was used as
the external reference electrode with the outer junction containing
0.10 M KCl and inner reference containing saturated KCl. The representative electrochemical cell for the electromotive force (EMF)
measurement is as follows:
Ag/AgCl|Internal solution (0.10 MPb(NO3 )2 )|Selective membrane

2. Experimental
2.1. Reagents
1-Aminoanthraquinone
(AAQ),
ammonium
persulfate
((NH4 )2 S2 O8 ), HClO4 (70%), acetonitrile, high molecular weight
polyvinyl chloride (PVC), dioctylphthalate (DOP), sodium
tetraphenyl borate (NaTPB), tetrahydrofuran (THF) and lead
nitrate of analytical reagent grade were commercially obtained
and used as received. 0.1 M Pb(NO3 )2 stock solution was prepared
by dissolving lead nitrate in distilled deionized water and standardized whenever necessary. The working solutions of different
concentrations were confected by gradually diluting the stock
solutions.
2.2. Synthesis of ne PAAQ microparticles as ionophore
PAAQ particles used as ionophore were simply prepared by a
chemically oxidative polymerization of the AAQ monomer [26]. A
typical procedure for the preparation of PAAQ particles is as follows: AAQ monomer (446 mg, 2 mmol) and HClO4 (11.8 M, 0.17 mL)
were added into acetonitrile (40 mL) in a 100 mL glass ask in a
water bath at 20 C and the mixture was then stirred vigorously for
30 min. (NH4 )2 S2 O8 (456 mg, 2 mmol) was dissolved separately in
deionized water of 0.75 mL to prepare an oxidant solution. The oxidant solution was then added dropwise into the monomer solution
at a rate of one drop (around 60 L) every 3 s. The reaction mixture
was stirred continuously for 24 h at 20 C. After reaction, the resulting polymer particles as precipitates were isolated from the reaction
mixture by ltration and washed with ethanol to remove the residual oxidant, remaining monomers and soluble oligomers. The PAAQ
particles were left to dry in ambient air at 50 C for 3 days. The
bluish black PAAQ particles obtained have nominal macromolecular structural formula in Scheme 1 and bulk electrical conductivity
of 5.0 105 S/cm at 15 C.

|Sample solution|Salt bridge(0.10 MKCl)|SCE


The performance of the electrodes was examined by measuring
the EMFs of the Pb2+ solutions in a concentration range of 108
to 101 M. The pH values were adjusted by HNO3 and NaOH when
considering the applicable pH range of the electrode. The response
time of the electrode was determined based on the potentials at different times for various concentrations of Pb(NO3 )2 solution until
the EMFs value kept constant in 5 min.
3. Results and discussion
3.1. Doping state of PAAQ microparticles
For the PAAQ ionophore of Pb2+ in this study, it is certain that the
N and O atoms could be soft coordination centers to complex Pb2+
ions. Besides, the as-prepared PAAQ salt particles contain a certain
amount of doped anions just like ClO4 and SO4 2 , which come
from the HClO4 and reducing products of (NH4 )2 S2 O8 used for the
polymerization. Both of them could be removed from the polymer
chains through dedoping process in ammonia, accompanying with
the signicant decline of conductivity. The reversible conversion
between salt and base states of the PAAQ chains is a unique nature of
aromatic conducting polymers [2023]. Concerning this, the electrodes based on PAAQ salt and base particles have been constructed
separately to study the effect of doping state on the response performance to Pb2+ ions. It is apparently observed from Fig. 1 that the
PAAQ salt particles-based electrode performs better response properties in all aspects, especially broader working concentration range

2.3. Electrode fabrication


To prepare a selective membrane, an appropriate amount of
PAAQ particles were dispersed in 5 mL THF by an intermittently
ultrasonic treatment. Proper amounts of PVC, DOP and NaTPB were
gradually added into the PAAQ dispersion and subsequently stirred
for 10 min to ensure a uniform mix. The mixture obtained thus
was poured on a smooth plate glass and then allowed to evaporate
for 24 h at room temperature. The translucent elastic membranes
with different compositions and thicknesses were obtained after
the evaporation of THF. The thickness of the membrane was measured by a roller type thickness gauge with the minimum scale of

Scheme 1. The macromolecular structural formula of PAAQ polymer together with


the exchange mechanism involved in producing potential signal.

500

X.-G. Li et al. / Talanta 78 (2009) 498505

Fig. 2. FTIR spectra of PAAQ ionophores (base and salt) before and after the exposure
of PAAQ in the 0.001 M Pb(NO3 )2 solution for 1 h.
Fig. 1. Potentiometric responses of Pb2+ selective electrodes based on PAAQ
microparticles at two doping states of salt and base at PAAQ:PVC:DOP:NaTPB of
1:33:66:1 and membrane thickness of 160 m.

from 4.0 106 to 0.1 M, higher slope (28.1 mV/decade), lower


detection limit (1.86 106 M), and shorter response time (14 s)
than the PAAQ base particles-based electrode (working concentration range: 1.0 105 to 0.1 M; slope: 27.3 mV/decade; detection
limit: 6.67 105 M; response time: 17 s). It should be noted that
the possibility of forming PbSO4 precipitate during detection process has been ruled out because possibly maximal SO4 2 and Pb2+
concentrations are 1011 M and 0.1 M, respectively, and the product
of SO4 2 and Pb2+ concentrations would be 1012 that is much lower
than the solubility product constant of PbSO4 (1.6 108 ). The
residual SO4 2 and ClO4 ions existing along the PAAQ salt chains
as the counter-anions are loose associated with protonated NH+
sites on the macromolecular chains, forming a negatively charged
layer around PAAQ chains. The static attraction of the negatively
charged layer to Pb2+ ions, as well as complexation between N/O
sites and Pb2+ , could hold the Pb2+ ions within the membrane and
therefore suppress the zero-current ux of the primary ions from
membrane into sample solution of low Pb2+ concentration, which
is considered as one important factor restricting the detection limit
[2729], ultimately leading to a low detection limit and broad linear range. On the other hand, the presence of SO4 2 and ClO4 in
the PAAQ salts would prevent the extraction of anions and then the
primary ions from the inner reference solution, i.e., the Pb2+ in the
inner reference solution would not be extracted into the membrane
phase. As we know, the absence of primary ion ux is necessary
to achieve low detection limit. Generally, the primary ions always
tend to transit from bulk membrane into the sample solution at
a low Pb2+ concentration, leading to a local higher concentration
than the sensing concentration. Consequently, true concentration
of sample solution could not be satisfactorily detected. Vigorous
agitation could improve the homogeneity of the local concentration
to some extent. Additionally, much higher electrical conductivity
of PAAQ(salt) (5.0 105 S/cm) than PAAQ(base) (3.7 108 S/cm)
would remarkably enhance the sensitivity of electrode as well, contributing to a nearer Nernstian slope and quicker response.
Fig. 2 shows the FTIR spectra of the ionophores of PAAQ base
and salt before and after the exposure of PAAQ in 0.001 M Pb(NO3 )2
solution for 1 h. The peaks at 3440, 1640 and 1270 cm1 due to
the NH, C O and CN, respectively, become weaker after contacting Pb2+ ions, indicating an interaction between NH/C O groups
and Pb2+ ions, i.e., complexation between the Pb2+ ions and the

PAAQ. Furthermore, the PAAQ saltPb2+ complex illustrates weaker


NH absorbance at 3440 cm1 than the PAAQ basePb2+ complex,
suggesting stronger interaction between the PAAQ salt and Pb2+
ions. Therefore, the as-prepared PAAQ salt particles with stronger
response to Pb2+ ions are selected as lead(II) ionophore in the following study.
3.2. PAAQ microparticle content
It is well known that the sensitivity and selectivity of ISEs depend
signicantly on the nature of ionophore and the membrane composition. Thus the loading content of ionophore should be optimized
to produce the best performance of the proposed electrode. The
complexation function of ionophore with Pb2+ cannot be visualized sufciently at low ionophore content. On the contrary, high
ionophore content is unfavorable to the transportation of ions in
membrane and may even cause pinholes in membrane, leading to
a deterioration of performance ultimately. To study the effect of
ionophore content, three electrodes loading different amounts of
PAAQ ionophore were prepared and their potential responses are
shown in Fig. 3 and Table 1. Interestingly, PAAQ:PVC:DOP:NaTPB

Fig. 3. Potentiometric responses of the Pb(II) ISEs based on different contents of


PAAQ in a 160 m-thick matrix membrane.

X.-G. Li et al. / Talanta 78 (2009) 498505

501

Table 1
Composition and performance characteristics of Pb(II) ISEs in Fig. 3.
PAAQ:PVC: DOP:NaTPB (weight ratio)

Working concentration range (M)

Working linear equation

1:25:50:1
1:33:66:1
1:50:100:1

1.0 105

E = 129.6 + 26.7 log[Pb2+ ]

to 0.1
4.0 106 to 0.1
5
1.0 10 to 0.024

E = 126.1 + 28.1 log[Pb2+ ]


E = 112.6 + 27.9 log[Pb2+ ]

of 1:33:66:1 (wt) exhibits the best performance involving the


widest working concentration range from 4.0 106 to 0.1 M with
the largest Nernstian slope of 28.1 mV/decade, the lowest detection limit of 1.86 106 M, and the shortest response time of
14 s. This great improvement of electrode performance should
originate from the presence of PAAQ ionophore. In fact, the
PAAQ:PVC:DOP:NaTPB (0:33:66:1) membrane with the thickness
of 160 m has much lower potential stability and lower conductivity (2.13 108 S/cm) than the PAAQ:PVC:DOP:NaTPB (1:33:66:1)
membrane (1.35 107 S/cm). The PAAQ-free electrode might not
be suitable to detect potential response. It is reported that the
addition of only a small amount of conducting polymers could
signicantly enhance the electrical and other properties of matrix
polymers [24,25]. Excellent performance has also been realized at
a high ionophore content for some PVC membrane ISEs based on
hexathia-18-crown-6-tetraone ionophore:PVC of 1:6(wt) [2] and
N,N -dimethylcyanodiaza-18-crown-6 ionophore:PVC of 1:5(wt)
[4]. However, the optimal weight ratio of PAAQ ionophore and
PVC is 1:33 for most electrodes involving the PAAQ-based electrode in this study. Much less PAAQ particles are required to
produce the best performance of the ISEs, i.e., PAAQ might be more
efcient ionophore than hexathia-18-crown-6-tetraone and N,N dimethylcyanodiaza-18-crown-6.
3.3. Nature of plasticizer
The plasticizer is considered to play an important role in optimizing the performance of ISEs through inuencing the dielectric
constant of the membrane phase [57]. Four plasticizers of different
polarity were used to study the effect of plasticizer on the selective response of the Pb2+ ISEs based on PAAQ. As shown in Fig. 4,
the ISE with DOP as plasticizer obviously performs a higher Nernstian response slope of 28.1 mV/decade over a wider range from

Fig. 4. Potentiometric responses of Pb(II) ISEs based on PAAQ microparticles with


different plasticizers at PAAQ:PVC:plasticizer:NaTPB of 1:33:66:1 and membrane
thickness of 160 m.

Slope (mV/decade)

Detection limit (M)

Response time (s)

26.7
28.1
27.9

3.29 106

16
14
17

1.86 106
2.92 106

4.0 106 to 0.1 M. The ISEs based on the other three plasticizers
only give a diminished response slope (<27 mV/decade) over narrow working concentration range. Therefore, DOP was proved to be
the most effective plasticizer in preparing the proposed Pb2+ ISEs
based on PAAQ.
3.4. Lipophilic anion additive content
The presence of lipophilic anion additive in cation-selective
membrane electrodes can not only reduce the ohmic resistance and
anion interference but also improve the sensitivity and selectivity
[11]. The potential responses of the Pb2+ ISEs containing lipophilic
anion additive (NaTPB) were investigated to examine the effect
of additive content. It can be seen from Fig. 5 that the potential
response of the Pb2+ ISE without NaTPB gives the weakest sensitivity with a diminished slope of 25.3 mV/decade over a narrow linear
range from 105 to 102 M. The sensitivity can be greatly improved
by the addition of NaTPB. The Pb2+ ISE with PAAQ:NaTPB of 1:1
(wt) displays the best response characteristics with a higher Nernstian response slope of 28.1 mV/decade over a wider linear range
of 4.0 106 to 0.1 M and a lower detection limit of 1.86 106 M.
Higher content of additive (PAAQ:NaTPB = 1:1.5) is also unfavorable
for the improvement of sensitivity. More additive may deteriorate
the mechanical property of membrane and the lifetime of electrode. Therefore, the optimal lipophilic additive content is 100 wt%
relative to the ionophore.
3.5. Thickness of the membrane containing PAAQ microparticles
Three electrodes with different membrane thicknesses but
the same membrane composition of PAAQ:PVC:DOP:NaTPB = 1:33:
66:1 were fabricated to examine the effect of membrane thickness.
As shown in Fig. 6 and Table 2, a signicant improvement is achieved
in the response characteristics of electrodes including linear range,

Fig. 5. Potentiometric responses of Pb(II) ISEs based on PAAQ microparticles with


different NaTPB contents at PAAQ:PVC:DOP of 1:33:66 and membrane thickness of
160 m.

502

X.-G. Li et al. / Talanta 78 (2009) 498505

Table 2
Performance characteristics of Pb(II) ISEs in Fig. 6.
Membrane thickness (m)

Working concentration range (M)

Working linear equation

220
160
90

1.0 105

E = 120.5 + 28.7 log[Pb2+ ]

to 0.1
4.0 106 to 0.1
6
2.5 10 to 0.1

E = 126.1 + 28.1 log[Pb2+ ]


E = 132.5 + 28.9 log[Pb2+ ]

Fig. 6. Potentiometric responses of the Pb(II) ISEs based on different membrane


thicknesses at PAAQ:PVC:DOP:NaTPB of 1:33:66:1 (wt).

detection limit, slope and response time with decreasing the membrane thickness. At the smallest membrane thickness of 0.09 mm,
the electrode exhibits an almost Nernstian slope of 28.9 mV/decade
and a widest working concentration range from 2.5 106 to 0.1 M,
which has been broadened nearly by four times compared with that
at the thickness of 0.22 mm. The electrode also possesses the lowest
detection limit of 7.76 107 M and the shortest response time of
12 s. It seems that the optimal PAAQ membrane exhibits comparable or even better performance than 10 representative lead(II)ISEs
with PVC matrix in Table 3.
It is reported that the membrane thickness of the PVC-based
Pb2+ ISEs is mostly between 0.2 and 0.5 mm [1,10,11,27,28]. The
sensing unit supported by a highly porous polymeric layer in these
electrodes is practically liquid state comprising a lipophilic organic
ionophore dissolved in oil phase such as DOP. The detection limits
of these electrodes commonly keep at the order of 106 or 107 M,

Slope (mV/decade)

Detection limit (M)

Response time (s)

28.7
28.1
28.9

4.47 106

15
14
12

1.86 106
7.76 107

beyond which the potential response would deviate from the Nernstian equation, because a zero-current ux of the primary ions from
the membrane into sample solutions would result in ion activity
of the primary ions maintaining as high as 106 or 107 M in the
local domain near the membrane surface at the sample side. The
ux of primary ions is generated by a transmembrane concentration gradient which occurs when the concentrations of the sample
and inner solutions are not the same. The detection limit could be
improved through reducing the transmembrane concentration gradient by adjusting the concentration of the inner solution to the
value of sample solution, decreasing the total ion concentration,
or increasing the thickness of the membrane [29]. However, the
result here is reverse to the above-mentioned circumstance. The
potential response characteristics of the proposed electrodes can be
enhanced by decreasing the thickness of the membrane. This may
be relevant to the existing state of the ionophore. The ne PAAQ
particle ionophore that is insoluble in DOP oil phase, disperses uniformly in the PVC membrane as the solid state. The transporting
rate of lead(II) ions in solid phase is much slower and more arduous than that in liquid phase. Hence the ux of lead(II) ions is not
encouraged in solid membrane. Nevertheless, the response is still
very fast due to the relatively thin membrane, as discussed below.
Considering the remarkable decrease of membrane strength, it is
not suggested to further reduce the membrane thickness. All further
detailed studies were carried out on the electrode with the composition of PAAQ:PVC:DOP:NaTPB (1:33:66:1) and the membrane
thickness of 0.09 mm.
Another result which disagrees with the theory that increasing membrane thickness might improve the detection limit was
also reported in the investigation on the electrode based on
diphenylmethyl-N-phenylhydroxamic acid ionophore [30]. Among
three styrene/acrylonitrile copolymer membranes with different
thicknesses of 0.14, 0.21, and 0.45 mm, the 0.21 mm-thick membrane demonstrates the lowest detection limit.
3.6. Effect of pH
The dependence of the potentiometric response of the proposed
ISE on the pH value of the Pb2+ solution was tested at three Pb2+

Table 3
Comparison of performance characteristics of the proposed electrode with the PVC membrane lead(II) ISEs reported in literature.
Ionophore

Working concentration
range (M)

Slope (mV/decade)

Detection limit (M)

Response time (s)

Lifetime (month)

Refs.

N,N -Dibenzyl-1,4,10,13-tetraoxa7,16-diazacyclooctadecane
Hexathia-18-crown-6-tetraone
2 -Methoxyethoxyl
sym-dibenzo-16-crown-5-ether
Tetrakis(p-carboxyphenyl)azo]-8tetrahydroxy
calix[4]arene
Meso-tetrakis(2-hydroxy-1naphthyl)
porphyrin
Chiral 2,6-bis-pyridinecarboximide
Piroxicam
Tetrabenzyl pyrophosphate
Capric acid
Phenyl disulde
Solid PAAQ particles

8.2 106 to 0.1

30.0

8.2 106

10

[1]

1.0 106 to 0.008


5.0 105 to 0.5

29.0
28.9

8.0 107
1.0 106

40
30

2
3

[2]
[3]

1.0 106 to 0.1

29.4

8.0 107

20

[6]

3.2 105 to 0.1

29.2

3.5 106

10

[11]

3.5 106 to 0.01


1.0 105 to 0.1
1.0 105 to 0.01
1.0 105 to 0.01
2.0 106 to 0.01
2.5 106 to 0.1

27.9
30.0
28.7
29.0
29.3
28.9

2.2 106
4.0 106
3.0 106
6.0 106
1.2 106
7.8 107

25
45
10
15
45
12

0.5
3
0.7
3
1.7
4

[12]
[13]
[14]
[15]
[16]
This study

X.-G. Li et al. / Talanta 78 (2009) 498505

Fig. 7. The pH dependence of the PAAQ:PVC:DOP:NaTPB (1:33:66:1) membrane


electrode with the thickness of 90 m on the potentiometric response under three
constant concentrations of lead ion.

concentrations (1.0 102 , 1.0 103 and 1.0 104 M) over the pH
range between 1 and 6. It is seen from Fig. 7 that the potential
response remains almost constant over the pH range from 2.8 to
5.2, beyond which a gradual change in potential can be observed.
As a result, this range can be taken as the working pH range of the
proposed electrode. The declined potential at higher pH values may
be ascribed to the formation of some hydroxy complexes of Pb2+
such as Pb(OH)+ and Pb(OH)2 , leading to a decreased Pb2+ concentration, while at lower pH, the abundant H+ ions can protonate the
N atoms of PAAQ and even cause the decomplexation of Pb2+ PAAQ
complex. The H+ ions itself can make interference to the electrode
response simultaneously. Both of them could result in the rise of
the potential.
The wide working pH range of the proposed electrode is comparable to the reported Pb2+ ISEs based on the following ionophores:
[1],
N,N -dibenzyl-1,4,10,13-tetraoxa-7,16-diazacyclooctadecane
hexathia-18-crown-6-tetraone [2], phosphorylated calyx[4]arene
[31], and 5,11-dibromo-25,27-dipropoxycalix[4]arene [32]. The
lower limit of the working pH range is expected to be further
improved in case that the binding ability of Pb2+ onto PAAQ
chains is strengthened. The electrode based on 2,12-dimethyl7,17-diphenyltetrapyrazole with strong complex capability shows
wider working pH range of 1.66.0, whereas the electrode based
on 5,11-dibromo-25,27-dipropoxycalix[4]arene with a relatively
weaker complex capability has narrower pH range of 2.36.0 [32].
3.7. Response and lifetime of the electrode
The response time of the PAAQ ISE was determined by measuring the time required to achieve a steady potential in Pb2+
solution with three concentrations of 1.0 102 , 1.0 103 and
1.0 104 M. It is found from Fig. 8 that the response time
is 12 s, approaching to the fastest response of Pb2+ ISEs based
on N,N -dibenzyl-1,4,10,13-tetraoxa-7,16-diazacyclooctadecane[1]
and N,N -bis(salicylidene)-2,6-pyridinediamine [33] as ionophores.
This fast response is relevant to the fast kinetic process of complexation between Pb2+ and PAAQ ionophore in the doped state. On
the other hand, the ne particles of the PAAQ salts, together with
their higher intrinsic conductivity (5.0 105 S/cm), may promote
the formation of a uniform, thin, and highly conducting composite membrane. All of these features are benecial to the transition
of charges in membrane, resulting in the fast response of the electrode. There was no difference for the potential responses for each

503

Fig. 8. Response time prole of the Pb(II) ISE based on PAAQ:PVC:DOP:NaTPB


(1:33:66:1) membrane with the thickness of 90 m.

concentration recorded from high concentration to low concentration and from low concentration to high concentration, indicating
that the ISE has an excellent reversibility. The standard deviations
of ten replicate potential measurements of 1 mM and 0.1 mM sample solutions were 0.58 and 0.65 mV, respectively, indicating the
good reproducibility of the prepared ISE. It is also observed that
the electrode could be satisfactorily used over a period of 4 months
without any signicant loss of the performance characteristics such
as working concentration range, slope and response time. Dislike
the ordinary ionophore, PAAQ salts existing as ne solid particles do
not leak from the matrix while contacting with aqueous solution,
ensuring a reasonably long lifetime of the proposed electrode.
3.8. Electrode selectivity
The selectivity is one of the most important characteristics for
selective electrodes, because it determines the feasibility and utility of the electrodes in real sample analysis. The potentiometric
selectivity coefcient K POT
has been served to reect the rela2+
Pb

,B

tive response of the membrane selective electrode for the primary


ion over other interfering ions present in solution and can be used
to predict response functions in mixed samples. It can be determined mainly through three different methods including separate
solution, xed interference and matched potential methods. The
xed interference method is preferable since it more closely mimics a practical application of the ISE. In the present study, the xed
interference method was employed to assess the selectivity of the
fabricated Pb2+ ISE over other commonly interfering ions. The selectivity coefcients were calculated through the equation:
K POT
2+
Pb

,B

= aPb2+ (DL)/(aB )2/z

where Pb2+ (DL) is the detection limit of Pb2+ ion, B the activity
of the interfering ion, and z the charge of the interfering ion [34]. It
is noteworthy that the selectivity coefcients for all diverse cations
listed in Table 4 are in the order of 102 or 103 , indicating that
the proposed electrode is highly selective over all the interfering
ions studied. The interferences of alkali metal ions (Na+ and K+ ) and
alkaline-earth metal ions (Ba2+ and Ca2+ ) are almost negligible with
the selectivity coefcients in the order of 103 . Some heavy metal
ions and noble metal ions (Cu2+ , Hg2+ , Au3+ and Ag+ ) might cause a
weak interference, which may arise from the complex capability of
PAAQ to some transition metal ions. It is discovered that aromatic
amine polymers can complex Hg2+ and Ag+ ions as well as Pb2+
ions through the NH2 and NH groups in the chains [2225].

504

X.-G. Li et al. / Talanta 78 (2009) 498505

Table 4
Potentiometric selectivity coefcients of the PAAQ salt particles-based Pb2+ ISE
obtained by the xed interference method at an interfering ion concentration of
1.0 103 M.
Interfering ions

Selectivity coefcient K POT


2+

Hg2+
Ag+
Cu2+
Au3+
NH4 +
Na+
K+
Ba2+
Ca2+

8.51 102
6.97 102
5.82 102
4.68 102
7.59 103
7.18 103
4.27 103
3.55 103
3.09 103

Pb

,B

Nevertheless, with the selectivity coefcients in the order of 102 ,


these metal ions could not make an obvious disturbance to the
determination of Pb2+ ions. Even so, it should be noted that much
more severe interference from Ag+ , Cu2+ , Hg2+ and some other
cations was always found in other Pb2+ ISEs [9,11,14,27,31].
3.9. Application of the electrode
The analytical application of the electrode was investigated
as an indicator electrode for the potentiometric estimation of
1.0 103 M Pb2+ solution by titrating with two EDTA concentrations of 1.0 103 M and 1.0 102 M. The titration plots presented
in Fig. 9 do not show a standard sigmoid curve, probably due to
some interference caused by Na+ ions from disodium EDTA salt.
When Pb2+ solution was titrated by EDTA, the Pb2+ concentration
decreased, accompanying by an increased concentration of Na+ ions
from EDTA with the titration proceeding. In the later period, the
amount of Na+ ions can greatly surpass the amount of Pb2+ ions,
causing interference to the potential. Similar titration plots were
observed when the ISEs based on N,N -dibenzyl-1,4,10,13-tetraoxa7,16-diazacyclooctadecane [1], 4-t-butylcalix[4]arene [35], and
4-t-butylcalix[6]arene [36] were served as indicator electrodes. The
stoichiometry of Pb2+ EDTA complex can be judged from the sharp
break point in Fig. 9. It must be appreciated that the potentiometric
titration of Pb2+ also performs well when the EDTA concentration
is in the same order of Pb2+ . Under this circumstance, the error
of reading of volume of EDTA solution used could be diminished

Table 5
Recovery studies of the PAAQ-based Pb2+ ISE on the detection of lead(II) in tap and
rain waters.
Water sample

Added Pb2+ (M)

Detected Pb2+ (M)

Recovery (%)

Tap water 1
Tap water 2
Rain water 1
Rain water 2

50
200
50
200

48.9
205
51.5
195

97.8
102.5
103.0
97.5

to some extent as compared to the titration just consuming ca.


1 mL titrant of much higher EDTA concentration than the Pb2+ concentration [3036]. Obviously, the prepared electrode in this study
shows the practical applicability as an indicator electrode in the
potentiometric titration to Pb2+ solutions.
Since no Pb2+ traces in tap water and rainwater samples were
detected by the prepared Pb2+ ISE based on PAAQ, the electrode
was applied to detect Pb2+ concentration in the water samples
by standard addition method. To prepare the tap water samples,
the collected tap water was rst boiled for about 5 min to remove
Cl2 . The pH values of the tap water and rainwater samples were
both adjusted to around 4.5 by HNO3 . The ve-replicate detection
and recovery results are presented in Table 5. It can be seen that
the Pb2+ ISE performs satisfactorily with the reasonable recovery,
indicating the feasibility of the electrode in the detection of Pb2+
concentration in the environmental water like tap and rainwaters.
4. Conclusions
The ne particles of aromatic PAAQ in PVC membrane well perform as a novel Pb2+ sensing ionophore with strong sensitivity and
high selectivity because of their intrinsic electroconductivity and
strong afnity towards Pb2+ ions. The Pb2+ ISEs based on the particles rather than traditional ionophores are high-performance ISEs
with wide working range from 2.5 106 to 1.0 101 M, low detection limit down to 7.76 107 M, and long lifetime. The response
time of shorter than 12 s has approached to the fastest response
rate of the existing electrodes. 1% PAAQ particles-containing ISE
demonstrates much better performance than the particles-free ISE.
In particular, the proposed electrode shows an excellent selectivity
for Pb2+ over alkali, alkaline earth, and heavy metal ions. And the
PAAQ ionophore is easily available since the PAAQ particles can be
facilely and productively synthesized through chemically oxidative
polymerization of aminoanthraquinone. Thus the prepared electrode with a good combination of excellent performance, and long
lifetime as well as low cost has a promising application in the
determination of Pb2+ ions. The electrode can be used to accurately
monitor the Pb(II) pollution in environmental waters just like tap
and rain waters. The results in this report have indicated the great
potential of other aromatic conducting polymers with strong ability to chelate heavy metal ions to act as new ionophores in ISEs and
thus the relevant research should be highly encouraged.
Acknowledgements
The project is supported by the National Natural Science Foundation of China (20774065) and the Foundation of Key Laboratory
of Molecular Engineering of Polymers, Fudan University, China.
References

Fig. 9. Potentiometric titration plots of l.0 103 M Pb(NO3 )2 solution of 10 mL


with 1.0 102 M and 1.0 103 M EDTA, respectively by the ISE based on
PAAQ:PVC:DOP:NaTPB (1:33:66:1) membrane with the thickness of 90 m.

[1]
[2]
[3]
[4]

V.K. Gupta, A.K. Jain, P. Kumar, Sens. Actuators B 120 (2006) 259.
M. Shamsipur, M.R. Ganjali, A. Roihollahi, Anal. Sci. 17 (2001) 935.
C.C. Su, M.C. Chang, L.K. Liu, Anal. Chim. Acta 432 (2001) 261.
M.R. Ganjali, M. Hosseini, F. Basiripour, M. Javanbakht, O.R. Hashemi, M.F. Rastegar, M. Shamsipur, G.W. Buchanen, Anal. Chim. Acta 464 (2002) 181.
[5] F. Cadogan, P. Kane, M.A. McKervey, D. Diamond, Anal. Chem. 71 (1999) 5544.
[6] J.Q. Lu, R. Chen, X.W. He, J. Electroanal. Chem. 528 (2002) 33.

X.-G. Li et al. / Talanta 78 (2009) 498505


[7] L.X. Chen, J. Zhang, W.F. Zhao, X.W. He, Y. Liu, J. Electroanal. Chem. 589 (2006)
106.
[8] H. Kim, H.K. Lee, A.Y. Choi, S. Jeon, Bull. Korean Chem. Soc. 28 (2007) 538.
[9] Z. Pilehvari, M.R. Yaftian, S. Rayati, M. Parinejad, Anal. Chim. 97 (2007) 747.
[10] W.J. Zhang, C.Y. Li, X.B. Zhang, Z. Jin, Anal. Lett. 40 (2007) 1023.
[11] H.K. Lee, K. Song, H.R. Seo, Y.K. Choi, S. Jeon, Sens. Actuators B 99 (2004)
323.
[12] S.S.M. Hassan, M.H.A. Ghalia, A.G.E. Amr, A.H.K. Mohameda, Talanta 60 (2003)
81.
[13] S. Sadeghi, G.R. Dashti, M. Shamsipur, Sens. Actuators B 81 (2002) 223.
[14] D.F. Xu, T. Katsu, Talanta 51 (2000) 365.
[15] M.F. Mousavi, M.B. Barzegar, S. Sahari, Sens. Actuators B 73 (2001) 199.
[16] A. Abbaspour, B. Khajeh, Anal. Sci. 18 (2002) 987.
[17] M.B. Gholivand, A. Mohammadi, Chem. Anal. (Warsaw) 48 (2003) 305.
[18] M. Casado, S. Daunert, M. Valiente, Electroanalysis 13 (2001) 54.
[19] M.R. Ganjali, A. Rouhollahi, A.R. Mardan, M. Hamzeloo, A. Mogimi, M. Shamsipur, Microchem. J. 60 (1998) 122.
[20] X.G. Li, M.R. Huang, W. Duan, Y.L. Yang, Chem. Rev. 102 (2002) 2925.

505

[21] S. Majid, M.E. Rhazi, A. Amine, A. Curulli, G. Palleschi, Microchim. Acta 143
(2003) 195.
[22] X.G. Li, R. Liu, M.R. Huang, Chem. Mater. 17 (2005) 5411.
[23] M.R. Huang, Q.Y. Peng, X.G. Li, Chem. Eur. J. 12 (2006) 4341.
[24] Q.F. L, M.R. Huang, X.G. Li, Chem. Eur. J. 13 (2007) 6009.
[25] X.G. Li, H. Li, M.R. Huang, Chem. Eur. J. 13 (2007) 8884.
[26] X.G. Li, H. Li, M.R. Huang, China Patent CN1810852, 2008.
[27] M.F. Mousavi, S. Sahari, N. Alizadeh, M. Shamsipur, Anal. Chim. Acta 414 (2000)
189.
[28] A. Rouhollahi, M.R. Ganjali, M. Shamsipur, Talanta 46 (1998) 1341.
[29] Z. Szigeti, T. Vigassy, E. Bakker, E. Pretsch, Electroanalysis 18 (2006) 1254.
[30] K.C. Gupta, M.J. DArc, IEEE Sens. J. 1 (2001) 275.
[31] M.R. Yaftian, S. Rayati, D. Emadi, D. Matt, Anal. Sci. 22 (2006) 1075.
[32] A.K. Jain, V.K. Gupta, L.P. Singh, J.R. Raisoni, Electrochim. Acta 51 (2006) 2547.
[33] T. Jeong, H.K. Lee, D.C. Jeong, S. Jeon, Talanta 65 (2005) 543.
[34] E. Bakker, E. Pretsch, P. Bu1hlmann, Anal. Chem. 72 (2000) 1127.
[35] V.K. Gupta, R. Mangla, S. Agarwal, Electroanalysis 14 (2002) 1127.
[36] V.S. Bhat, V.S. Ijeri, A.K. Srivastava, Sens. Actuators B 99 (2004) 98.

Você também pode gostar