Analytica Chimica Acta 578 (2006) 59–74
Prospects of conducting polymers in biosensors
Bansi D. Malhotra a,∗ , Asha Chaubey b , S.P. Singh a
a
Biomolecular Electronics and Conducting Polymer Research Group, National Physical Laboratory,
Dr. K.S. Krishnan Marg, New Delhi 110012, India
b Regional Research Laboratory, Jammu, India
Received 18 February 2006; received in revised form 17 April 2006; accepted 20 April 2006
Available online 29 April 2006
Abstract
Applications of conducting polymers to biosensors have recently aroused much interest. This is because these molecular electronic materials
offer control of different parameters such as polymer layer thickness, electrical properties and bio-reagent loading, etc. Moreover, conducting
polymer based biosensors are likely to cater to the pressing requirements such as biocompatibility, possibility of in vivo sensing, continuous
monitoring of drugs or metabolites, multi-parametric assays, miniaturization and high information density. This paper deals with the emerging
trends in conducting polymer based biosensors during the last about 5 years.
© 2006 Elsevier B.V. All rights reserved.
Keywords: Conducting polymer; Biosesnor; Soliton; Polaron; Bipolaron; Enzyme; Glucose oxidase; Urease; Cholesterol oxidase; Deoxyribonucleic acid (DNA);
Peptide nucleic acid (PNA); Microorganism; Amperometric; Potentiometric; Immunosensor
1. Introduction
Biosensors have recently attracted much interest. This is
because these interesting bio-devices have been shown to have
applications in clinical diagnostics, environmental monitoring,
food freshness and bioprocess monitoring [1–249]. A number
of materials such as polymers, sol–gels and conducting polymers have been used to improve the stability of the biomolecules
used in the fabrication of the desired biosensors. In this context,
polymers have become the materials of choice for recent technological advances in biotechnology. Initially polymers were
thought to be insulators and were utilized in various electrical
and electronic devices. The discovery of poly (sulphur nitride)
[(SN)x ] which becomes superconducting at low temperatures
led to the renewed interest in polymers [10]. The electrical
conductivity of [(SN)x ] can be enhanced by several orders,
i.e. 105 S cm−1 by simple doping with oxidizing agents, e.g.
I2 , AsF5 , NOPF6 (p-doping) or reducing agents (n-doping),
e.g. sodium napthalide. Subsequently, a new class of polymers
such as poly-para-phenylene (PPP), polyphenylene sulphide
(PPS), polythiophenes and polypyrroles (PPy) were reported
∗
Corresponding author. Tel.: +91 11 25734273; fax: +91 11 25726938.
E-mail address:
[email protected] (B.D. Malhotra).
0003-2670/$ – see front matter © 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.aca.2006.04.055
[11–13]. Diaz et al. produced coherent films of PPy with conductivity of 100 S cm−1 and this conducting polymer exhibits
excellent air stability [14]. Chen et al. have estimated the concentration of cytochrome C using electrochemically prepared
conducting polymers based on ferrocene substituted thiophene
and terthiophene [44]. Many other conducting polymers such as
poly(3,4-ethylenedioxythiophene) (PEDOT), polyfuran, polyindole, polycarbazole, polyaniline, etc. have been synthesized and
studied extensively [12–16,45].
Compared to saturated polymers, conducting polymers have
different electronic structures. Chemical bonding in conducting polymers provides one unpaired electron, i.e. electron
per carbon atom in the backbone of the polymer. Carbon atoms
are in sp2 pz configuration in bonding and orbitals of successive carbon atoms overlap providing delocalization of electrons
along the backbone of polymer [12]. This delocalization provides the charge mobility along the backbone of the polymer
chain and induces unusual properties such as electrical conductivity, low ionization potential, low energy optical transitions
and high electron affinity. The bonds in conjugated polymers
are highly susceptible to chemical or electrochemical oxidation
or reduction. The origin of electrical conduction in conducting polymers has been ascribed to the formation of non-linear
defects such as solitons, polarons or bipolarons formed either
during doping or polymerization of a monomer [19,46,47,57].
60
B.D. Malhotra et al. / Analytica Chimica Acta 578 (2006) 59–74
The conductive and semiconducting properties of these polymers make them an important class of materials for a wide
range of electronic, optoelectronic and biotechnological applications such as in rechargeable batteries, molecular electronics,
electronic displays, solar cells, ion exchange membrane in fuel
cells, diodes, capacitors, field-effect-transistors, printed circuit
boards, chemical sensors, drug release systems and biosensors,
etc. [14–17]. It is being projected that conducting polymers
can be used to transport small electronic signals in the body,
i.e. act as artificial nerves. Perhaps modifications to the brain
may eventually be contemplated. Scientists have used films
in a neurotransmitter as a drug release system into the brain
[18].
Conducting polymers have emerged as potential candidates
for biosensors. Gerard et al. have reviewed the literature on applications of conducting polymers to biosensors [57]. Geetha et al.
[77] have discussed the applications of conducting polypyrrole
to drug delivery. Andreescu and Sadik have reviewed the challenges and trends in biosensors for environmental and clinical
monitoring [78]. Cosnier [79,80] has discussed the analytical
applications of affinity biosensors based on electropolymerized
films. Ramaniviciene and Ramanavicius have reported an interesting overview on the potential use of conducting polymers
as electrochemical based affinity biosensors [81]. Malinauskas
et al. have reviewed the electrochemical aspects of conducting polymer-based nano-structured materials for application to
super-capacitors, energy conversion systems, batteries and sensors [82]. Wanekaya et al. have reviewed recent advances in
biosensors based on one-dimensional (1-D) nanostructures [83].
Ming-Hung Lee et al. have highlighted the current developments
of DNA-based bioanalytical microsystesm for point-of-care
diagnostics [84]. Adhikari and Majumdar have discussed the
role of non-conducting and intrinsically conducting polymers in
sensor devices [56]. Terry et al. [86] have assessed the future
and current trends of biosensors in food industry. Drummond
et al. have discussed numerous approaches to electrochemical
detection based on modified electrodes, electrochemical amplifications with nanoparticles and electrochemical devices using
DNA-mediated charge transport chemistry and electrochemistry
of DNA-specific redox reporters [87]. Habermuller et al. have
reported on the various electron-transfer mechanisms operating
in amperometric biosensors [88]. Kerman et al. have predicted
that electrochemical DNA biosensors with suitable microfabrication techniques are likely to be increasingly popular in the
near future [89]. The present paper focuses on the prospective
applications of conducting polymers in biosensors.
2. Importance of conducting polymers to biosensors
Conducting polymers [48–114] have attracted much interest as suitable matrices of biomolecules and can be used
to enhance stability, speed, sensitivity and hence are finding
increasing use in medical diagnostics [20–23,51–96]. A number of techniques such as physical adsorption, electrochemical
entrapment and covalent attachment based on ethyl-dimethylaminopropylcarbodiimide (EDC) and N-hydroxy-succinimide
(NHS) coupling chemistry, have been used to improve the sta-
bility of the desired biomolecules onto conducting polymers
[57,80,170,171,242–245]. Schuhmann et al. reported that functionalization of conducting polypyrrole films provides suitable
surface for covalent linkage of enzymes after carbodiimide
activation [224]. Cosnier et al. have discovered that oxidative
electropolymerization of a new (dicarbazole) derivative functionalized by N-hydroxysuccinimide group in acetonitrile results
in electroactive poly (dicarbazole) films. Chemical functionalization of the poly (dicarbazole) film could be easily performed
by successive immersions in aqueous and mediator solutions
[170]. Devouge et al. have reported a practical molecular clip
for immobilization of receptors and biomolecules on device
surface based on the photographting of o-succinimidyl-4-(pazido-phenyl)butanoate [241]. Liu et al. have constructed a
bio-electrode in which a self-assembled monolayer containing
a novel norbornylogous bridge by covalently attaching to flavin
adenine dinucleotide (FAD), the redox active centre of several
oxidase enzymes [243]. It has been proposed that a decrease
in the electronic coupling between the redox active FAD and
the electrode following reconstitution of the glucose oxidase is
responsible for the inability of the enzyme to be turned over
under anaerobic conditions.
The techniques of incorporating biomolecules into electrodepositable conducting polymeric films permit the localization
of biologically active molecules on electrodes of any size or
geometry and are particularly appropriate for the fabrication of
multi-analyte micro-amperometric biosensors [24]. Conducting
polymers can act as transducers in biosensors. A transducer converts a biochemical signal resulting through the interaction of a
biological component into an electronic signal. Suitable transducing systems can be adapted in a sensor assembly depending
on the nature of the biochemical interaction with the species
of interest. The physical transducers vary from electrochemical, spectroscopic, thermal, piezoelectric and surface wave
technology [25,26]. The most common electrochemical transducers being utilized are amperometric and potentiometric. An
amperometric biosensor measures the current produced during
the oxidation or reduction of a product or reactant at a constant applied potential. Such sensors have fast response times
and good sensitivities. However, the excellent specificity of a
biological component can be compromised by the partial selectivity of the bio-electrode. This lack of specificity requires sample preparation, separation or some compensation for interfacing signals. Potentiometric biosensors relate electrical potential
to the concentration of analytes by using ion-selective electrodes or gas-sensing electrodes as the physical transducers
[27,28].
Electrically conducting polymers are known to have considerable flexibility in chemical structures, which can be modified
as required. By chemical modeling and synthesis, it is possible
to modulate the required electronic and mechanical properties of
conducting polymers. Moreover, the polymer itself can be modified to bind protein molecules [29–31,116]. Another advantage
offered by conducting polymers is that the electrochemical synthesis allows direct deposition of a polymer on the electrode
surface while simultaneously trapping the protein molecules
[32,33]. It is thus possible to control the spatial distribution of
B.D. Malhotra et al. / Analytica Chimica Acta 578 (2006) 59–74
the immobilized enzymes, the film thickness and modulation
of enzyme activity by changing the state of the polymer. The
development of a technology in this field heavily depends on
the understanding of interactions at the molecular level between
the biologically active protein, either as a simple composite or
through chemical grafting. For the proper relay of electrons from
the surface of an electrode to the enzyme active site, the concept
of ‘electric wiring’ has been reported [34,35].
Conductive polymers can be reversibly doped and undoped
using electrochemical techniques accompanied by significant
changes in conductivity. Besides this, the optical properties
of these films can be used as a signal for investigating biochemical reactions [36,37]. The electrical conductivity of conducting polymers changes over several orders of magnitude in
response to change in pH, redox potential or their environment [38]. Conducting polymers have the ability to efficiently
transfer electric charge produced by a biochemical reaction
[39]. Moreover, conducting polymers can be deposited over
defined areas of electrode. This unique property of conducting polymers along with the possibility to entrap enzymes
during electrochemical polymerization has been exploited for
the fabrication of amperometric biosensors [40–52]. Besides
this, conducting polymers exhibit exchange and size exclusion
properties due to which these are highly sensitive and specific
towards desired substrates [53–55]. Bioelectrochemistry and
electroanalysis of biologically important substances are being
intensively studied for application as chemically modified electrodes [57–60]. Conducting polymers are also known for their
ability to be compatible with biological molecules in neutral
aqueous solutions [59]. Wrighton et al. demonstrated sensor
response to oxidizing and reducing species in solution [60].
These sensors were based on changes in the electrical conductivity that accompanies oxidation and reduction of polymers
such as polypyrrole, poly-(3-methylthiophene), and polyaniline
[61–63].
The electrochemical conversion is known to occur at a conducting polymer/solution interface when the overall rate of both
the biochemical reaction and charge transport exceeds substantially the rate of mass transport of solution species to the
electrode. This phenomenon is a diffusion controlled electrocatalytic process. The combination of relative rates for the three
processes favors an efficient, reversible anodic oxidation of analyte and can be used for the amperometric detection at relatively
low electrode potential.
Redox mediators can be effective in the technical development of biosensors. The redox-active sites in redox proteins
are generally shielded by insulating peptides and sugars. The
electron transfer between enzymes and conducting polymers (or
electrodes) can be accelerated by using small electron transfer
mediators or shuttles, such as quinones and ferrocenes, which
enter the clefts of proteins and shuttle between the redox sites
of enzymes and electrode surfaces [64]. Gorton et al. have
discussed the role of direct electron transfer between hemecontaining enzymes and electrodes as the basis of third generation biosensors [65]. Schuhmann has described biosensors based
on (i) electron-transfer via conducting polymers, (ii) electrontransfer cascades via redox hydrogels, (iii) anisotropic orienta-
61
tion of redox proteins and (iv) direct electron transfer between
redox proteins and electrodes modified with self-assembled
monolayers [66].
The interest in nanomaterials for biosensing applications
has recently emerged [233–239]. Nanostructures are inorganic,
organic or composite materials synthesized in various forms with
sizes down to the nanometer. The size dependent changes in
the physical and chemical properties of materials make them
different than that of their bulk counterpart. The change in physical properties such as electrical, electronic, magnetic, optic and
thermodynamic properties makes nanomaterials important candidate for technological development of biodevices. The large
surface to volume ratio provides substantial changes in chemical properties. When size of these materials is of the order of
de Broglie wavelength, quantum confinement effect becomes
prominent which also gives rise to the unusual electrical and
optical properties. The examples of materials falling in this category (with different molecular symmetries) are fullerenes and
carbon nanotubes. Like conjugated polymers, molecular structure of fullerenes and carbon nanotubes also consist of number
of C C double and single bond. The fullerenes and carbon
nanotubes can be termed as a class of conjugated polymers.
Semiconducting nature and possibility of p-type and n-type
doping which can be made possible due to the presence of
conjugated carbon structure with unusual molecular symmetries makes them compatible with conducting polymers and
promising materials from technological point of view. Biological
molecules can be covalently attached to the carbon nanotubes.
Single wall and multi-wall carbon nanotubes were made water
soluble via esterification of nanotube bound carboxylic acid
by oligomeric polyethylene glycols which were then biofunctionalized by bovine serum albumin [68,69]. The functionalization of carbon nanotubes with bovine serum albumin (BSA)
has been found to be highly water soluble. The majorities of
carbon nanotubes are intimately associated with the BSA protein and remain bioactive. This method may be used to introduce carbon nanotubes into other biological and biomedically
important systems [68]. Ramanathan et al. have functionalized single wall carbon nanotubes with amino groups, which
could further be covalently attached to polymers or biological
systems such as DNA and carbohydrates [67]. Varfolomeyev
et al. have reviewed the recent developments in electrochemistry and electro-analytical chemistry of carbon nanotubes as
new nanomaterials [68]. Unique structural property of CNTS
makes these materials attractive to accommodate electrochemically or biologically functional elements, e.g. quantum dots,
organic compounds, biomaterials like enzymes, DNA, proteins
and antibodies paving the way for the fabrication of highly sensitive and stable biosensors. Ramanathan et al. have fabricated
bioaffinity sensors using biologically functionalized conducting
polymer nanowire [225,226]. It has been shown that sensitivity of potentially single-molecule detection can be achieved by
adjusting the nanowire’s conductivity closer to the lower end
of a semiconductor. It has been demonstrated that similar to
biological modification of PPy nanowires, monomers such as
thiophene and aniline that can be electropolymerized from an
aqueous environment can also be used. Carrara et al. have uti-
62
B.D. Malhotra et al. / Analytica Chimica Acta 578 (2006) 59–74
lized improved nanocomposite materials of poly(o-anisidine)
(POA) containing titanium dioxide nanoparticles (TiO2 ), carbon black and multi-walled carbon nanotubes (MWNT) for
biosensing applications based on electrochemical impedance
spectroscopy [232].
The biotin-labeled biomolecules can be immobilized on the
avidin-modified electrode surface through avidin–biotin complexation [104,136,216,246–248]. Nzai et al. have described
various techniques for the surface derivatization within biotin
and avidin and for the coupling of the enzymes [246]. It is
proposed that the possibility of constructing protein architecture is based on the non-covalent interaction of avidin and
biotin. Nobs et al. have covalently bound NeutrAvidinTM to the
surface of poly (dl-lactic acid) (PLA) nanoparticles with the
aim of attaching targeting compounds such as proteins to their
surface [247]. Their studies indicate that other proteins such
as antibodies could be coupled to the nanoparticles for active
targeting. Furthermore, PLA nanoparticles are interesting candidates for active targeting with biotinylated antibodies using
the biotin–avidin interaction in a two-step procedure. Gref et
al. have recently shown that biotin-poly (ethylene glycol)-poly
(-caprolactone) (B-PEG-PCL) can be helpful for studying the
interaction between cells and functionalized nanoparticles with
surface characteristics (ligand type and density, PEG layer density and thickness) [248].
The impedance spectroscopy has been used to fabricate a
number of conducting polymer and nanoparticles biosensors
based on enzymes, antibodies and mico-organisms, respectively.
Li et al. have recently reviewed the literature on impedimetric
biosensors [69]. Heeger and Heeger have revealed that conducting polymer luminescence can be quantitatively manipulated to
fabricate a variety of real-time biosensing applications including
medical diagnostics and toxicology [70].
Molecular imprinting is gradually becoming a versatile technique for the preparation of artificial receptors based on molecularly imprinted conducting polymers (MIPs) containing tailormade sites [74]. Molecular imprinting technology can be used
for the manufacture of synthetic polymers with pre-determined
molecular recognition properties. MIPS are highly stable synthetic polymers having molecular recognition properties due
to cavities in the polymer matrix that are complementary to
the analyte (ligand) both in shape and the positioning of the
functional groups and hence have been used to obtain stabilized biological response [75]. Some of these polymers exhibit
very high affinity constants and selectivity comparable to naturally occurring molecular recognition systems such as antibodies. Over-oxidized polypyrrole shows improved selectivity
due to oxygen functionality signed to the removal of positive
ions from PPy films. Sensors based on methyl pyrrole (mPPy)
have been fabricated for detection of 1-naphthalenesulphonate
and 1-serotonin [75,76]. Ho et al. have utilized poly(3,4ethylenedioxythiophene) (PEDOT) to immobilize MIP particles
of morphine onto the indium–tin-oxide (ITO) glass. The sensitivity, detection limit, linear range and the signal/noise ratio of
the modified morphine MIP/PEDOT electrode were experimentally estimated as 41.63 A cm−2 mM−1 , 0.3 mM, 0.1–2 mM
and 3, respectively [76].
Fig. 1. Schematic of a conducting polymer based biosensor.
Fig. 1 describes the schematic of a conducting polymer based
biosensor.
3. Conducting polymer based biosensors
3.1. Conducting polymer based enzyme biosensors
Enzymatic biosensors utilize the biospecificity of an enzymatic reaction, along with an electrode reaction that generates an
electric current or potential difference for quantitative analysis
[89–175]. The biomolecules such as glucose, cholesterol, urea,
triglycerides, creatinines, pesticides are important analytes due
to their adverse effects on health. Enzymatic biosensors utilize
the biochemical reactions, i.e. analyte and enzyme resulting in
a product (hydrogen/hydrogen peroxide/hydroxyl/ammonium
ion) that can be detected and quantified using a transducer
(amperometric/potentiometric/optical thermal/piezoelectric). In
general, many oxidoreductases including glucose oxidase catalyze the oxidation of substrates by electron transfer to oxygen
to form hydrogen peroxide. These oxidoreductase enzymes can
be immobilized on conducting polymer surfaces and the H2 O2
formed as a result of enzyme and the corresponding analytes may
be measured amperometrically [98–110]. However, it has not
been possible to discriminate between the direct electron transfer from the oxidation of hydrogen peroxide at polymer surface
and that at the underlying electrode [99–111]. Since conducting
polymers are insoluble in aqueous solutions, electropolymerization has been frequently used to create a matrix for immobilization of enzymes at the electrode surface, and the sensor response
was obtained by the oxidation of hydrogen peroxide [112,113].
Belanger et al. suggested that the reaction of hydrogen peroxide
and polypyrrole decreases the electrical conductivity [117]. Over
potential can be reduced by using mediators; small molecules
that shuttle between electrode and analyte to accelerate electron flows and the formal potential of the mediator should be
close to or positive to that of the analyte. Tian et al. [131] have
described an amperometric biosensor for the detection of H2 O2
based on horseradish peroxidase/polypyrrole (PPy) membrane
deposited onto the surface of ferrocenecraboxylic acid mediated derived sol–gel derived composite carbon electrode. This
biosensor had a linear range from 2 × 10−5 to 2.6 × 10−3 M),
detection limit as 0.2 mM and retained its 90% of the initial sensitivity (52.2 A mM−1 ) for about 1 week. It may be remarked
that there are some dehydrogenases that may directly transfer
B.D. Malhotra et al. / Analytica Chimica Acta 578 (2006) 59–74
electrons to conducting polymers without producing hydrogen
peroxide [223].
Sharma et al. have demonstrated that electrochemically
prepared poly(aniline-co-fluoroaniline) films for immobilization of glucose oxidase using physical adsorption technique
[94]. The shelf-life of these conducting polymer glucose
electrodes was found to be about 15 days. Singhal et al.
prepared poly(3-hexylthiophene) Langmuir–Blodgett films by
simultaneous entrapment of glucose oxidase and transferred
onto indium–tin-oxide glass [120]. This glucose electrode was
found to have detection limit of 50 mg dL−1 and sensitivity
of 0.75 nA mg−1 dL−1 and linearity from 100 to 500 mg dL−1
of glucose and stability of 75 days at 4 ◦ C [120]. Subsequently, these researchers reported a glucose biosensor based
on Langmuir–Blodgett films of polyvinylcarbazole [121]. These
glucose electrodes could be repeatedly used for 15 days for the
estimation of glucose from 100 to 500 mg dL−1 and were stable
for about 5 months at 4 ◦ C.
Uang and Chou have investigated the pH effect on the electropolymerization in the presence of glucose oxidase on the
characteristics of glucose oxidase/polypyrrole (PPy) biosensor.
This glucose sensor had linearity from 0 to 10 mM, sensitivity as
7 nA mM−1 and could be used for about 2 weeks [123]. Borale
et al. have constructed glucose biosensors based on electropolymerized polyaniline, poly (o-toluidine) and poly (anilinie-cotoluidine), respectively [124]. Sekar and Becerik have electrochemically entrapped glucose oxidase into polypyrrole matrix
alongwith p-benzoquinone in piperazine ethyl sulphate (PIPES)
buffer [125]. It is revealed that application of the artificial network analysis applied to the non-linear calibration plot can be
used to predict the sensor failure detection. Ramanavicius et al.
have prepared polypyrrole in the presence of glucose oxidase
from Pencillum vitale, glucose and oxygen revealing that optimal conditions of glucose oxidase activity (pH 6.0) are similar
to pyrrole polymerization reaction (pH 6.5) indicating possible biomedical applications of this conducting polymer [126].
These authors later reported the polypyrrole coated glucose oxidase nano-particles by self encapsulation of GOX in polypyrrole
during polymerization for application as amperometric glucose
biosensor and showed that the Km value of polypyrrole coated
GOX is 10 times larger than that for the native GOX [166]. The
shape/size and other properties of these polypyrrole coated glucose oxidase nano-particles have not yet been investigated. Pan
et al. have reported CNT based glucose biosensor using immobilization of glucose oxidase in poly (o-aminophenol) and carbon
nano-tube composite through electrochemical polymerization
onto gold electrodes [118]. They have shown that this biosensor has detection limit of 0.01 M, sensitivity of 0.114 A cm−2
and is stable for about 30 days. These authors, however, did
not investigate the thermal stability of this CNT glucose electrode. Curulli et al. have studied the influence of electrolyte
nature and its concentration on the kinetics of electropolymerization of monomers such as 1,2- and 1,4-diaminobenzene,
2,3- and 1,8-diaminonaphthalene, o-anisidine and on the resulting morphology of nanotubules [133]. Glucose oxidase (GOx),
lactate oxidase (LOD), l-aminoacid oxidase (l-AAOD), alcohol oxidase(AOD), glycerol-3-phosphate oxidase (GPO), lysine
63
oxidase (LyOx) and choline oxidase (ChOx) were immobilized
on Prussian Blue (PB) layer supported on 1,2-diaminobenzene
(1,2-DAB) nanotubes onto platinum electrodes. The nanostructured poly (1,2-DAB) based glucose biosensor showed detection
limit of 5 × 10−5 mol L−1 , a wide linear range (5 × 10−5 to
5 × 10−3 mol L−1 ) and stability for 4 weeks at room temperature. Similar results were obtained for choline, l-leucine, l-(+)lactic acid, ethanol, lysine and glycerol-3-phoaphate, respectively. Callegavi et al. have fabricated the amperometric biosensors for lactate, phenol, catechol and ethanol by incorporation
of their respective enzyme in carbon nanotubes (CNT) matrix
[122]. They have shown that the use of CNT electrode yields better results as compared to normal carbon paste electrode. The
observed stability of the H2 O2 sensor has been attributed to the
nano-tubular structure. It was shown that these nano-structured
polymers can be utilized for fabrication of enzyme biosensors.
Trojanowicz and Miernik have employed avidin–biotin interactions for the immobilization of glucose oxidase on bilayer
lipid membrane (BLM) formed on polypyrrole and polyo-phenylylenediamine) electrodeposited onto platinum wire,
respectively [136]. A lipid solution comprising of 1.6 mg of
biotin DHPE (N-(biotinoyl)-1,2-dihexadecaneoyal-sn-glycero3-phosphor ethanolo-amine, trethylammonium salt) and 0.4 mg
of cholesterol dissolved in 1 cm3 of a mixture of n-decane and
n-butanol (8:1, v/v) was used for lipid formation. The solid
supported BLMs were formed on a bare or polymer-modified
platinum surface (0.6 mm in diameter in each case) by immersing the tip of the electrode in the lipid solution for 5 min and
then in avidin–enzyme conjugate for 5 min. These BLM based
glucose biosensors resulted in stable and sensitive response with
significant reduction from electro-active species such as ascorbic acid, cholesterol and uric acid, respectively.
Mylar et al. have reported improved signal linearity of
enzyme (glucose oxidase) ultra micro-electrodes fabricated
via sonication and deposition of polysiloxane coating onto
the working glucose oxidase/polyaniline electrode coated with
insulating diaminobenzene [137]. He et al. [139] have developed a sensitive hydrogen peroxide probe based on amplified fluorescence quenching ability of poly(9,9-bis(6′ -N,N,Ntrimethylammonium-hexyl) fluorene phenylene (PFP-NMe3 + ,
a cation conjugated polymer) and peroxyfluor-1 borate protective groups. The hydrogen peroxide probe has a detection range
of 15–600 nM and can be used for glucose detection.
Conducting polymer based self-regulating insulin delivery
system has recently been reported [140,141]. It has been brought
out that conducting polypyrroles provide good alternatives due
to biocompatibility and in vivo stability. Experiments are however being conducted to choose a suitable conducting polymers
or composites for in vivo applications.
Arslan et al. have fabricated a polysiloxane/polypyrrole/
tyrosinase electrode by entrapment of tyrosinase in conducting matrix by electrochemical copolymerization for determining
phenolic content of green and black tea [128]. Compared to
sodium dodecyl sulfate (SDS) doped electrodes, the p-toluene
sulfonic acid (PTSA) doped enzyme electrodes show good pH
resistance and temperature stability and the effect of supporting electrolyte (PTSA). However, the PTSA doped conducting
64
B.D. Malhotra et al. / Analytica Chimica Acta 578 (2006) 59–74
polymers could not be used to detect phenolic components in tea
because of the apparent high Km value. Vedrine et al. have fabricated an amperomteric tyrosinase biosensor based on conducting
poly(3,4-ethylenedioxythiophene) (PEDOT) for estimation of
herbicides and phenolic compounds [140]. The detection limits
for monophenol and di-phenol ranged from 5 to 500 nM and the
detection limit for diuron and attrazine were found to be 0.5 and
1 mg L−1 , respectively. Böyükbayram et al. have made a comparative study of immobilization methods of tyrosinase (Tyr) on
electrolpoymerized conducting and non-conducting polymers
for application to the detection of dichlorvos organaophosphorous insecticide [141]. A detection limit of 0.06 m was obtained
for dichlorvos using mediator 1,2-naphthoquinone-4-sulphonate
(NQS) and Tyr. Jiang et al. have demonstrated that incorporation
of polyvinylalcohol (PVA) onto a polymer film caused higher
sensitivity than that of pure PPy sensor [130]. Rahman et al. have
fabricated the amperometric biosensor for choline by immobilizing choline oxidase (ChOx) with and without horseradish
peroxidase (HRP) onto carboxylated thiophene modified electrodes [135]. They have shown that the electrode with HRP has
better performance and these biosensors have the response time
of 5 s and selectivity to interfering species.
Kan et al. have reported stronger affinity between uricase
and polyaniline prepared by template process resulting in the
increased stability of this polyaniline–uricase biosensor [134].
However, the activation energy of this uricase electrode has been
found to be larger by 29.9 kJ mol−1 than the uricase–polyaniline
biosensor fabricated using two-step processes.
Haccoun et al. have fabricated a reagentless lactate biosensor using electro-copolymerized copolymer film of poly(5hydroxy-1,4-naphthoquinone-co-5-hydroxy-3-acetic acid-1,4naphthoquinone) [142]. These studies show that the presence
of interferents like acetaminophen, glycine and ascorbic acid
does not influence the response of this mediated (quinone
group) copolymer electrode. However, this biosensor can be
used for l-lactate estimation upto 1 mM of l-lactate. Bartlett et
al. have found that poly(aniline)–poly(anion) composites films
can be utilized for electrochemical oxidation of NADH at around
50 mV versus SCE and pH 7 [143]. These results have implications towards the technical development of microelectrodes,
biofuel cells and amperometric biosensors. Chaubey et al. coimmobilized lactate oxidase (LOD) and lactate dehydrogenase
(LDH) on electrochemically prepared polyaniline (PANI) films
by physical adsorption technique [145]. The LOD/LDH/PANI
electrodes were shown to have linearity from 0.1 to 1 mM
of lactate and detection limit of 5 × 10−5 M and stability of
3 weeks at 4–10 ◦ C. Subsequently, these researchers electrochemically entrapped polyaniline onto sol–gel derived tetraethylorthosilicate (TEOS) films deposited onto indium–tin-oxide
coated glass electrodes for immobilization of lactate dehydrogenase (LDH) [146]. The application of external layer of
polyvinyl chloride (PVC) onto sol–gel/PANI/LDH electrodes
resulted in the extended linearity from 1 to 10 mM of l-lactate.
These electrodes had a shelf-life of 8 weeks at 0–4 ◦ C and the
common interferents such as glucose (100 mg dL−1 ), uric acid
(35 mg dL−1 ), glutamic acid (25 mg dL−1 ) and ascorbic acid
(25 mg dL−1 ) affected the observed amperomteric response.
Singh et al. have reported conducting polypyrrole/cholesterol
oxidase electrodes based on conducting polypyrrole films
for estimation of cholesterol ester concentrations of 1–8 mM
[147,148]. The applicability of these electrodes to serum samples has not yet been demonstrated. These researchers have
co-immobilized cholesterol oxidase, cholesterol esterase and
peroxidase onto electrochemically prepared polyaniline films.
This polyanline based cholesterol biosensor has a response time
of 240 s, an apparent Km value as 75 mg dL−1 , sensitivity value
as 0.042 A mg dL−1 , detection limit as 25 mg dL−1 , shelf-life
of 6 weeks and can be used to estimate cholesterol concentration
up to 500 mg dL−1 .
Asberg and Inganas have cross-linked horseradish peroxidase in highly conducting poly (3,4-ethylene dioxy thiophene)
(PEDOT)/(polystyrene sulphonate) (PSS) dispersion using
poly-4-vinylpyridine for estimation of hydrogen peroxide
in the concentration range of 0–30 M [149]. Grennam et
al. have shown that horseradish peroxidase immobilized
screen-printed electrodes based on chemically polymerized polyaniline/polyvinylsulphonate films can be used
for the mass production of biosensors [150]. Morrin et al.
have electrochemically applied nanoparticulate polyaniline
(PANI) doped with dodecylbenzenesulphonic acid (DBSA)
to glassy carbon electrode surface for physical adsorption
of horseradish peroxidase [151]. Compared to electrochemically prepared polyaniline/polyvinylsulphonate films, the
nanoPANI/dodecylbenzenesulphonic acid (DBSA) showed
faster response time (0.62 s) and improved linearity for estimation of H2 O2 and better signal to noise ratio (61 ± 3). These
results have been attributed to the highly ordered structure of
nanoPANI/DBSA system. Jia et al. have fabricated horseradish
peroxidase (HRP) biosensor by self-assembling gold nanoparticles to a thiol-containing sol–gel network of 3-mercaptaopropyltrimethoxysilane (MPS) [152]. The response time, detection
limit, linear range and stability of the HRP biosensor for H2 O2
were found to be 2.5 s, 2.0 mol L−1 , 5.0–10.0 mol L−1
and 120 days, respectively. Ngmana et al. have immobilized
horseradish peroxidase poly (2-methsulphonated polyaniline5-sulphonic acid)/l-lysine composite [153]. The correlation
coefficient sensitivity, detection limit and the linear range of the
amperometric HRP biosensor were determined to be 0.9966,
24.91 A cm−2 , 0.01 mM, 0.01 mM of H2 O2 , respectively.
Mathebe et al. have electrostatically immobilized horseradish
peroxidase on the surface of polyaniline film electrochemically
deposited onto platinum disc electrode [154]. The correlation
coefficient and linear range of the polyaniline/peroxidase based
biosensor were found to be 0.995 and 2.5 × 10−4 to 5 × 10−3 M,
respectively. Zhou et al. have observed enhanced peroxidase
activity in hemoglobin in sulfonated polyaniline on glassy carbon electrodes [155]. This result has been attributed to the facile
interfacial transfer of hemoglobin mediated by sulphonated
polyaniline. Lindgren et al. have immobilized horseradish
peroxidase (HRP), sweet potato peroxidase (SPP) and peanut
peroxidase (PNP) tobacco peroxidase (TOP) on graphite
electrodes [156]. The peroxidase electrodes were used for
estimation of H2 O2 . SPP was found to have the lowest detection
limit.
B.D. Malhotra et al. / Analytica Chimica Acta 578 (2006) 59–74
Dai et al. electrochemically immobilized cytochrome C on a
NaY Zeolite modified electrode [157]. This electrode displayed
excellent response to the reduction of H2 O2 without the use of
electron mediator and hence could be used for H2 O2 detection.
Brahim et al. have developed a p (2-hydroxyethyl methacrylate/polypyrrole system containing glucose oxidase, cholesterol
oxidase and galactose oxidase [158]. The observed stability
of 9 months for this system indicates that this biosensor can
be subcutaneously used to monitor glucose, cholesterol and
galactose. Pijanowska et al. have made a comparison of urea
estimation in blood plasma and in dialysate for the physiological and the pathological range by enzyme-field-effecttransistor (EnFET) based detection of pH and pNH4 [159]. It
is concluded that pNH4 based biosensors require pre-dilution
whereas for pH biosensors both for blood plasma and dialysate
do not require dilution. Zhang et al. have discovered that
stable mixed amphiphile(octadecylamine, ODA/behenic acid,
BA)/urease Langmuir–Blodgett films can be formed onto the
hydrolyzed surface of a pH-ISFET [160]. This urea biosensor has response time of 15 s, detection limit of 0.2 mM and
linearity of 0–20 mM. Gambhir et al. have covalently immobilized urease on polypyrrole microspheres linked to conducting polypyrrole–polyvinylsulphonate (PPy–PVS) films [161].
These conducting polymer electrodes had a response time of
about 40 s, were thermally stable for about 40 days at 25–50 ◦ C
and could be used for estimation of urea from 5 × 10−3 to
6 × 10−2 mol L−1 using potentiometric technique for application to urea biosensor.
Singhal et al. immobilized urease in mixed monolayers of
polyvinylcarbazole (PNVK) and stearic acid (SA) [162]. Two
values (9 and 30 mM) of Michaelis–Menton constant (Km )
were obtained at lower and higher concentrations of urea.
These PNVK/SA/urease electrodes, stable for about 5 weeks
at 4 ◦ C, were found to have detection limit and sensitivity as
be 5 mM and 10 mV mM−1 , respectively. Rajesh et al. have
fabricated an amperometric urea biosensor using covalent
immobilization of urease onto poly (N-3-aminopropyl pyrroleco-pyrrole (PAPCP) copolymer film [163,242,244]. The copolymer electrode (PAPCP) was immersed in a phosphate buffer
solution (0.1 M, pH 7.0) containing 0.015 M ethyl-dimethylaminopropylcarbodiimide (EDC) and 0.03 M and N-hydroxysuccinimide (NHS) for 1.5 h and was immediately placed in an
enzyme solution of 10 mg mL−1 in the same buffer solution for
another 1.5 h. These urease/PAPCP/ITO electrodes show stability for more than 2 months when stored at 4–6 ◦ C, linearity in the
range of 0.16–5.02 mM, response time of 40 s, current sensitivity
as 0.47 A mM−1 cm−2 . The efforts are currently being made
to improve the stability and ascertain the effect of interferents on
these urea bioelectrodes. Rebriiev et al. have fabricated a urea
biosensor based on ion-sensitive-field-effect-transistor (ISFET)
and photopolymeric membrane obtained by irradiation of a
liquid mixture comprising of vinylpyrrolidone, oligourethane
methacryalate and oligocarbonate methacrylate [164]. This urea
biosensor exhibited high sensitivity, response time as 20 min,
linearity as 0.05–20 mM and stability for about 40 days.
Sharma et al. have developed a monoenzyme amperomteric
biosensor for the estimation of galactose in milk by immobi-
65
lizing galactose oxidase with poly (3-hexyllthiophene)/stearic
acid onto indium–tin-oxide (ITO) coated glass plates using
Langmuir–Blodgett film deposition technique [93–95]. The
biosensor shows response time of 60 s, detection limit as
1 g dL−1 , linearity from 1 to 4 g dL−1 of galactose and shelflife as 90 days. These investigators later co-immbilized lactase
and galactose oxidase onto Langmuir–Blodgett films of poly(3hexylthiophene)(P3HT)/stearic acid (SA) for estimation of lactose in milk and its products to prevent lactose intolerance [96].
These enzyme electrodes show a linearity of 1–6 g dL−1 of lactose, has a shelf-life of 120 days and could be used about ten
times. These galactose oxidase (GaO) immobilized P3HT-SA
electrodes were used for estimation of galactose in human blood
[95]. The amperometric galactose biosensor based on P3HT had
detection limit of 0.5 g L−1 of galactose and were stable for about
30 days. There was no interference with uric acid (0.1 mM),
CaCl2 (0.1 mM) and ascorbic acid (0.2 mM).
Pinto and Schanze have demonstrated the application of
conjugated polyelectrolytes (CPEs) based on pendent ionic
sulphonate and carboxylate groups for estimation of protease
activity. The sensor has been used for estimation of low enzymeconcentrations and enzyme-catalyzed kinetics [168]. This technique can be used to develop assays for kinases, phosphates,
lipases and esterases, etc. Ivanov et al. have used glassy carbon
electrodes modified with polyaniline for the immobilization of
cholinesterase using cross-linking technique [169]. Compared
to other detectors, these polyaniline modified potentiometric
biosensors utilized for the estimation of pesticides (trichlorfon, coumaphos, metiocarb and aldicarb) exhibit higher storage
stability, increased sensitivities and lower detection limits. The
degree of inhibition referred to as the total inhibitory effect of
all the toxic compounds could be determined in about 15 min.
Cosnier et al. have demonstrated that the immobilized thionine in the poly (dicarbazole-N-hydroxysuccinimide) results in
the improved sensitivity (0.1457 A M−1 cm−2 ) and the maximum current (15.43 A cm−2 ) of catechol of the polyphenol
oxidase based biosensor [170,171]. These authors later fabricated an enzyme electrode by coupling of avidin and biotinylated
polyphenol oxidase with biotinylated polycarbazole film resulting in improved biosensor stereoselectivity [170,171].
Rahman et al. have covalently immobilized pyruvate oxidase
onto nanoparticles comprising of poly 5,2′ :5,2′′ -terthiophene3′ -craboxylic acid, poly-TTCA (nano-CP) on a glassy electrode
for amperometric detection of phosphate ions [172]. The glassy
carbon electrode (GCE) coated with nano-CP was immersed in
a 20 mM HEPES buffer solution (pH 7.0) containing 10.0 mM
ethyl-dimethyl-aminopropylcarbodiimide (EDC) for 6 h to activate the carboxylic groups of the nano-CP layers. The nanoCP/GCE was washed with HEPES buffer solution and subsequently incubated in 100 units of pyruvate oxidase in a 20 mM
HEPES buffer (pH 7.0) solution for 12 h at 4 ◦ C. The immobilization of pyruvate oxidase occurred through the formation of
covalent bond between carboxylic acid groups of the conducting polymer and amine groups of the enzyme. The detection
limit, linearity, the response time and the shelf-life of this nanoCP biosensor was observed to be 0.3 M, 1.0–100 M, 6 s and 1
month, respectively.
66
B.D. Malhotra et al. / Analytica Chimica Acta 578 (2006) 59–74
Langer et al. have immobilized choline oxidase in nanostructured polyaniline layers of controlled nano and micro porosity for estimation of choline in food [173]. The choline oxidase
based polyaniline biosensor is stable for about 30 days, has
sensitivity of 5–10 mV mM−1 in the potentiometric mode and
5 A mM−1 in the amperometric mode. Qu et al. have reported
an amperomteric biosensor for detection of choline based on
polyaniline multilayer film and layer-by-layer assembled functionalized carbon nanotube [174]. The choline biosensor showed
response time as 3 s, detection limit as 0.3 M and linear range
as 1 × 10−6 to 2 × 10−3 M.
Biloivan et al. demonstrated that protein selective microelectrodes can be fabricated by immobilization of different proteinases such as trypsin, pronase E and carboxypeptidase B on
surface of platinum electrodes modified by conducting PPy[3,3Co(1,2-C2 B9 H11 )]2 . This biosensor had a detection limit of
1–2 g mL−1 in phosphate buffer [175].
3.2. Conducting polymer based microbial biosensors
Microorganisms as sensing elements provide advantages
like over-coming the process of purification for enzymes,
multi-enzyme functions and cofactor/coenzyme addition,
etc. [176–180]. The recent progress in molecular biology/recombinant DNA technologies has opened enormous possibilities of tailoring microorganisms to improve the activity of
an existing enzyme or express foreign enzyme/protein in a host
cell such that it can be utilized to enhance specific activity. There
have been various strategies to modify the microbes for application to microbial biosensors such as lux gene expressions for
fluorescent protein, gfp gene for green fluorescence protein. The
enhanced expressions of gene of interest have been reported to
improve the sensitivity for target analytes. Besides this, various electrochemical detection methods have been used to integrate microbes with various immobilization matrices. Kwok
et al. immobilized Bacillus subtilis in a sol–gel derived biocompatible material of silica and poly (vinyl alcohol)-graftedpoly (4-vinylpyridine) (PVA-g-PVP) copolymer resulting in fast
diffusion of organic compounds [177]. This prototype optical
biosensor can be used for the estimation of biochemical oxygen
demand (BOD) in waste water samples. Lei et al. have reported
a whole cell biosensor using PNP-degrader P. putida JS444 for
estimation of organo-phosphorus nerve agent with p-nitrophenyl
constituent [178]. This biosensor was found to be stable for
only 5 days. Applications such as BOD sensing, toxicity sensing, microbial fuel cell and various other important biological
parameters such as alcohol, sugar, phenol, methanol, heavy metals, cyanide, etc. appear to be the most promising. Efforts are
being made to commercialize some of microbial sensors through
optimization of the microbial species, immobilization medium,
oxygen sensing film and suitable engineering design. Minet
et al. have reported that a mediated (toluidine blue) antibody
(Salmonella and E. Coli)-immobilized polypyrrole electrode can
be used for detection of Listeria monocytogenes using amperometric method [180]. This biosensor could reproducibly and
selectively detect Listeria after incubation in the micro-organism
at levels of 105 cells mL−1 in 30 min.
3.3. Conducting polymer based immunosensors
Immunosensors can be used to detect the presence of an
analyte in minutes using small volumes of sample [181–190].
An effective combination of immunochemistry coupled with
electro-chemistry in an analytical device could provide the basis
of direct electrical detection for a wide range of analytes with
specificity and great sensitivity. A number of immunosensors
based on conducting polymers have been reported. A unique
approach for conducting polymers reversible immunosensors
using pulsed electro-chemical detection has been developed.
The strategy involves the detection of organochlorine pesticides including PCBs, atrazines and chlorinated phenols. Porter
has investigated electro-polymerized conducting polymers as
antibody receptors in [181]. They have shown that antibodies against conducting polymers (carbazole as a hapten) may
react to modulate the polymer electrochemistry. The reaction
of the anti-serum was found to influence the polymer electrochemistry by amperometric response and therefore can be utilized as a sensor for amino acid. Grant et al. have reported a
label free and reagentless immunosensor based on direct incorporation of antibodies into conducting polymer films using ac
impedometric electrochemical interrogation [182]. It has been
demonstrated that faradic real component of the impedometric response acts as a dominant component of AC impedometric response of anti BSA loaded conducting polypyrrole
(PPy) film on its exposure to different concentration of BSA.
The nature of observed Faradaic response current, perhaps
arising due to antibody–antigen interaction in this system is
still a matter of speculation. Ramanaviciene and Ramanavicius have discussed the use of conducting polymers thin films
for application as electrochemical affinity sensors with the
emphasis on design and applications of novel immunosensors [184]. They have briefly discussed the biological active
component for the creation of polypyrrole based immunosensors. Gooding et al. have fabricated the glassy carbon electrode modified with anti-rabbit IgG antibody entrapped in an
electrodeposited polypyrrole membrane for label free amperometric detection of rabbit IgG antigen in flow injection system [183]. The observed reversible antibody–antigen binding was assigned to the short anodic pulse (200 ms). The
changes in the concentration of the immobilized antibody within
a polymer matrix significantly affected the stability of the
immuno-electrode.
Zhang et al. have fabricated a low-cost label-free amperometric immunosensor based on anti-rubella serum immobilized
onto nano-Au/poly-o-phenyldiamine doped with Prussian blue
for the detection of rubella vaccine [185]. This immunosensor had a signal to noise ratio of 3 and exhibited response to
rubella vaccine in the range of 8.1 × 10−8 to 8.0 × 10−6 cell
culture infective dose (CCDI) mL−1 . Tahir et al. have described
the characteristics of polyaniline compounds in different protonic acid for application to diarrhea virus detection based on
polyclonal and monoclonal BVDV antibodies [186]. They have
shown that PANI with perchloric acid show highest conductivity
in pH 6.6 and the sensitivity of the biosensor ranged from 103
to 104 cell culture infective dose (CCDI) mL−1 . They however
B.D. Malhotra et al. / Analytica Chimica Acta 578 (2006) 59–74
did not investigate the effect of polymerization conditions on the
characteristics of the biosensor.
Darain et al. have developed a mediator-less immunosensor
for the detection of rabbit IgG (RIgG) by covalently bonding horseradish peroxidase and streptavidin onto conducting
polymer(5′ ,2:5′ 2′′ -terthiophene-3′ -carboxylic acid) coated on a
screen-printed electrode [188]. The catalytic current measured
amperometrically at −0.35 V versus Ag/AgCl showed detection
limit as 0.33 g mL−1 and a linear range of RIgG concentration
from 0.5 to 2 g mL−1 .
Lillie et al. fabricated simple immunosensor formats by polymerizing pyrrole loaded with avidin or antibody to luteinising
harmone (LH) on a gold inter-digited electrode and demonstrated that impedance spectroscopy can be used to detect LH
between 1 and 800 IU L−1 [189]. Farace et al. have developed a
reagentless immunosensor for detection of luteinising harmone
based on antibody entrapped in a conducting polypyrrole matrix
using impedance spectroscopy [190]. Interestingly, the interaction of DNA with conducting polypyrrole could be used to
differentiate single- and double-stranded DNA.
3.4. Conducting polymer based nucleic acid biosensors
DNA probes and biosensors have widely attracted much
attention for the diagnosis of various disorders [191–219].
DNA biosensors have enormous applications in clinical diagnostics for inherited diseases, rapid detection of pathogenic
infections, and screening of c-DNA colonies required in molecular biology. Conventional methods for the analysis of specific
gene sequences are based on either direct sequencing or DNA
hybridization. Lucarelli et al. have recently reviewed the literature in relation to the application of gold and carbon electrodes
as electrochemical transducers for DNA hybridization sensors
[202]. Because of its simplicity, most traditional techniques in
molecular biology are based on hybridization. Marrazza et al.
have reported an electrochemical DNA biosensor based on covalent immobilization of double stranded calf thymus DNA on the
surface of a disposable screen-printed graphite electrode using
avidin–biotin coupling [187]. Differential pulse voltammetry
(DPV) and potentiometric stripping analysis at constant current were used to detect a specific sequence of bases and the
presence of a gene or an infectious agent of the specific oligonucleotide target in solution to the concentration of 2 g mL−1
necessitating amplification of hybridization signal for real samples. In this context, conducting polymers are reported to be good
candidates for DNA immobilization [161]. Gambhir et al. have
immobilized calf thymus DNA on electrochemically prepared
conducting polypyrrole/polyvinyl sulphonate films [193]. Ban
et al. have developed a method for investigating DNA–protein
interaction using electrochemical techniques [210].
Yu et al. have developed an amperometric enzyme-linked
immunoassays based on single-wall carbon nanotubes (SWNT
forests) on pyrolytic graphite surfaces for estimation of human
serum albumin (HSA) using horseradish peroxidase (HRP)
labels [194]. Using mediator (hydroquinone) alongwith SWNT
provided the detection limit of HSA to 1 pmol mL−1 (1 nM).
Some of the issues currently being addressed relate to the pat-
67
terning of SWNT forests into arrays for multiple protein assays
and the strategy to avoid soluble mediation.
Cha et al. have reported the application of conducting poly
(thiophen-3-yl-acetic acid 1,3-dioxo-1,3-dihydro-isoindol-2-ylester) (PTAE) to DNA hybridization electrochemical sensor
[199]. This sensor has a sensitivity of 0.62 A nmol−1 and detection limit as 1 nmole of target oligonucleotides (ODN). The limited selectivity of this sensor has been assigned to the increased
steric hindrances resulting due to partial destruction of double
helix structure. The physical and chemical conditions affecting
the response have not yet been studied. Youssufi and Makrouf
have prepared conducting polypyrrole substituted with ferrocenyl groups as DNA electrochemical sensor [203]. This sensor
has a detection limit of less than 1 pmol of DNA target. Both the
thermal stability and the effect of interferents on the response of
the DNA biosensor have not yet been reported. Gu and Hasebe
have incorporated horseradish peroxidase (HRP) and methylene
blue (MB) in polyion complex membrane comprising of DNA
(ds-DNA) and poly (allylamine) (PAA) on the surface of gold
(Au) disk electrode to fabricate second generation H2 O2 biosensor [204]. This HRP/DNA-MB/PAA/Au bioelecrode had high
sensitivity (14.6 A−1 mM−1 ) and detection limit as 0.3 M.
The attempts are being made to apply this electrode to other
oxidases that produce H2 O2 . Shi et al. have coated poly (anilineaniline boronic acid) on DNA templates to fabricate redox-active
polymer wires that electrically contact glucose oxidase reconstituted on the polymer wires with the electrode [205]. It is
envisaged that glucose oxidase may be similarly contacted with
negatively charged polyelectrolyte templates like polystyrene
sulphonate. Arora et al. have physically immobilized double
stranded calf thymus deoxyribonucleic acid (DNA) onto electrochemically prepared polypyrrole-polyvinylsulphonate (PPyPVS) films [90]. The amperometric response studies of the
DNA/PPY-PVS electrodes carried out at 25 ◦ C as a function
of 2-aminoanthracene (2 AA) concentration (0.01–20 ppm) and
o-chlorophenol (OCP, 0.1–30 ppm) reveal that 10 ppm is sufficient to reduce the observed guanine oxidation current and
that 25 ppm of OCP reduced the oxidation current to zero.
The effect of interferents like calcium, magnesium carbaryl and
bisphenols on the response of the DNA/PPY-PVS electrodes has
not yet been investigated. Wu et al. have reported a biosensor
based on polyaniline intercalated graphite oxide nanocomposite (PAI/GO) for monitoring DNA hybridization [206]. These
workers observed that these electrodes without ss-DNA exhibited linear behaviour between 34 and 241 g mL−1 of ss-DNA.
However, when the PAI/GO electrode was immobilized with
ss-DNA, the hybridization current was found to be linear from
275 to 241 g mL−1 of complementary ss-DNA. These results
showed that paraffin–graphite–PAI/GO electrode can be used for
the estimation of ss-DNA. The effect of temperature and the various interferents on the response characteristics of the modified
PAI/GO electrode has not yet been studied.
Kerman et al. have fabricated a magnetic bead-based
hybridization assay for application as peptide nucleic acid
probes. This method can be used to detect specific sequences
in PCR amplified DNA samples [211]. Fan et al. have fabricated
an electrochemical DNA sensor based on molecular beacon
68
Table 1
Characteristics of various conducting polymer based biosensors
S. no.
Bioensing molecule
Analyte
Immobilization method
Method of detection
Response time
Detection limit
Linearity
Interference
Stability
Reference
Polypyrrole
Virostate anti-AZT
AZT
10 pg mL−1
–
–
–
[58]
Poly(oaminophenol/carbon
nanotubes)
Ferocenyl carbon
nanotubes/polypyrrole
Polysiloxane/polypyrrole
Glucose oxidase
Glucose
Electrochemicalimpedance
Amperometric
–
2
Electrochemical
entrapment
Electro-Polymerization
3s
0.01 mm
Upto 5 mm
30 days
[118]
Glucose oxidase
Glucose
Electro-polymerization
Amperomteric
–
–
–
No interference from
ascorbic acid, uric acid
and acetaminophen
–
–
[122]
Tyrosinase
Catechol
Electrochemical
entrapment
Spectro-photometric
–
–
–
–
35 days
[128]
5
Polypyrrole-poly(vinyl
alcohol)
Methanol
Resistance change
5 min approx
–
50–1059 ppm
–
2 weeks
[130]
6
Polyaniline
Polypyrrolepoly(vinyl alcohol)
film
Uricase
60 days
[134]
7
No interference from
ascorbic acid
–
–
[149]
3
4
8
9
10
11
12
13
Uric acid
Template
Amperometric
10 s
–
0.0036–1 mmol dm3
HRP
Poly(3,4ethelenedioxythiophene)poly(styrenesulfonate) + poly4-vinylpyridine
Lactate oxidase
Poly(5-hydroxy-1,4naphthoquinone-co-5hydroxy-3-acitic
acid-1,4-naphthoquinone)
Polyaniline/polyvinylsulphonate
Alcohol
dehydrogenase
Octadecylamine
Urease
H2 O2
Entrapment
–
–
0–30 M
–
Lactate
Covalent
Amperometric
–
50 M
50–1500 M
No interference from
ascorbic acid, glycine
and acetaminophen
9 weeks
[142]
-nicotinamide
adenine dinucleotide
Urea
Solution phase
Amperometric
–
–
–
–
–
[143]
Potentiometric
15 s
0.2 mm
0–20 mm
–
–
[160]
Polyanilinedodecylbenzenesulphonic
Poly(phenyleneethynyleneSO3 /CO2
Glassy carbon-polyaniline
HRP
H2 O2
Adsorption in the
subhase of Langmuir
Layer
Electrostatic adsorption
Colorimetric
–
–
–
–
–
[151]
Protease
P-Nitranilide and
rhodamine-Arg-2
Organophosphorus
and carbamic
pesticides
BOD in waste water
–
3 min
–
–
–
–
[168]
Cross-linking
Fluorosscence
spectroscopy
–
–
Coumaphos
–
–
–
[169]
–
Optical
20 min
–
Heavy metals
(transition metals)
45 days
[177]
0.2–50 M−1
5 days
[178]
0–75 ppm
No interference from
phenolic compounds
and pesticides
–
–
[182]
–
[183]
60 days
[185]
–
[187]
Cholinesterase
5 × 10−9
Oxygen sensing film
15
Silica-tris(4,7-diphenyl1,10-phenanthroline)
ruthenium (II)
Oxygen sensing film
Pseudomonas putida
JS444
Organophosphates
–
O2 consumption
Less than 5 min
25 mg L−1 for B.
subtilis and 60 mg L−1
for activated sludge
Lower than 60 ppb
16
Polypyrrole
Anti-BSA
BSA
ac Impedance
–
–
17
Glassy carbon/polypyrrole
Anti-rabbit IgG
Rabbit IgG
Electrochemical
entrapment
Entrapment
Amperometric
–
–
8.1 × 10−8
14
18
Nano Au-ophenylenediamine/prussian
blue
Anti-rubella
Rubella
Physical
Amperometric
6 min
4.010–8 lgccid 50 mL−1
19
Graphite screen printed
electrode
DNA
DNA
Biotin–avidin/adsorption
–
1 g mL−1
–
20
Polypyrrole-polyvinyl
sulphonate
Single wall carbon
Nanotube
PTAE
Ferrocenyl popypyrrole
DNA
OCP
Physical
Differential pulse
voltammetry and
chronopotetiometry
Amperometric
No interference from
human serum
No significant
interference rom
Mumps, varicella,
human serum albumin,
etc.
–
30 s
–
0.1–25 ppm
–
4 months
[90]
Anti-human serum
albumin (HSA)
DNA
Amino-DNA
Human serum albumin
(HSA)
DNA
DNA
Covalent
Amperometric
1 nm
–
–
–
[193]
Substitution
Covalent
Cyclic voltammetry
Cyclic voltammetry
1 nmole
10−14 mol
–
–
–
–
–
–
[199]
[203]
21
22
23
to
8 × 10−6 lgccid 50 mL−1
B.D. Malhotra et al. / Analytica Chimica Acta 578 (2006) 59–74
Matrix
1
24
25
PNA/polymer/nuclease
enzyme
DNA–poly(allylamine)
Ferrocene tagged
PNA
HRP/methylene blue
26
Polyaniline
27
Polyaniline–graphite oxide
nano composite in carbon
paste electrode
Poly(9,9-bis(6′ -N,N,Ntrimethylammoniumhexyl)fluorene
phenyle)
Polyttca/glassycarbon
Polyttca/carbon
Cholesterol oxidase,
cholesterol esterase
and peroxidase
DNA
28
29
30
Fluoroscence
2–3 s for each
addition
<10 s
–
–
–
–
[205]
0.3 M
Upto 0.1 mm
5 days
[206]
6 weeks
[249]
H 2 O2
Solution cast
Square wave
voltammetry
Cholesterololeate and
cholesterol
Covalent
Spectroscopic and
amperometric
240 s
25 mg dL−1
25–500 mg dL−1
No interference from
ascorbic acid,
l-cysteine and uric acid
–
DNA
Physical adsorption
Conductometry
–
–
–
–
–
[208]
Peptide nucleic acid
DNA
Solution phase
Fluorescence detection
–
10 pm
–
–
–
[209]
Pyruvate oxidase
HRP/streptavidin/Anti
rabit igg
Polyphenol oxidase
Phosphate ions
Rabit igg
Covalent immobilization
Avidin–biotin coupling
Amperometric
Amperometric
6s
35 min
0.3 M
0.33 g m L−1
1.0–100 M
0.5–2 g mL−1
No interference
–
1 month
–
[172]
[188]
Thionine
Grafting
Amperometric
–
–
–
–
–
[170]
HRP
H 2 O2
Drop coating
Amperometric
–
0.01 mm
0.01–0.1 mm
No interference of
ascorbic acid
6 months
[153]
Glucose oxidase
Glucose
Galvanostatic
–
–
0–10 mm
14 days
[123]
–
–
0–35 mm
1 month
[173]
200 s
0.3 mm
0.1–2 mm
–
[76]
1 month
[174]
33
Poly(dicarbazole-Nhydroxysuccinim
ide)
Poly(2-methoxyaniline-5sulfonic
acid)
Polypyrrole
34
Polyaniline
Choline oxidase
Choline
Galvanostatic
electropolymerization
Diffuion
35
PEDOT
MIP
Morphine
Precipitation
Amperometric and
potentiometric
Amperometric
36
Polyanilne/multi-wall
nanotube (MWNT)
Polyanilie
Poly 3,4-ethylenedioxythi
ophene
Polypyrrole
Polypyrrole
Thiophene capped
polytetrahydrofuran
(TPTHF)-co-pyrrole (Py)
1,2-Diaminobenzene
Polypyrrole
Choline oxidase
Choline
Cross linking
Amperometric
3s
0.3 M
1 × 10−6 to 2 × 10−3 M
No interference of
ascorbic acid
No interference of
codeine
No interfere nce
HRP
Tyrosinase
H2 O2
Phenolic compounds
Electrostatic
Entrapment
Amperometric
Amperometric
–
20–40 s
2.5 × 10−4 M
–
2.5 × 10−4 to 5 × 10−3 M
5–500 nm
–
–
–
12 days
[154]
[140]
HRP
Anti-listeria
Polyphenol oxidase
H2 O2
Listeria
Phenolic compounds
Entrapment
Entrapment/covalent
Entrapment
Amperometric
Cyclic voltammetry
Spectrophotometric
method
<20 s
30 min
–
5 × 10−5
105 cell mL−1
–
9.0 × 10−7 –2.0 × 10−4 M
–
–
–
–
–
2 weeks
–
20 days
[131]
[180]
[141]
Glucose oxidase
Trypsin
Covalent
Cross-linking
Amperometrically
Potentiometric
15 s
–
0.05 mmol L−1
0.05 mmol
0.05–5 mmol L−1
0.05–0.5 mmol
No interference
–
90 days
–
[119]
[175]
DNA
Entrapment
ac Impedance
30 min
–
1.0 × 10−5 –3.0 × 10−8 mol L−1–
–
[219]
45
46
Polypyrrole–MWNTCOOH
Polypyrrole
Polypyrrole
Glucose
N␣-benzoyl-l-arginine
ethylester hydrochlori
de
DNA
DNA
DNA
DNA
DNA
Entrapment
Covalent
–
–
–
–
–
–
–
–
–
–
[217]
[192]
47
Polyterthiophene
DNA
DNA
Covalent
–
–
–
–
–
[218]
48
Polyaniline
DNA
DNA
Covalent
Microgravimetric
Photocurrent
spectroscopy
Impedance and
admittance
Differential pulse
voltammetry
–
1.0 × 10−12 mol L−1
–
–
–
[212]
32
37
38
39
40
41
42
43
44
B.D. Malhotra et al. / Analytica Chimica Acta 578 (2006) 59–74
31
Snps
69
70
B.D. Malhotra et al. / Analytica Chimica Acta 578 (2006) 59–74
like DNA stem–loop labeled with an electro-active reporter as
the hybridization sensing element [207]. Such an E-DNA sensor can be used to electrochemically detect a target-induced
conformation change in a biopolymer. The synthetic DNA [peptide nucleic acid (PNA)] due to its superior enzymatic stability
and hybridization characteristics coupled with unique biochemical and physico-chemical properties probe has been used to
estimate DNA target upto about 10 femto moles (fM) using
conductance measurements. Gaylord et al. have demonstrated
that fluorescently labeled PNA coupled to conducting poly
(9,9bis (6′ -N,N,N-trimethylammoniumhexylbromide)fluorine)co-phenylene and a nuclease can be used to identify single
nucleotide polymorphism (SNP) [209]. The addition of an
enzyme prior to the addition of a conjugated polymer has been
found to increase the DNA target lengths and the selectivity of
the assay. Besides this, conducting polymer/PNA system provides added sensitivity to standard single fluorophore-labeled
probes without the need for complex/expensive optical setups for identification of neurodegenerative diseases. It may be
remarked that PNA is uniquely placed compared to many other
nucleic acid derivatives. Analysis using bead-probe techniques
and light-up probes reveal the degree of improvement that can be
achieved. In spite of the high cost of the synthesis, PNA provides
a unique opportunity for advanced high biosensor technology
with regards to sensitivity, stability, automation and system integration.
Zhu et al. have developed a novel and sensitive electrochemical DNA biosensor based on electrochemically fabricated
polyaniline nanowire and methylene blue for detection of DNA
hybridization [212]. The detection limit of the DNA biosensor
for discrimination of complementary and non-complementary
DNA sequences was found to be 1 × 10−12 mol L−1 .
Taira et al. have immobilized single stranded DNA probe on
self-assembling polymer based on polyallyamine modified with
thioctic acid for hybridization assays [213]. The efficiency and
selectivity of hybridization assay could be affected by adjusting the ionic strength of sodium chloride. Ioannou et al. have
reported an electrochemical DNA biosensor based on carbon
paste electrodes modified by a conducting composite for investigation of DNA-drug interactions [214]. Lori et al. developed
a conducting polymer surface by doping N-nitriloacetic acid
(NTA) into the electropolymerized polyvinylsulphonate doped
polyaniline (PANI/PVS) at a screen printed carbon electrode
for the immobilization of his-tagged biomolecules [215]. The
resulting NTA-PANI-/PVS film was shown to have interesting
electrochemical properties.
Crucifix et al. have investigated the immobilization of double stranded DNA molecules and of nucleoprotein complexes
on 2D-strepavidin crystals [216]. The binding efficiency and
specificity were examined using radioactively labeled oligonucleotides by direct visualization of unstained and hydrated
nucleic acid molecules using cryo-electron microscopy. It is
shown that RNA polymerase, once immobilized, femtomolar
amounts of DNA template can suitably interact with the enzyme.
Lassalle et al. have investigated the electrosynthesis of conducting copolymer using pyrrole and pyrrole-oligonucleotide
(ODN) at platinum disc electrodes [192]. The poly (pyrrole/
pyrrole-ODN) films could be used to detect DNA hybridization in real time using quartz crystal microbalance (QCM). The
affinity between avidin and biotin modified DNA target was
used to validate the effectiveness of transduction by fluorescence
microscopy [217]. Lee et al. have developed an electrochemical method to directly detect DNA hybridization using a new
conducting polymer, which was polymerized on the glassy carbon electrode with a terthiophene monomer having a carboxyl
group (3′ -carboxyl-5,2′ ,5′ ,2′′ -terthiophene) [218]. The highest
difference in admittance was seen at 1 kHz before and after
hybridization. Peng et al. have reported a novel oligonucleotide
(ODN) sensor based on electropolymerization of polypyrrole
in the presence of a sample containing ODN(s) [240]. The
resulting trapped ODN(s) have been probed by addition of complementary sequence ODN using impedance spectroscopy. The
impedance signal was found to be almost linear in the range of
3.7–370 nM with detection limit of ∼1 nM. By incorporating
cadmium sulphide (CdS) nanoparticles with the probe, a significant improvement in sensitivity was obtained. Cai et al. have
demonstrated that electrochemical impedance spectroscopy can
be used to visualize charge transfer through conducting polypyrrole films loaded with oligonucleotides probes formed on the
carbon nanotube modified electrodes as a basis for reagentless
protocol [219]. This technique has several advantages such as
high selectivity, reduced reaction time without the use of mediators or fluorescent materials for complementary and mismatched
target sequences.
Table 1 gives the characteristics of the various conducting
polymer based biosensors discussed in the above sections.
4. Conclusions
An attempt has been made to present an overview on the
prospects of conducting polymers in biosensors reported since
2000 till date. The extended systems in conjugated polymers,
highly susceptible to oxidation or reduction, provide precise control to the electrical and optical properties as biochemical reactions are often reversible in nature. The electro-active property of
these polymers makes them good candidates for various biosensing applications including drug release systems. Conducting
polymers provide an excellent opportunity for highly selective, specific, stable, economic and handy biosensing devices.
Many efforts are, however, required before these interesting
conducting biomolecular electronic materials are widely used
in biosensor industries.
Acknowledgements
We are grateful to Dr. Vikram Kumar, Director, NPL, New
Delhi, India for his interest in this work. We thank Mr. Sunil
K. Arya and other members of the group for interesting discussions. B.D.M. thanks the Department of Biotechnology, Govt. of
India, the Department of Science & Technology, Govt. of India
and the Indian National Science Academy, New Delhi, for the
financial assistance relating to the participation at the Pacifichem
2005.
B.D. Malhotra et al. / Analytica Chimica Acta 578 (2006) 59–74
References
[1] S. Cosnier, Ch. Gondran, Analysis 27 (1999) 558.
[2] S. Dong, X. Chen, Rev. Mol. Biotechnol. 82 (2002) 303.
[3] S.K. Arya, P.R. Solanki, R.P. Singh, M.K. Pandey, M. Datta, B.D. Malhotra, Talanta (2006), in press.
[4] A.B. Sanghvi, K.P.-H. Miller, A.M. Belcher, C.E. Schmidt, Nat. Mater. 4
(2005) 496.
[5] V. Saxena, B.D. Malhotra, in: B.D. Malhotra (Ed.), Advances in Biosensors, Elsevier, Netherlands, 2003, p. 63.
[6] M. Komiyama, S. Ye, X. Liang, Y. Yamamoto, T. Tomita, J.-M. Zhou, H.
Aburatani, J. Am. Chem. Soc. 125 (2003) 3758.
[7] S. Milardovic, Z. Grabaric, B.S. Grabaric, Food Technol. Biotechnol. 38
(2000) 203.
[8] J.D. Rasinger, G. Marrazza, F. Briganti, A. Scozzafava, M. Mascini, A.P.F.
Turner, Anal. Lett. 38 (2005) 1531.
[9] N. Adanyi, M. Varadi, N. Kim, I. Szendro, Curr. Appl. Phys. 6 (2006)
279.
[10] S. Suman, R. Singhal, A.L. Sharma, B.D. Malhotra, C.S. Pundir, Sens.
Actuators B 107 (2005) 768.
[11] R.L. Green, G.B. Street, L.J. Suter, Phys. Rev. Lett. 34 (1975) 577.
[12] D.M. Ivory, G.G. Miller, J.M. Sowa, L.W. Shacklette, R.R. Chance, R.H.
Baughman, J. Chem. Phys. 71 (1979) 1506.
[13] J.F. Rabolt, T.C. Clarke, K.K. Kanazawa, J.R. Reynolds, G.B. Street, J.
Chem. Soc. Chem. Commun. (1980) 347.
[14] A.F. Diaz, K.K. Kanazawa, G.P. Gardini, J. Chem. Soc. Chem. Commun.
(1979) 635.
[15] R. Saraswathi, M. Gerard, B.D. Malhotra, J. Appl. Polym. Sci. 74 (1999)
145.
[16] T. Kawai, T. Kuwabara, S. Wang, K. Yoshino, Jpn. J. Appl. Phys. 29
(1990) 602.
[17] H.H. Weetall, A. Druzhko, A. Rde Lera, R. Alvarez, B. Robertson, Bioelectrochemistry 51 (2000) 27.
[18] B. Zinger, L.L. Miller, J. Am. Chem. Soc. 106 (1984) 6861.
[19] A.J. Heeger, in: T.A. Skotheim (Ed.), Handbook of Conducting Polymers,
vol. II, Marcel Dekker, New York, 1986, p. 729.
[20] S.B. Adeloju, Analyst 121 (1996) 699.
[21] P.N. Bartlett, P.R. Birkin, Synth. Met. 61 (1993) 15.
[22] W.J. Sung, Y.H. Bae, Anal. Chem. 72 (2000) 2177.
[23] A. Heller, Acc. Chem. Res. 23 (1990) 128.
[24] P.R. Unwin, A.J. Bard, Anal. Chem. 64 (1992) 113.
[25] J. Svorc, S. Miertu, J. Katrlik, M. Stred’ansk, Anal. Chem. 69 (1997)
2086.
[26] G. Urban, A. Jachimowicz, F. Kohl, H. Kuttner, F. Olcaytug, H. Kamper,
F. Pittner, E.M. Buxbaum, T. Schalkhammer, O. Prohaska, M. Schonauer,
Sens. Actuators A 21–23 (1990) 650.
[27] D. Lewenstam, J. Babacka, A. Ivaska, J. Electroanl. Chem. 368 (1994)
23.
[28] K. Domansky, D.L. Baldwin, J.W. Grate, T.B. Hall, J. Li, M. Josowicz,
J. Janata, Anal. Chem. 70 (1998) 473.
[29] W. Lu, H. Zhao, G.G. Wallace, Anal. Chim. Acta 315 (1995) 27.
[30] A.K. Mulchandani, C.L. Wang, Electroanalysis 8 (1996) 414.
[31] M. Situmorang, D.B. Hibbert, J.J. Gooding, Electroanalysis 12 (2000)
111.
[32] P.N. Bartlett, R.G. Whitaker, Biosensors 3 (1988) 359.
[33] A. Gambhir, M. Gerard, A.K. Mulchandani, B.D. Malhotra, Appl.
Biochem. Biotechnol. 96 (2001) 249.
[34] B.A. Gregg, A. Heller, Anal. Chem. 62 (1990) 258.
[35] H.H. Weetall, K. Rogers, Anal. Lett. 35 (2002) 1342.
[36] M.S. Wrighton, Science 231 (1986) 32.
[37] G.P. Kittlesen, H.S. White, M.S. Wrighton, J. Am. Chem. Soc. 106 (1984)
7389.
[38] E.W. Paul, A.J. Ricco, M.S. Wrighton, J. Phys. Chem. 89 (1985) 1441.
[39] P. De Taxis du Poet, S. Miyamoto, T. Murakami, J. Kimura, I. Karube,
Anal. Chim. Acta 235 (1990) 255.
[40] M. Umana, J. Waller, Anal. Chem. 64 (1986) 2979.
[41] N.C. Foulds, C.R. Lowe, J. Chem. Soc. Faraday Trans. I 82 (1986)
1259.
71
[42] C. Iwakura, Y. Kajiya, H. Yoneyama, J. Chem. Soc. Chem. Commun.
(1988) 1019.
[43] J. Wang, M. Musameh, Anal. Chem. 75 (2003) 2075.
[44] J. Chen, A.K. Burrell, G.E. Collis, D.L. Officer, G.F. Swiegers, C.O. Too,
G.G. Wallace, Electrochim. Acta 47 (2002) 2715.
[45] V. Syritski, R.E. Gyurcsanyi, A. Öpik, K. Toh, Synth. Met. 152 (2005)
133.
[46] V. Saxena, B.D. Malhotra, in: B.D. Malhotra (Ed.), Handbook of Polymers
in Electronics, Rapra Technology Pvt. Ltd., UK, 2002, p. 3.
[47] J. Tschmelak, P. Guenther, J. Riedt, J. Kaiser, P. Kraemmer, L. Barzaga,
J.S. Wilkinson, P. Hua, J.P. Hole, R. Nudd, M. Jacson, R. Abuknesh, D.
Barcelo, S.R. Mozaz, M.J. Lopez de Alda, F. Sacher, J. Stien, J. Slobodnik,
P. Oswald, H. Kozmenko, E. Korenkova, L. Tothova, Z. Krascsenits, G.
Gauglitz, Biosens. Bioelect. 20 (2005) 1499.
[48] L. Zhai, R.D. McCullogh, J. Mater. Chem. 14 (2004) 141.
[49] D. Cunliffe, S. Pennadam, D.L. Raffa, F. Battaglini, J. Electroanal. Chem.
504 (2001) 120;
C. Alexander, Eur. J. 40 (2004) 5.
[50] P.C. Pandey, J. Chem. Soc. Faraday Trans. I 84 (1988) 2259.
[51] G.E. Collis, A.K. Burrell, S.M. Scott, D.L. Officer, J. Org. Chem. 68
(2003) 8974.
[52] W.E. Price, G.G. Wallace, H. Zhao, J. Membr. Sci. 87 (1994) 47.
[53] N. Mirmohseni, W.E. Price, G.G. Wallace, Polym. Gels Networks 1
(1993) 61.
[54] P.R. Teasdale, G.G. Wallace, Analyst 118 (1993) 329.
[55] M. Trojanowicz, T.K. vel Krawcyzk, Chem. Anal. 42 (1995) 199.
[56] B. Adhikari, S. Majumdar, Prog. Polym. Sci. 29 (2004) 699.
[57] M. Gerard, A. Chaubey, B.D. Malhotra, Biosens. Bioelectron. 17 (2002)
345.
[58] M. Situmorang, J.J. Gooding, D.B. Hibbert, D. Barnett, Electroanalysis
14 (2002) 17.
[59] A. Chaubey, B.D. Malhotra, Biosens. Bioelectron. 17 (2002) 441.
[60] M.S. Wrighton, H.S. White Jr., G.P. Kittlesen, Molecule-based microelectronic devices, US Patent, 192 (1991) 5034.
[61] C.P. Sun, J.C. Liao, Y.H. Zhang, V. Gau, M. Mastali, J. Babbitt, W. Grundfest, B.M. Churchill, E.R.B. McCabe, D.A. Haake, Mol. Gen. Metab. 84
(2005) 90.
[62] E. Cass, G. Davis, G.D. Francis, H.A.O. Hill, W.J. Alton, I.J. Higgins, E.V. Plotkin, L.D.L. Scott, A.P.F. Turner, Anal. Chem. 56 (1984)
667.
[63] K.F. Fu, W.J. Huang, Y. Lin, D.H. Zhang, T.W. Hanks, A.M. Rao, Y.P.
Sun, J. Nanosci. Nanotechnol. 2 (2002) 457.
[64] K.F. Fu, W.J. Huang, Y. Lin, D.H. Zhang, T.W. Hanks, A.M. Rao, Y.P.
Sun, Nano. Lett. 2 (2002) 311.
[65] L. Gorton, A. Lindgren, F.D. Muntaenu, T. Ruzgas, I. Gazaryan, Anal.
Chim. Acta 400 (1999) 91.
[66] W. Schuhmann, Rev. Mol. Biotechnol. 82 (2002) 425.
[67] T. Ramanathan, F.T. Fisher, R.S. Ruoff, I.C. Brinson, Chem. Matter. 17
(2005) 1290.
[68] S. Varfolomeyev, I. Kurochkin, A. Ermenko, E. Efremenko, Pure Appl.
Chem. 74 (2002) 2316.
[69] C.M. Li, W. Chen, X. Yang, C.Q. Sun, C. Gao, Z.X. Zhang, J. Sawyer,
Front. Biosci. 10 (2005) 2518.
[70] P.S. Heeger, A.J. Heeger, PNAS 96 (1999) 12219.
[71] P.N. Bartlett, R.G. Whitaker, J. Electroanal. Chem. 224 (1987) 37.
[72] Z. Sun, H. Tachikawa, Anal. Chem. 64 (1992) 1112.
[73] B. Yu, Y. Moussy, F. Moussy, Electroanalysis 17 (2005) 1771.
[74] S. Kroger, A.P.F. Turner, K. Mosbach, K. Haupt, Anal. Chem. 71 (1999)
3698.
[75] H. Shiigi, K. Okamura, D. Kijima, A. Hironaka, B. Deore, U. Sree, T.
Nagoka, Electrochem. Solid State 6 (2003) H1.
[76] K.-C. Ho, W.M. Yeh, T.S. Tung, J.-Y. Liao, Anal. Chim. Acta 542 (2005)
90.
[77] S. Geetha, C.R.K. Rao, M. Vijyan, D.C. Trivedi, Anal. Chim. Acta 568
(2006) 119.
[78] S. Andreescu, O.A. Sadik, Pure Appl. Chem. 76 (2004) 861.
[79] S. Cosnier, Electroanalysis 17 (2005) 1701.
[80] S. Cosnier, Anal. Bioanal. Chem. 377 (2003) 507.
72
B.D. Malhotra et al. / Analytica Chimica Acta 578 (2006) 59–74
[81] A. Ramanaviciene, A. Ramanavicius, Advanced Biomaterials for Medical Applications, Kluwer Academic Publishers, Netherlands, 2004,
p. 111.
[82] A. Malinauskas, J. Malinauskiene, A. Ramanavicius, Nanotechnology 16
(2005) R51.
[83] A.K. Wanekaya, W. Chen, N.V. Myung, A. Mulchandani, Electroanalysis
18 (2006) 533.
[84] T. Ming-Huang Lee, I.-M. Hsing, Anal. Chim. Acta 556 (2006) 26.
[85] I.M. Mackay, K.E. Arden, A. Nitsche, Nucleic Acids Res. 30 (2002) 6.
[86] L.A. Terry, S.F. White, L.J. Tigwell, J. Agric. Food Chem. 53 (2005)
1309.
[87] T.G. Drummond, M.G. Hill, J.K. Barton, Nat. Biotechnol. 21 (2003) 1192.
[88] K. Habermuller, M. Mosbach, W. Schuhmann, Fresnius J. Anal. Chem.
366 (2000) 560.
[89] K. Kerman, M. Kobayashi, E. Tamiya, Meas. Sci. Technol. 15 (2004) R1.
[90] K. Arora, S. Chand, B.D. Malhotra, Anal. Chim. Acta 568 (2006)
259.
[91] S. Singh, P.R. Solanki, M.K. Pandey, B.D. Malhotra, Anal. Chim. Acta
568 (2006) 126.
[92] M. Gerard, B.D. Malhotra, Curr. Appl. Phys. 5 (2005) 174–177.
[93] S.K. Sharma, R. Singhal, B.D. Malhotra, N. Sehgal, A. Kumar, Electrochim. Acta 49 (2004) 2479.
[94] A.L. Sharma, R. Singhal, A. Kumar, K.K. Pande, B.D. Malhotra, J. Appl.
Polym. Sci. 91 (2004) 3999.
[95] S.K. Sharma, R. Singhal, B.D. Malhotra, N. Sehgal, A. Kumar, Biotechnol. Lett. 26 (2004) 645.
[96] S.K. Sharma, R. Singhal, B.D. Malhotra, N. Sehgal, A. Kumar, Biosens.
Bioelectron. 20 (2004) 651.
[97] B.D. Malhotra, R. Singhal, Pramana 61 (2003) 331.
[98] B.D. Malhotra, A. Chaubey, Sens. Actuators B 91 (2003) 17.
[99] V. Saxena, B.D. Malhotra, Curr. Appl. Phys. 3 (2003) 293.
[100] A.L. Sharma, S. Annapoorni, B.D. Malhotra, Curr. Appl. Phys. 3 (2003)
3.
[101] A.L. Sharma, S. Annapoorni, B.D. Malhotra, Polymer 42 (2001) 8307.
[102] A. Chaubey, M. Gerard, V.S. Singh, B.D. Malhotra, Appl. Biochem.
Biotechnol. 96 (2001) 293.
[103] R. Doong, H. Shih, S. Lee, Sens. Actuators B 111–112 (2005) 323.
[104] A.L. Sharma, M. Gerard, R. Singhal, B.D. Malhotra, S. Annapoorni, Appl.
Biochem. Biotechnol. 96 (2001) 155.
[105] A.L. Sharma, V. Saxena, S. Annapoorni, B.D. Malhotra, J. Appl. Polym.
Sci. 81 (2001) 1460.
[106] K. Ramanathan, S.S. Pandey, R. Kumar, Bansi D. Malhotra, A.S.N.
Murthy, J. Appl. Polym. Sci. 78 (2000) 662.
[107] A. Chaubey, K.K. Pande, V.S. Singh, B.D. Malhotra, Anal. Chim. Acta
407 (2000) 97.
[108] S. Tombelli, M. Mascini, C. Sacco, A.P.F. Turner, Anal. Chim. Acta 418
(2000) 1.
[109] T. Roopa, N. Kumar, S. Bhattacharya, G.V. Shivshanakar, Biophys. J. 87
(2004) 974.
[110] J.H. Ng, L.L. Ilag, J. Cell. Mol. Med. 6 (2002) 329.
[111] S.L. Beaucage, Curr. Med. Chem. 89 (2001) 1231.
[112] J. Wang, Analyst 130 (2005) 421.
[113] J.Q. Brown, J. McShane, Biosens. Bioelectron. 21 (2006) 1760.
[114] R.A. Potyrailo, Angew. Chem. Int. Ed. 45 (2006) 702.
[115] P.N. Bartlett, P. Tebbutt, C. Tyrrel, Anal. Chem. 64 (1992) 138.
[116] V. Gau, S.C. Ma, H. Wang, J. Tsukuda, J. Kibler, D.A. Haake, Methods
37 (2005) 73.
[117] D. Belanger, J. Nadreau, G. Fortier, J. Electroanal. Chem. 274 (1989) 143.
[118] D. Pan, J. Chen, S. Yao, W. Tao, L. Nie, Anal. Sci. 21 (2005) 367.
[119] A. Curulli, F. Valentini, S. Orlanducci, M.L. Terranova, G. Palleschi,
Biosens. Bioelectron. 20 (2004) 1223.
[120] R. Singhal, A. Chaubey, K. Kaneto, W. Takashima, B.D. Malhotra,
Biotechnol. Bioeng. 85 (2004) 277.
[121] R. Singhal, B.D. Malhotra, Asian J. Phys. 14 (2005) 151.
[122] A. Callegavi, S. Cosnier, M. Marcaccio, D. Paolucci, F. Paolucci, V. Creorgakilas, N. Tagmatarchis, E. Vazquez, M. Prato, J. Mater. Chem. 14
(2004) 807.
[123] Y.-M. Uang, T.-C. Chou, Biosens. Bioelectron. 19 (2003) 141.
[124] D.D. Borole, U.R. Kapadi, P.P. Mahulikar, D.G. Hundiwale, Polym. Adv.
Technol. 15 (2003) 306.
[125] S. Sekar, I. Becerik, Electroanalysis 16 (2004) 1542.
[126] A. Ramanavicius, A. Kausaite, A. Ramanavicience, J. Acaite, A. Malinauskas, Synth. Met. 156 (2006) 409.
[127] Z.M. Tahir, E.C. Alocilija, D.L. Grooms, Biosens. Bioelectron. 20 (2005)
1690.
[128] A. Arslan, S. Kiralp, L. Toppare, Y. Yagci, Int. J. Biol. Macromole. 35
(2005) 163.
[129] S.D. Varfolomeev, I.N. Kurochkin, Biosens. Bioelectron. 11 (1996) 863.
[130] L. Jiang, H.K. Jun, Y.S. Hoh, J.O. Lim, D.D. Lee, J.S. Huh, Sens. Actuators B 105 (2005) 132.
[131] F. Tian, B. Xu, G. Zhu, L. Zhu, Anal. Chim. Acta 443 (2001) 9.
[132] G. Drummond, M.G. Hill, J.K. Barton, Nat. Biotechnol. 21 (2003) 10.
[133] A. Crulli, F. Valentini, S. Orlanduci, M.L. Terranova, G. Palleschi,
Biosens. Bioelectron. 20 (2004) 1223.
[134] J. Kan, X. Pan, C. Chen, Biosens. Bioelectron. 19 (2004) 1635.
[135] M.A. Rahman, D.S. Park, Y.B. Shim, Biosens. Bioelectron. 19 (2004)
1565.
[136] M. Trojanowicz, A. Miernik, Electrochim. Acta 46 (2001) 1053.
[137] S. Myler, F. Davis, S.D. Collyer, S.P.J. Higson, Biosens. Bioelectron. 20
(2004) 408.
[138] A.C. Barton, S.D. Collyer, F. Davis, D.D. Gornall, K.A. Law, E.C.D.
Lawrence, D.W. Mills, S. Myler, J.A. Prtichand, M. Thomson, S.P.J. Higson, Biosens. Bioelectron. 20 (2004) 408.
[139] F. He, Y. Tang, M. Yu, S. Wang, Y. Li, D. Zhu, Adv. Funct. Mater. 16
(2006) 91.
[140] C. Vedrine, S. Fabiano, C. Tran-Minh, Talanta 59 (2003) 535.
[141] A.E. Böyükbayram, S. Kiralp, L. Toppare, Y. Yaĝci, Bioelectrochemistry
69 (2006) 164.
[142] J. Haccoun, B. Piro, V. Noel, M.C. Pham, Biochemistry 68 (2005) 223.
[143] P.N. Bartlett, E. Simon, C.S. Toh, Bioelectrochemistry 56 (2002) 117.
[144] Y. Lu, J. Liu, P.J. Bruesehoff, C.M.B. Parot, A.K. Brouwn, Biosens.
Bioelctron. 18 (2003) 529.
[145] A. Chaubey, K.K. Pande, V.S. Singh, B.D. Malhotra, Anal. Chim. Acta
407 (2000) 97.
[146] A. Chaubey, K.K. Pande, B.D. Malhotra, Anal. Sci. 19 (2003) 1477.
[147] S. Singh, A. Chaubey, B.D. Malhotra, J. Appl. Polym. Sci. 91 (2004)
3769.
[148] S. Singh, A. Chaubey, B.D. Malhotra, Anal. Chim. Acta 502 (2004) 229.
[149] P. Asberg, O. Inganas, Biosens. Bioelectron. 19 (2003) 199.
[150] K. Grennan, A.J. Killard, M.R. Smyth, Electroanalysis 17 (2005) 1360.
[151] A. Morrin, O. Ngamna, A.J. Kllard, S.E. Moulton, M.E. Smyth, G.G.
Wallace, Electroanalysis 17 (2005) 423.
[152] J. Jia, B. Wang, A. Wu, G. Cheng, Z. Li, S. Dong, Anal. Chem. 74 (2002)
2217.
[153] O. Ngamna, A. Morrin, S.E. Moulton, A.J. Killrad, M.R. Smyth, G.G.
Wallace, Synth. Met. 153 (2005) 185.
[154] N.G.R. Mathebe, A. Morrin, E.I. Iwuoha, Talanta 64 (2005) 115.
[155] B. Zhou, R. Sun, X. Hu, L. Wang, H. Wu, S. Song, C. Fan, Int. J. Mol.
Sci. 6 (2005) 303.
[156] A. Lindgren, T. Ruzgas, L. Gorton, E. Csoregi, G.B. Ardila, I.Yu.
Sakharov, I.G. Gazaryan, Biosens. Bioelectron. 15 (2000) 491.
[157] Z. Dai, S. Liu, H. Ju, Electrochim. Acta 49 (2004) 2139.
[158] S. Brahim, D. Narinesinh, A. Gueseppi-Elie, Biosens. Bioelectron. 17
(2005) 53.
[159] D.G. Pijanowska, M. Dawgul, T. Torbiz, Sensors 3 (2003) 160.
[160] A. Zhang, Y. Hou, N. Jaffrezic-Renault, J. Wan, A. Soldatkin, J.-M. Chovelon, Bioelectrochemistry 569 (2002) 157.
[161] A. Gambhir, A. Kumar, B.D. Malhotra, B. Miksa, S. Slomkowski, ePolymers 52 (2002) 1.
[162] R. Singhal, A. Gambhir, S. Annapoorni, B.D. Malhotra, Biosens. Bioelectron. 7 (2002) 697.
[163] Rajesh, V. Bisht, W. Takashima, K. Kaneto, Biomaterials 26 (2005) 3683.
[164] A.V. Rebriiev, N.F. Starodub, Electroanalysis 16 (2004) 1891.
[165] A. Ramanavicius, A. Kausaite, A. Ramanaviciene, Sens. Actuators B 111
(2005) 532.
[166] S. Mu, J. Kan, J. Zhou, J. Electroanl. Chem. 370 (2002) 135.
B.D. Malhotra et al. / Analytica Chimica Acta 578 (2006) 59–74
[167] E. Smela, Adv. Mater. 15 (2003) 481.
[168] M.R. Pinto, K.S. Schanze, PNAS 101 (2004) 7505.
[169] A.N. Ivanov, G.A. Evtugyn, L.V. Lukachova, E.E. Karyakina, H.C. Budnikov, S.G. Kiseleva, A.V. Orlov, G.P. Karpacheva, A.A. Karyakin, IEEE
Sens. J. 3 (2003) 333.
[170] S. Cosnier, D. Fologea, S. Szunerits, R.S. Marks, Electrochem. Commun.
2 (2000) 827.
[171] S. Cosnier, A.L. Pellec, R.S. Marks, K. Perie, J.-P. Lellouche, Electrochem. Commun. 5 (2003) 973.
[172] M.A. Rahman, D.-S. Park, S.-C. Chang, C.J. McNeil, Y.-B. Chang, C.J.
McNeil, Y.-B. Shim, Biosens. Bioelectron. 21 (2006) 1116.
[173] J.J. Langer, M. Filipiak, J. Kecinska, J. Jasnowska, J. Wlodareczak, B.
Buladowski, Surf. Sci. 573 (2004) 140.
[174] F. Qu, M. Yang, J. Jiang, G. Shen, R. Yu, Anal. Biochem. 344 (2005) 108.
[175] O.A. Biloivan, S.V. Vervka, S.V. Dzyavych, J.-N. Zine, J. Bausells, J.
Samitier, A. Errachid, Mater. Sci. Eng. C 26 (2006) 574.
[176] Q. Fang, D.G. Chetwynd, J.W. Gardner, C. Toh, P.N. Bartlett, Mater. Sci.
Eng. A 355 (2003) 62.
[177] N.Y. Kwok, S. Dongb, W. Loa, K.Y. Wong, Sens. Actuators B 110 (2005)
289.
[178] Y. Lei, P. Mulchandani, W. Chen, A. Mulchandani, J. Agri. Food Chem.
53 (2005) 524.
[179] A. Ramanavicius, A. Finkelsteinas, A. Ramanavicius, Mater. Sci. 10
(2004) 1392.
[180] A.I. Minet, J.N. Barisci, G.G. Wallace, Anal. Chim. Acta 475 (2003) 37.
[181] R.A. Porter, J. Immunoassay 21 (2000) 51.
[182] S. Grant, F. Dauris, K.A. Law, A.C. Berton, S.D. Collyer, S.P.J. Higson,
T.D. Gibson, Anal. Chim. Acta 537 (2005) 163.
[183] J.J. Gooding, C. Wasiowych, D. Barnett, D.B. Hibbert, J.N. Barisci, G.G.
Wallace, Biosens. Bioelectron. 20 (2004) 260.
[184] A. Ramanaviciene, A. Ramanavicius, Crit. Rev. Anal. Chem. 32 (2002)
245.
[185] L. Zhang, R. Yuan, X. Huang, Y. Chai, D. Tang, S. Cao, Anal. Bioanal.
Chem. 381 (2005) 1036.
[186] Z.M. Tahir, E.C. Alcocilza, D.L. Grooms, Biosens. Bioelectron. 20 (2005)
1690.
[187] G. Marrazza, I. Chianella, M. Mascini, Biosens. Bioelectron. 14 (1999)
43.
[188] F. Darain, S.-U. Park, Y.-B. Shim, Biosens. Bioelectron. 18 (2003)
773.
[189] G. Lillie, P. Payne, P. Vadgama, Sens. Actuators 78 (2001) 249.
[190] G. Farace, G. Lillie, T. Hianik, P. Payne, P. Vadgama, Bioelectrochemistry
55 (2002) 1.
[191] P. Leonard, Enzyme Microbiol. Technol. 32 (2003) 3.
[192] N. Lassalle, P. Mailley, E. Vieil, T. Livache, A. Roget, J.P. Correia, L.M.
Abrantes, J. Electroanal. Chem. 509 (2001) 48.
[193] A. Gambhir, M. Gerard, S.K.J. Ain, B.D. Malhotra, Appl. Biochem.
Biotech. 96 (2001) 303.
[194] X. Yu, S.N. Kim, F. Papadimitrkopoulos, J.M. Ruling, Mol. Biosyst. 1
(2005) 70.
[195] S. Palanti, G. Marrazza, M. Mascini, Anal. Lett. 29 (1996) 2309.
[196] A.G. Frutos, S. Pal, M. Quesada, J. Lahiri, J. Am. Chem. Soc. 124 (2002)
2396.
[197] F. Patolsky, A. Lichtenstein, J. Am. Chem. Soc. 123 (2001) 5194.
[198] R.G. Endres, Colloquium: Rev. Modern Phys. 76 (2004) 195.
[199] J. Cha, J.I. Han, Y. Choi, D.S. Yoon, K. Who, G. Lim, Biosens. Bioelectron. 18 (2003) 1241.
[200] P. Dutta, S.K. Mandal, J. Phys. 37 (2005) 2908.
[201] S. Zhang, G. Wright, Y. Yang, Biosens. Bioelectron. 15 (2000) 273.
[202] F. Lucarelli, G. Marrazza, A.P.F. Turner, M. Mascini, Biosens. Bioelectron. 19 (2004) 515.
[203] H.K. Youssoufi, B. Makrouf, Anal. Chim. Acta 469 (2002) 85.
[204] T. Gu, Y. Hasebe, Anal. Chim. Acta 525 (2004) 191.
[205] L. Shi, I. Willner, Electrochem. Commun. 6 (2004) 1057.
[206] J. Wu, Y. Zhou, X. Li, H. Liu, G. Shen, R. Yu, Sens. Actuators B 104
(2005) 43.
[207] C. Fan, K.W. Plaxo, A.J. Heeger, PNAS 100 (2003) 9134.
[208] J. Hahm, C.M. Lierber, Nano Lett. 4 (2003) 51.
73
[209] B.S. Gaylord, M.R. Massie, S.C. Feinstein, G.C. Bazan, PNAS 102 (2005)
34.
[210] C. Ban, S. Chung, D.-S. Park, Y.-B. Shim, Nucl. Acids Res. 32 (2004)
1.
[211] K. Kerman, Y. Matsubara, Y. Morita, Y. Takamura, E. Tamiya, Sci. Technol. Adv. Mater. 5 (2004) 351.
[212] N. Zhu, Z. Chang, P. He, Y. Fang, Electrochim. Acta 18 (2006)
3758.
[213] S. Taira, K. Yokoyama, Biotechnol. Bioeng. 89 (2005) 835.
[214] A.K. Ioannou, A.A. Pantazaki, S.T.H. Girousi, M.-C. Millet, C.V. Madjar,
A.N. Voulgaropoulos, Electroanalysis 18 (2006) 456.
[215] J. Lori, A. Morrin, A.J. Killard, M.R. Smyth, Electroanalysis 18 (2006)
77.
[216] C. Crucifix, M. Uhring, P. Schultz, J. Struct. Biol. 146 (2004) 441.
[217] N. Lassale, A. Roget, T. Livache, P. Mailley, E. Vieil, Talanta 55 (2001)
993.
[218] T.-Y. Lee, Y.-B. Shim, Anal. Chem. 73 (2001) 5629.
[219] H. Cai, Y. Xu, P.-G. He, Y.-Z. Fang, Electroanalysis 15 (2003)
1864.
[220] J. Livage, T. Coradin, C. Roux, J. Phys. Condens. Matter 13 (2001)
R673.
[221] J. Wang, L. Chen, A. Mulchandani, P. Mulchandani, W. Chen, Electroanalysis 11 (1999) 866.
[222] J.C. Vidal, S. Esteban, J. Gil, J.R. Castillo, Talanta 68 (2006) 791.
[223] A. Ramanavicius, K. Habermuller, E. Csoregi, V. Laurinavicius, W.
Schuhmann, Anal. Chem. 71 (1999) 3581.
[224] W. Schuhmann, R. Lammert, B. Uhe, H.-L. Schmidt, Sens. Actutaors 1
(1990) 537.
[225] K. Ramanathan, M. Bangar, M. Yun, W. Chen, A. Mulchandani, N.V.
Myung, Nano Lett. 4 (2004) 1237.
[226] K. Ramanathan, M. Bangar, M. Yun, W. Chen, A. Mulchandani, N.V.
Myung, J. Am. Chem. Soc. 127 (2005) 496.
[227] O. Brandt, J.D. Hoheisel, Trends Biotechnol. 22 (2004) 617.
[228] M. Mravcakova, K. Boukerma, M. Omastova, M.M. Chehimi, Mater. Sci.
Eng. C 26 (2006) 306.
[229] K. Boukerma, J.Y. Piquemal, M.M. Chehimi, M. Mravcakova, M. Omastova, P. Beaunier, Polymer 47 (2006) 569.
[230] M. Okochi, H. Ohta, T. Taguchi, H. Ohta, T. Matsunaga, Electrochim.
Acta 51 (2005) 952.
[231] J. Li, J.-D. Huang, B.-Y. Wu, Q. Chen, Acta Pharmacol. Sin. 28 (2005)
1212.
[232] S. Carrara, V. Bavastreello, D. Ricci, E. Stura, C. Nicolini, Sens. Actuators
B 109 (2005) 221.
[233] K. Karnicka, M. Chojak, K. Miecznikowski, M. Skunik, B. Barnowska,
A. Kolary, A. Piranska, L. Adamczyk, P.J. Kulesza, Bioelectrochemistry
66 (2005) 79.
[234] R. Ade Barros, C.R. Martins, W.M. de Azevedo, Synth. Met. 155 (2005)
35.
[235] H.P. de Oliveira, C.A.S. Andrade, C.P. de Melo, Synth. Met. 155 (2005)
631.
[236] P. Murugaraj, D.E. Mainwaring, T. Jakubov, N.E. Mora-Huertas, N.A.
Khelil, R. Siegele, Solid State Commun. 137 (2006) 422.
[237] Z. Zhang, F. Wang, F. Chen, G. Shi, Mater. Lett. 60 (2006) 1039.
[238] E.R. Zubarev, J. Xu, A. Sayyad, J.D. Gibson, J. Am. Chem. Soc. 128
(2006) 4958.
[239] H. Peng, C. Soeller, M.B. Cannel, G.A. Bowmaker, R.P. Cooney, J.
Travas-Sejdic, Biosens. Bioelectron. 15 (2006) 1727.
[240] A. Ramanaviciene, W. Schuhmann, A. Ramanavicius, Colloids Surf. B:
Bioinform. 48 (2006) 159.
[241] S. Devouge, C. Salvagnini, J. M-Brynaert, Bioorg. Med. Chem. 15 (2005)
3252.
[242] Rajesh, V. Bisht, W. Takashima, K. Kaneto, Surf. Coat. Technol. 198
(2005) 231.
[243] J. Liu, M.N. Paddon-Row, J. Gooding, Chem. Phys. (2006), in press.
[244] Rajesh, V. Bisht, W. Takashima, K. Kaneto, React. Funct. Polym. 62
(2005) 51.
[245] P. Tengvall, E. Jansson, A. Askendal, P. Thomsen, C. Gretzer, Colloids
Surf. B: Bioinform. 28 (2003) 261.
74
B.D. Malhotra et al. / Analytica Chimica Acta 578 (2006) 59–74
[246] J.-I. Nzai, T. Hoshi, T. Osa, TrAc Trends Anal. Chem. 13 (1994)
205.
[247] L. Nobs, F. Buchegger, R. Gurny, E. Allemann, Eur. J. Pharm. Biopharm.
58 (2004) 483.
[248] R. Gref, P. Couvreur, G. Barratt, E. Mysiakine, Biomaterials 24 (2003)
4529.
[249] S. Singh, P.R. Solanki, M.K. Pandey, B.D. Malhotra, Sens. Actuators B
115 (2006) 534.