Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/222392344

Recent developments in urea biosensors

Article  in  Biochemical Engineering Journal · April 2009


DOI: 10.1016/j.bej.2008.07.004

CITATIONS READS

157 3,680

3 authors:

Gunjan Dhawan Gajjala Sumana


University of North Dakota National Physical Laboratory - India
19 PUBLICATIONS   808 CITATIONS    102 PUBLICATIONS   3,189 CITATIONS   

SEE PROFILE SEE PROFILE

Bansi Malhotra
Delhi Technological University
380 PUBLICATIONS   20,079 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Role of Gut in AD View project

Amyloid-b mediated microglial activation View project

All content following this page was uploaded by Gunjan Dhawan on 21 February 2018.

The user has requested enhancement of the downloaded file.


Biochemical Engineering Journal 44 (2009) 42–52

Contents lists available at ScienceDirect

Biochemical Engineering Journal


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

Review

Recent developments in urea biosensors


Gunjan Dhawan, Gajjala Sumana ∗ , B.D. Malhotra
Biomolecular Electronics & Conducting Polymer Research Group, National Physical Laboratory, Dr. K. S. Krishnan Marg, New Delhi 110012, India

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

Article history: This paper reviews recent developments in urea biosensors, as reported in the literature. The advantages
Received 10 March 2008 and roles of various matrices, different strategies for biosensor construction, analytical performance and
Received in revised form 16 May 2008 applications are discussed. The prospects of urea biosensors for medical applications are also discussed.
Accepted 1 July 2008
© 2008 Elsevier B.V. All rights reserved.

Keywords:
Urea
Urease
Matrix
Biosensor

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
2. Historical background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3. Urea biosensors based on polymeric matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.1. Urea biosensors based on non-conducting polymer matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.2. Urea biosensors based on conducting polymer matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
3.3. Sol–gel-based urea biosensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
4. Urea biosensors based on Langmuir–Blodgett films . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
5. Urea biosensors based on nanomaterials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
5.1. Nanoparticles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
5.2. Quantum dots (QDs) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
5.3. Nanofibers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
5.4. Nanoporous silicon technology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
6. Miscellaneous matrices for urea biosensing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
7. ISFET-based urea biosensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
8. Demands, challenges and commercialization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
9. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

levels. The normal level of urea in serum is from 15 to 40 mg/dl


1. Introduction
(2.5–7.5 mM/l). In patients suffering from renal insufficiency, urea
concentrations in serum vary from 180 to 480 mg/dl and, at ele-
Urea is widely distributed in nature and its analysis is of con-
vated levels above 180 mg/dl, hemodialysis is required. Apart from
siderable interest in clinical and agricultural chemistry [1,2]. It
clinical applications, there is a growing demand for robust, reli-
is known to be an important marker for evaluating uremic toxin
able instrumentation toward the estimation of urea in other fields
(e.g., food science and environmental monitoring). As the principal
component of non-protein nitrogen in cow milk, urea is utilized
∗ Corresponding author. Tel.: +91 011 25734273. as an indicator of protein-feeding efficiency [3]. Improved feeding
E-mail address: sumanagajjala@gmail.com (G. Sumana). practices through appropriate urea testing over a period of several

1369-703X/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.bej.2008.07.004
G. Dhawan et al. / Biochemical Engineering Journal 44 (2009) 42–52 43

months can result in lower feed costs, greater milk-protein yield, Enzymatic biosensors utilize the biospecificity of an enzymatic
increased reproductive performance, and minimal nitrogen excre- reaction. In the fabrication of a urea biosensor, urease is most
tion into the environment [4,5]. Urea also plays a strategic role in commonly used as the biosensing element. Enzymatic urea biosen-
the marine nitrogen cycle [6,7], including sources of excretion by sors utilize biochemical reactions. In other words, analyte (urea)
invertebrates and fish, and bacterial decomposition of nitrogenous and enzyme (urease) result in a product (ammonium ion) that
material and terrestrial drainage. Urea estimation is important dur- can be detected and quantified using a transducer (amperomet-
ing environmental monitoring. Annual world production of urea ric/potentiometric/optical thermal/piezoelectric). The enzymatic
exceeds 100 million metric tonnes, the majority of which is used reaction of urea with urease is shown in the following equation:
as fertilizer. Excessive nitrogen fertilizer applications can lead to Urease
pest problems by increasing the birth rate, longevity and overall NH2 CONH2 + 3H2 O −→ 2NH4 + + HCO−3 + OH− (1)
fitness of certain pests. Urea can be responsible for reduction in Many matrices (e.g., polymers, sol–gels, conducting poly-
soil pH. Many methods are available for urea estimation, including mers, Langmuir–Blodgett films, nanomaterials and self-assembled
gas chromatography [8], calorimetry [9], and fluorimetry [10]. How- monolayers (SAMs)) have been used to (i) provide support; (ii)
ever, these methods suffer from complicated sample pretreatment impart stability to biomolecules towards variations in temperature,
steps and are unsuitable for on-site monitoring. It is envisaged that pH and ionic strength; (iii) increase shelf-life; and (iv) reduce cost
biosensors can be utilized to overcome some of these problems. for the fabrication of a urea biosensor. For optimum biostability
Biosensors are analytical devices that incorporate biological and reaction efficiency, the preferred host matrix appears to be the
materials such as enzymes, nucleic acids (DNA and RNA), micro- one that isolates the desired biomolecule, protecting it from self-
organisms, whole cells, antibodies, cell receptors or biologically aggregation and microbial attack, while providing essentially the
derived materials, which in most cases are specific to an analyte in same local aqueous microenvironment as in biological media. In
intimate contact with a physico-chemical transducer. Transducers recent years, nanomaterials such as carbon nanotubes (CNTs), metal
are components that convert a biochemical signal into a measur- nanoparticles and semiconducting nanoparticles have emerged
able signal. The transducing microsystem may be electrochemical, as new classes of nanomaterials that are receiving considerable
thermometric, optical, piezoelectric or magnetic. Owing to their interest owing to their unique structure, high chemical stability
simplicity, high sensitivity and potential ability for real-time and and high surface-to-volume ratio. These properties make them
on-site analysis, biosensors have been widely applied in various extremely attractive for fabricating chemical sensors [14,15] (Lim
fields, including industrial processes, clinical detection and envi- et al., 2005; Wang et al., 2003). Recently, composite materials
ronmental control [11,12]. The immobilization technique and the based on conducting polymers, redox mediators, metal nanoparti-
support to impart stability, physical and chemical conditions (e.g., cles, nanocomposites and nanoclusters have been used to combine
pH, temperature, contaminants and thickness) play a major role properties of the individual components with a synergistic perfor-
in the performance of a biosensor. These interesting biodevices mance in biosensor fabrication. However, efforts are being made to
have been predicted to play a critical role towards the implanta- fabricate a cost-effective urea biosensor with improved sensitivity
tion of a chip into a body to monitor all biological parameters of and stability. The present article is an update on recent develop-
clinical importance, as well as for point-of-care devices [13]. There- ments concerning urea biosensors as reported in the literature.
fore, biosensors are presently the subject of extensive research
for applications like clinical diagnosis, food technology, military, 2. Historical background
industrial, biomedical and environmental monitoring using various
biomolecules (e.g., enzymes, antibodies, nucleic acids, receptors, In 1962, Professor Leland C. Clark Jr., father of the biosensor con-
bio-mimetic synthetic or semisynthetic molecules). The schematic cept, published a paper illustrating an experiment in which glucose
of a biosensor is shown in Fig. 1. oxidase was entrapped at a Clark oxygen electrode using a dialy-
sis membrane, resulting in a glucose biosensor. His development
became the basis of numerous variations on this basic design, as
well as many affinity-based biosensors (e.g., immunosensors and
nucleic acid biosensors), ultimately aiming to achieve lab-on-a-
chip microarrays [16]. Fig. 2 shows the schematic of a typical urea
biosensor.
Guilbault and Montalvo were the first to fabricate a potentio-
metric urease enzyme electrode for the estimation of urea through

Fig. 1. Schematic of biosensor. Fig. 2. Schematic of a typical urea biosensor.


44 G. Dhawan et al. / Biochemical Engineering Journal 44 (2009) 42–52

Fig. 3. Immobilization techniques used for the development of urea biosensors.

its enzyme-catalyzed hydrolysis [17,18]. The analytical signal of this and long-term stability as supports for the immobilization of the
early biosensor was the potential of a cation-selective glass elec- enzymes. There are a variety of methods by which enzymes can be
trode for detection of ammonium ion activity by urease-catalyzed immobilized, ranging from covalent chemical bonding to physical
hydrolysis. The urea biosensor was found to be stable for 3 weeks at entrapment. The different techniques for enzyme immobilization
25 ◦ C. Efforts were made to improve the applicability of a urea elec- are depicted in Fig. 3.
trode prepared in pioneering work by the same group in the late A potentiometric enzymatic biosensor, with a detection
sixties [19–23]. Many different versions of this biosensor have been range for urea between 8.9 × 10−5 and 1.1 × 10−3 M, consisting
described. The originally used glass electrode-based sensing ele- of urease immobilized on PVA activated with 2-flouro-1-
ment was soon replaced with a more selective, neutral carrier-type methylpyridiniumtoluene-4-sulfonate (FMP) was developed by
ion-selective ammonium electrode [24,25], with Severinghaus- Sehitogullari and Uslan [41]. This sensor has a detection limit
type ammonium [26] or carbon dioxide [22] gas measuring cells. of only 1 mM and does not cover the physiological range of
Miniaturized urea electrodes, several of them using ISFET transduc- urea. Lakard et al. used polyethylenimine films coated onto glass-
ers [27,28], have been constructed. Improvements on the enzyme sealed, metal microelectrodes for the potentiometric detection
layer have been achieved [29], resulting in higher enzyme stability of urea. They state the advantages of chemical immobilization
and lower limits of urea detection. The following sections contain methods over physical adsorption and entrapment as providing
details on various urea biosensors reported to date. mechanical strength, long-term stability and overcoming diffu-
sional limitations encountered by substrates and enzymes. Among
3. Urea biosensors based on polymeric matrices the various immobilization techniques used for the fabrication of
urease biosensing electrodes, glutaraldehyde immobilized urease
3.1. Urea biosensors based on non-conducting polymer matrices biosensors have been known to provide many advantages. These
include a short response time (15–30 s), sigmoidal response for urea
Polymers are finding wide usage in the field of electronic mea- concentrations in the working range from 1 × 10−2.5 to 1 × 10−1.5 M
suring devices, especially sensors owing to their ability to have and a lifetime of about 4 weeks. Also included are benefits such as
their chemical and physical properties tailored over a wide range low diffusional resistance, as well as strong binding forces between
of characteristics [30]. The suitably chosen polymeric matrices enzyme and matrix, thus reducing the loss of enzymes and increas-
are found to be biocompatible, flexible, and cost-effective. Besides ing stability.
this, they can be obtained in the form of free-standing films for The anion exchange method has been utilized for urease
the fabrication of biosensors. Extensive efforts have been made immobilization onto electroprecipitated poly(vinyl ferrocium) per-
to fabricate urea biosensors using different polymer matrices, chlorate electrodes to detect urea as low as 1 ␮M in the linear range
such as phosphine oxide polyether [31], polygalacturonate [32], from 1 to 250 ␮M with a shelf-life of 29 days. However, these matri-
polyurethane-acrylate [33], acrylonitrile [34], hydroxyapatite [35], ces suffer from toxicity, owing to polymer degradation products
alginate [36], bovine serum albumin [37], poly(vinyl ferrocenium) [38].
[38,39], polyethylenimine [40], synthetic latex polymer, activated Natural polymers (e.g., polysaccharides like chitin and cellulose)
polyvinyl alcohol (PVA) [41], polyvinyl chloride (PVC) [42–44], are biocompatible and have become the materials of choice for
and chitosan [45]. These materials provide mechanical strength recent technological advances in biosensor fabrication. Chitin and
G. Dhawan et al. / Biochemical Engineering Journal 44 (2009) 42–52 45

a deacetylated product, chitosan, have been used for the immo- poly(vinylchloride) film, has been developed. This urea biosen-
bilization of enzymes for biosensors and industrial applications sor has been used for flow injection urea measurements in
[47], as they are non-toxic and biocompatible [46]. Enzymatic sen- the 0.01–2 mM concentration range. These sensors suffer from
sors with chitin and chitosan supports have been reported for the the influence of interferants such as ammonium, potassium and
determination of glucose, lactate, ethanol and urea. Most research sodium ions. In this context, several strategies have been used
papers reported in the literature toward urease immobilization on to overcome the effect of interferences. For instance, differential
chitosan are dominated by the use of beads. The chitosan-based measurements including separation by gas diffusion and electronic
potentiometric urea biosensor is founded on the use of solid-state tongues have been used [60,61,64]. Ion-selective-electrode arrays
ammonium electrodes, which utilize four different immobiliza- and multivariate calibration models have proved to be effective for
tion procedures. These include: (A) adsorption; (B) adsorption increasing selectivity, enabling the simultaneous determination of
followed by reticulation with dilute aqueous glutaraldehyde solu- several species in different types of samples, ranging from biologi-
tions; (C) activation with glutaraldehyde followed by contact with cal fluids [65] to natural waters [66].
the enzyme solution; and (D) activation with glutaraldehyde, con- Despite these interesting developments, the development of
tact with the enzyme solution and reduction of the Schiff base simple assay methods is of importance in clinical studies. The
with sodium borohydride to obtain membranes with high enzyme miniaturization of biosensors is particularly important to reduce
loadings with adequate permeability towards ammonium ion and the amount of enzyme and reagents needed. More efforts must be
without significant enzyme leaching [45]. It has been reported that made to minimize the influence of interferants.
the biosensor fabricated using method (B) exhibits improved per-
formance with response characteristics including: (i) linearity in 3.2. Urea biosensors based on conducting polymer matrices
the range 10−4 to 10−2 M; (ii) a slope of up to 56 mV per decade;
(iii) response time between 30 s and 2 min; and (iv) a lifetime of 2 Conducting polymers have become the materials of choice for
months. However, more work must be done in this field to obtain recent technological advances in biotechnology and have been
a commercial urea biosensor based on natural polymers. extensively reviewed by various researchers [67–70]. These matri-
PVC-based ammonium ion-selective electrodes have been ces are used as supports for biomolecules, resulting in biosensors
extensively used for the fabrication of potentiometric urea biosen- that have enhanced speed, sensitivity and versatility in diag-
sors. These conventional potentiometric urea biosensors, pH nostics to measure desired analytes. These devices are finding
electrodes [48–50] and NH4 + -selective electrodes [51–53] detect ever-increasing use in clinical diagnostics. Conducting polymers are
changes in hydrogen ions or ammonium ions, respectively, pro- conjugated polymers that can be synthesized by chemical methods
duced in an enzymatic reaction. However, they suffer with the as well as electrochemical methods, providing easy modulation of
associated problem of strong dependence on sample buffer capac- various properties (e.g., film thickness, conductivity, functionaliza-
ity, which is known to be suppressed by the buffer used, leading tion, use of various supporting electrolytes and the ability of serving
to a narrow dynamic range and loss in sensor sensitivity [48]. as an electrochemical transducer itself). Additional merits are that
PVC membrane electrode-based biosensors [54–57], with suitable enzyme molecules can be entrapped during electro-polymerization
ionophores, offer many advantages that include short response in one step and also that the polymer film uniformly covers the
time, good reproducibility, simple design and longer lifetime, up surface of substrate electrodes of any shape or size [71,72].
to 12 months [58,59]. Various conducting polymers, like polyaniline (PANi), polypyr-
Functionalized PVCs have been investigated for use as electrode role and polythiophene, have been used for the fabrication of
membrane materials to facilitate efficient enzyme immobiliza- biosensors. Among them, polypyrrole is one of the most extensively
tion [50,60]. Also, suitable ionophores have been utilized with used conducting polymers in the fabrication of urea biosensors
PVC ion electrodes to enhance response characteristics, such as [73–75]. The versatility of this polymer is determined by (i) its
response time, reproducibility, simple design and longer lifetime. biocompatibility, (ii) capability to transduce energy arising from
The response characteristics [54,60,61] and potential use of ami- the interaction of analytes and analyte recognizing sites into
nated PVC membranes for urea biosensor fabrication has also been electrical signals that are easily monitored, (iii) capability to pro-
discussed. An ionophore-free PVC–NH2 pH membrane electrode tect electrodes from interfering material, and (iv) easy ways for
covalently immobilized with enzyme was applied by Walcerz et al. electro-chemical deposition on the surface of any type of electrode.
for urea estimation by measuring hydronium ions after enzymatic Among other conducting polymers, polyaniline is often used as the
hydrolysis [62]. Later, in 2002, Tinkilic et al. fabricated a minia- immobilizing substrate for biomolecules and as an efficient electro-
turized urea sensor by immobilizing urease directly onto an all catalyst. However, the necessity to detect bioanalytes in neutral pH
solid-state, contact PVC–NH2 membrane, ammonium ion-selective ranges leads to electro-inactivity of the deposited film, discouraging
electrode [44] with improved sensitivity, longer lifetime and low the use of polyaniline and polythiophene as biosensing materials
cost. The biosensors showed a linear range between 5 × 10−2 for the detection of urea.
and 5 × 10−4 M urea in unbuffered solution. However, it has been Adeloju et al. have fabricated amperometric, flow injection, urea
reported that sensitivity is highly dependent on the H+ , Na+ and K+ biosensors using electrochemically entrapped urease in polypyr-
concentrations in the sample solution, as severe interference from role [76–78]. The influence of flow rate and various other analytical
Na+ and K+ ions is observed. Efforts were made to reduce the effect parameters (e.g., applied potential and polymerization time) have
of interferants to achieve improved performance over carboxylated been studied. Under optimized conditions, the amperometric
PVC-based biosensors. This was in terms of sensitivity, dynamic response was linear between 3 and 15 mg/l of urea. The sensi-
stability over 2 months and a response time of 1–2 min. Also, tivity and detection range were further improved by increasing
this concerned the estimation of serum urea concentration using the enzyme loading and utilizing pulsed amperometric detec-
urease-biosensing electrodes that were based on PVC-membrane tion [79]. Using pulsed amperometric flow injection sensors based
ammonium-selective electrodes containing palmitic acid and non- on polypyrrole, detection limits of 60 mg/l and a linearity in the
actin [43]. 100–450 mg/l range has been observed. However, a shelf-life of only
An optical biosensor [63] based on an ion-selective optode mem- 2 weeks could be achieved using this technique. This sensor suf-
brane, which contained nonactin as the ion-selective ionophore fers from a limitation of quantitative urea estimation in biological
and ETH 5294 chromoionophore in a thin (1 mm) plasticized samples, as it needs pretreatment with anion exchange separators
46 G. Dhawan et al. / Biochemical Engineering Journal 44 (2009) 42–52

for the removal of interferants and needs further improvement to Polyaniline–Nafion-based composite films have been immobi-
minimize the interferant effect [79]. lized with urease via electrochemical and casting methods for
Efforts have been made to improve the sensitivity of the the fabrication of amperometric urea biosensors [75]. Urease
polypyrrole-based urea sensor by using composite films with immobilization on these composite films has been achieved by
polyion complexes. Komaba et al. have reported a potentiometric electrochemical immobilization and the casting method. It was
urea sensor based on nucleophilic electrolytes and urease, indi- found that the sensitivity of composite electrodes immobilized
cating improved urea response with a slope of 31.8 mV per decade with the casting method is greater than that of the electrochemical
[74]. However, it has been found that only a small number of urease immobilization method. The sensitivity and detecting limit of this
molecules could be immobilized using the electro-polymerization urea sensor were found to be 0.7 ␮A (mg dl−1 )−1 cm2 and 6 mg dl−1 ,
process. To overcome this problem, the polyanion complexes, respectively, when urease was immobilized by a glutaraldehyde
polyacrylic acids, have been used to enhance the degree of immobi- (GA) cross-linker; and 5.27 (mg dl−1 )−1 cm2 and 0.3 mg/dl, when
lization and the response was found to increase to 53 mV per decade immobilized by a Nafion network [85]. However, no studies have
[80]. It has been demonstrated that the sensitivity can be further been carried out for the stability of urease and the estimation of
increased by using polyion complexes (i.e., polycation polystyrene urea in clinical samples.
sulphonate and polyanion polyacrylic acid) and, using a polyion For practical applications of urea biosensors, problems like
complex, the sensitivity was found to be enhanced to 110 mV per reproducibility, long-term stability and clinical sample measure-
decade. This high sensitivity has been attributed to effective immo- ments (e.g., in urine and serum) remain the major thrust. Although,
bilization of the large amount of urease by electro-polymerization conducting polymer-based urea sensors have a lot of advantages
of the electrode with a precoated polyion complex [81]. However, (e.g., stability, tunable properties, flexibility and biocompatibility)
the experiments on shelf-life and sensor reproducibility have not they cannot work properly in high ionic strength samples. Fur-
been reported. thermore, the response of these sensors is highly dependent on
To minimize enzyme desorption from the desired immo- the buffer capacity of the sample, resulting in a narrow dynamic
bilization material and to obtain increased enzyme electrode range and low sensitivity. It may be noted that the use of additional
stability, various approaches and immobilization methods have perm-selective membranes on top of the enzymatic membrane
been used. A urea biosensor achieving the electrochemical entrap- could substantially reduce the dependence of sensor response on
ment of urease in a polypyrrole matrix, including a stability of buffer concentration and significantly extend its dynamic range.
around 2 months, has been reported [78]. This biosensor, how- For example, the use of Nafion and polyvinylene phenylene mem-
ever, cannot be reused. A functionalized copolymer of pyrrole and brane on top of enzymatic layers has been reported to substantially
N-3-aminopropylpyrrole [75] has been used for the fabrication improve the dynamic range of urea biosensors and reduce the
of thin film urea biosensors. In this system, urease was cova- dependence of buffer concentrations [86].
lently immobilized onto the copolymer matrix. The sensor acted Current developments in conducting polymer-based biosensors
via covalent binding with free –NH2 groups of the polymer. This show that electrochemical affinity sensors, based on molecularly
sensor exhibited improved characteristics, such as 2-month sta- imprinted conducting polymers, have a great potential for direct
bility at 4–6 ◦ C and urea detection in the range of 6.3 × 10−6 to electrochemical sensing. This is because they enable deliberate con-
4.07 × 10−4 M, with electrode sensitivity at 27.5 mV/dl. Sensor sta- trol of the molecular structure at the electrode surface, resulting
bility of this system has been attributed to a porous morphology, in the additional advantage of higher affinity to the target ana-
owing to the large, inserted PTS dopant. This results in high enzyme lyte. Therefore, nanostructured conducting polymers and tailored,
loading and covalent binding with copolymers having free –NH2 molecularly imprinted conducting polymers are likely to be the
groups. best choice to obtain enhanced stability and reproducibility in urea
A polypyrrole-based urea biosensor has been reported that pro- biosensors.
vides advantages via eliminating the need for an optical indicator,
dye and pH sensitive reagent. Polypyrrole itself acts as a pH sensitive 3.3. Sol–gel-based urea biosensors
indicator in the near-IR range, with the change in urea concentra-
tion at the linear range of 0.06–1 M urea [73]. Sol–gel matrices have been known for their rigidness, chemical
Polyaniline, among other conducting polymers, is known to inertness, thermal and photochemical stability, negligible swelling
have a broad range of tunable properties, desirable electrochem- in aqueous solution, tunable porosity and optical transparency.
ical activity, chemical stability, structural flexibility and, above all, Besides this, they have been widely used in the fabrication of chem-
two redox couples in a convenient potential range. A pH urea sen- ical opto-electronic sensors, as most of the biological materials tend
sor has been reported in both aqueous and non-aqueous media via to retain their activity owing to the attractive low temperature pro-
electro-polymerization of polyaniline in dry 0.5 M acetonitrile. This cess of immobilization for various biomolecules (e.g., enzymes and
sensor retains its reproducibility for more than 9 months. However, antibodies).
the sensor has not yet been used for clinical samples [82]. A conductometric urea biosensor, based on tetramethyl
A copolymer of polyaniline-poly(n-butyl-methacrylate) (Pn- orthosilicate sol–gel immobilized urease, on a screen-printed inter-
PBMA) homogeneous composite films has been obtained using digitated array (IDA) electrode, has been used to detect urea
poly(vinyl methyl ether) (PVME) and poly(vinyl ethyl ether) in linear dynamic range 0.03–2.5 mM with good reproducibly
(PVEE) as dispersants. This copolymer has been character- using the conductometric screen printed technique [87]. This urea
ized for its morphological, electrical and mechanical properties biosensor exhibits good sensor-to-sensor reproducibility, but has a
[83]. These composites have been used for the detection of poor storage stability of around 25 days. The stability of the sol–gel-
H2 O2 , NH3 and toward the fabrication of urea and uric acid encapsulated urease biosensor decreased over time, which was
biosensors. Polyaniline–Nafion composites have been used to attributed to denaturation during immobilization or restriction of
develop an amperometric urea biosensor that is based on the enzyme in an inactive conformation after encapsulation.
polyaniline–perfluorosulfonated ionomer composite electrodes. Cu(II) and Cd(II) array-based optical urease biosensors, which
This system uses a flow injection analysis system, has been fab- were found on tetra ethoxy orthosilicate, have been fabricated
ricated and is reported to have a detection limit of 0.5 ␮M as well with a FITC–dextran encapsulated fluorescent dye. Attempts have
as a response time of 40 s [84]. been made for its application to biological and environmen-
G. Dhawan et al. / Biochemical Engineering Journal 44 (2009) 42–52 47

tal samples [88]. Similarly, urease, acetyl choline esterase and were carried out using an ammonium ion analyzer [92]. The detec-
FITC–dextran co-entrapped TEOS sol–gel matrix have been used to tion limit and sensitivity of these electrodes was found to be 5 mM
detect heavy metals and urea (2.5–50 M), with an analytical range and 10 mV/mM, respectively. The shelf-life of these electrodes was
of 2–3 orders-of-magnitude. The applicability of this developed- found to be 5 weeks at 4 ◦ C. However, the detection limit does not
array biosensor toward simultaneous multi-analyte determination cover the physiological range of urea.
has been demonstrated. Although results provide a method for It may be noted that most studies discussed above were not
simultaneously analyzing multiple samples and multi-analytes in carried out using clinical samples and the influence of interferants
environmental and biological samples, the urease-based system (e.g., potassium and sodium ions) has also not been reported. As
exhibited relatively poor stability and its initial sensitivity dropped these materials show promise for applications toward the design
to 45% after 4 weeks [89]. and development of biosensors, more emphasis should be given to
The performance characteristics of urea biosensors based on explore these advantageous materials for improved characteristics
urease immobilization using physical adsorption, a physically of urea biosensors.
entrapped sandwich and microencapsulation in a tetramethy-
lorthosilicate (TMOS)-derived sol–gel matrix on the surface of 5. Urea biosensors based on nanomaterials
a glass–pH electrode, have been studied [90]. Various parame-
ters for biosensor performance (e.g., detection limit, sensitivity, Nanostructured materials have found a great variety of appli-
linear range, response time and regression) have been studied cations owing to characteristic properties, such as (i) in catalysis
using potentiometric techniques. The results indicate that the reactions, due to their high surface area and controlled crystal
microencapsulation technique is a better method for enzyme surfaces, (ii) in sun screen, hyperthermic cancer treatment and flu-
immobilization in sol–gel films derived from TMOS. The advan- orescent tags based on optical properties, and (iii) in smoke/fog
tage of microencapsulated biosensors over physically entrapped screens because of their light scattering effect. Nanomaterials
sandwich configuration biosensors includes: higher sensitivity have found applications in drug delivery (inhalation asthma and
(2.4 (dpH/dp(C))), lower detection limit (4.5 ␮M) and a larger lin- timed drug release), pesticide delivery (fogging and fumigation),
ear range of urea determination (0.01–30 mM). Microencapsulated magnetic recording (orient magnetic domain axis, hard drives,
enzyme electrode systems yield more satisfactory results with video and audio tapes), pigments, inks and paints. A number of
blood serum samples. However, microencapsulated enzyme elec- biosensors based on nanomaterials, such as nanoparticles [96],
trode systems have shown interference for various analytes (e.g., nanofibers, conducting polymers, nanofilms/nanocomposites [97],
glucose, FeSO4 , thiourea and HgCl) and suffer from a poor storage or self-assembled monolayers [98,99], have recently been reported.
stability of 2 weeks and a longer response time of 5 min. Sol–gel Biomolecules (e.g., enzymes, antigens/antibodies or DNA)
matrices based on molybdenum trioxide, for the development of exhibit nanoscale dimensions comparable to the dimensions of
urea biosensors, have been studied. It has been found that ure- metal or semiconductor nanoparticles. These size similarities pave
ase retains its activity inside molybdenum trioxide sol–gels that the way to combine the unique electronic, photonic, and catalytic
are sensitive to ammonia. These hybrid nanoporous composites are properties of nanoparticles with the nature-evolved, selective bind-
reported to be useful for biosensors and fuel cells [91]. ing and catalytic functions of biomolecules. This can be achieved
Apart from the entrapment of sensing agents without enzyme by integrating native and synthetic components into hybrid sys-
leaching, longer biomolecule stability, the advantages of sol–gel tems. Such hybrids may reveal new material functions, wherein
glasses include short response time and easy fabrication of multi- the nanoparticles would introduce new electronic, photonic or
analyte detection electrodes. Drawbacks, including diffusional opto-electronic properties into biomolecules, or alternatively, the
limitations, reproducibility of results, sensitivity, non-conductive biomolecules would affect the structure or properties of nanopar-
nature, unknown catalyst–matrix interactions and kinetics, chal- ticles. Extensive research efforts have been directed in the past
lenge researchers. Attempts and approaches should be directed few years towards the development of biomolecule–nanoparticle
toward solving these problems. hybrid assemblies and their application to construct biosensors,
nanoscale circuitry or nanodevices.
4. Urea biosensors based on Langmuir–Blodgett films
5.1. Nanoparticles
Langmuir–Blodgett technology has been considered a conve-
nient tool for designing artificial systems with biological functions Gold nanoparticles have been studied for biosensor fabrication,
(e.g., biosensors), where ultra-thin films containing biomolecules as biological molecules can retain their activity when adsorbed onto
are expected to preserve high enzymatic activity. Until now, many them [100]. Gold nanoparticles can immobilize enzymes based
kinds of biosensors (e.g., cholesterol, glucose and urea) have been on chemical adsorption onto self-assembled monolayers [101,102].
designed with LB films as the biosensitive part. However, very little Yang et al. have proposed a renewable inhibitive mercury sensor
work has been done in the area of urea biosensors based on LB films based on self-assembled gold nanoparticles on an ethylenediamine
[92,93]. poly(vinyl chloride) (PVC–NH2 ) membrane-based pH electrode
Urea biosensors based on octadecylamine and enzyme urease [103]. Gold nanoparticles have been assembled on the surface
on ISFET surfaces have been used to detect urea at levels as low as of a PVC–NH2 membrane. Urease was immobilized on the gold
0.2 mM, in the dynamic range of 0.2–20 mM, within 15 s. Monolayer nanoparticles through chemical adsorption. An advantage of using
stabilization has been achieved using glutaraldehyde [94]. assembled gold nanoparticles lies in the fact that the inactive
Highly stable urease monolayers have been immobilized onto enzyme layers denatured by Hg2+ can be rinsed out via a saline
hydrophobic surfaces of octadecyl silane-modified SiO2 /Si sub- solution using acid and alkali washes successively.
strates using the Langmuir–Blodgett deposition technique. The sen- Gouma et al. [104] have discussed the production of biological
sitivity of urea detection in this enzyme/insulator/semiconductor sensors using nanostructured metal oxides. This work was focused
(EnIS) system was found to be 22 mV/pUrea [95]. on the synthesis of novel structures of nanocomposite metal
Urease-immobilized, mixed monolayers of poly(N-vinyl car- oxide systems by combining sol–gel processing and electrospinning
bazole) (PNVK) and stearic acid (SA) LB monolayers have been with an application towards biosensing probes. Encapsulation and
studied. Potentiometric measurements on these urease electrodes incorporation of these biomolecules into highly surface-porous,
48 G. Dhawan et al. / Biochemical Engineering Journal 44 (2009) 42–52

biocompatible polymer mats enables their higher catalytic activity, sputtering and Ag evaporation. The working electrode was com-
faster response and possibly longer operation lifetime. Urease was prised of urease immobilized onto an organic polymeric conductor
chosen as the model enzyme system in these studies, as it acts as a (PPy). The electrochemical properties were characterized using
catalyst in the hydrolysis of urea to ammonia and carbon dioxide. A both amperometry and potentiometry for detection of urea. Yun
novel approach to biosensor development using hybrid nanomate- et al. [107] used a similar technique for the development of an
rials systems for the selective detection of biospecies (pathogens) of amperometric urea biosensor. In their work, the attachment of
interest has been demonstrated, such as electrospun polymer/oxide urease to a carboxylic-terminated SAM, which was chemisorbed
nanofiber structures in which proteins or cells are incorporated onto a gold electrode, was achieved using carbodiimide coupling.
(Fig. 2). The result of biochemical reactions between the react- The formation of a self-assembled, chemisorbed layer of ure-
ing biospecies and bioreceptors is a chemical signal that can be ase on gold was studied by X-ray photoelectron spectroscopy
sensed by the hybrid nanofiber platform through a conductimetric (XPS). The measured sensitivity of the urease/SAM/Au/planar sil-
mode. These new findings open the pathway for numerous appli- icon electrode was found to be 11.21 ␮A/(mM cm2 ) and that of the
cations of bio-nanocomposite materials, including non-invasive urease/SAM/Au/PlanarSi electrode was ca. 4.39 ␮A/(mM cm2 ). An
medical diagnostics tools, bio-fuel cell devices, protective cloth- approximately threefold sensitivity increase was observed in the
ing (electronic and interactive textiles), bioengineering scaffolds porous-silicon-based electrode.
and advanced sensor arrays. Further work must be performed to Nanopatterned porous alumina has been prepared by electri-
exploit these findings toward designing biosensors. cal anodization in acid solution and has been used as an enzyme
electrode and pH sensor for urea detection. The biocompatibil-
5.2. Quantum dots (QDs) ity of nano-size porous alumina as an immobilization matrix for
biocolloidal systems has stimulated interest for improved sen-
Quantum dots or semi-conducting nanoparticles are highly sor sensitivity. The sensor was tailored to enlarge the active
luminescent, photostable fluorophores that have recently been surface area, which in turn increased the sensitivity of the
used for sensing applications. QDs have much higher photolumi- sensor.
nescence quantum efficiency than their bulk counterparts. Hybrid
systems containing QDs coupled with various biomolecules have 6. Miscellaneous matrices for urea biosensing
stimulated research in biotechnology and nanotechnology. A sub-
tle change of the surface property of QDs can result in a dramatic Some other supports used for urea biosensor fabrication
change in their optical properties. In principle, this novel feature of include: bilayer lipid membranes [108], carbon paste electrodes
QDs can be extended for detecting specific analytes if appropriate [109], composite hydrogel membranes [110], gelatin membranes
conditions are established. An assay system with urease as catalyst [111], inorganic matrices (e.g., clays, laponite and Zn–Al layered
and CdSe/ZnS quantum dots as indicators has been developed for double hydroxides matrices) [112,113], SiO2 films, glass beads,
the quantitative analysis of urea by Huang et al. [105]. Detection diatomite, silica, porous glass, hornblende and molecular sieves
in this system is based on the enhancement of QD photolumines- [114–118].
cence intensity, which is correlated to the enzymatic degradation of Srivastava et al. [111] have described a urea biosensor utilizing
urea. By controlling buffer concentration and pH, PL enhancement urease immobilized on gelatin beads. The system was success-
due to urea degradation was found to be linear in the 0.01–100 mM ful in achieving a long storage stability, with a half-life of 240
concentration range. days. The characteristics of this sensor include a detection limit
Important advances using semiconductor QDs for biosensing from 0.8 to 23 mM and a response time of 6 min. Another sensing
processes have been reported in recent years. Nevertheless, this system has been proposed for urease immobilization on porous
field requires further development, and challenging goals still glass beads [119]. The sensor set-up proposed by Seki’s group used
require scientific answers. The development of chemical proce- urease immobilized on SiO2 films as the pH sensitive layer in light-
dures for the preparation of highly fluorescent QDs in aqueous addressable potentiometric sensors (LAPS) [120], with a detection
media, while retaining the photonic communication between the range from 5 to 15 mM. However, more studies must be done in this
QDs and the associated dye units, is essential. field to fabricate a cost-effective, commercial sensor using these
novel materials.
5.3. Nanofibers The use of enzyme mixtures can provide improved performance
in urea sensors. Seo et al. [122] determined NH4 + via amperometric
With increasing demands for nanotechnology, electro-spinning methods, using the immobilization of l-glutamate dehydrogenase
has become a novel technique for generating composite nanofibers on the immobilon-AV affinity membrane. Incorporation of the
[106]. Sawickaa et al. [97] have utilized the electro-spinning enzyme into a carbon paste was performed by Yang et al. [109]
technique to prepare nanocomposite fibers of urease and for the construction of an amperometric enzyme electrode. The
polyvinylpyrrolidone (PVP). The non-woven mat was found to have enzyme-modified carbon paste electrode resulted in improved
a potential use as a urea biosensor, showing characteristic features specificity and sensitivity. This system made use of bi-enzyme
of fast response time, sensitivity to lower concentrations of urea, systems involving glutamate dehydrogenase and urease for urea
and a more versatile design. The immobilized enzyme was found estimation. Response characteristics of the system were based on
to remain active inside the polymer solution and reactivity was the redox activity of NADH. The oxidation current of NADH was
maintained inside electrospun non-woven mats. The electrospun monitored at +1.10 volt vs. Ag/AgCl. The optimum pH range for cat-
membrane acts as a catalyst for hydrolysis in 0.5–2.5 mM urea solu- alyzed hydrolysis reactions of urea was found to be between 7.0 and
tions. 7.4. The linear response range and detection limit were found to be
2.0 × 10−5 to 2.0 × 10−4 M and 5.0 × 10−6 M, respectively. Liu et al.
5.4. Nanoporous silicon technology [123] have shown that the immobilization of enzymes on ␥-Al2 O3
is well-suited for the preparation of biosensors. Interestingly, the
Nanoporous silicon technology has been used for the fabrica- enzymes can be entrapped in internal surfaces by physical or chem-
tion of urea biosensors. Jin et al. [106] fabricated three thin film ical reaction (e.g., change in temperature, pH, or pretreatment of the
electrodes patterned on p-type silicon wafers using platinum RF substrate) without using cross-linking chemicals that can partially
G. Dhawan et al. / Biochemical Engineering Journal 44 (2009) 42–52 49

deactivate enzyme activity. A potentiometric biosensor for urea 7. ISFET-based urea biosensors
was found to be sensitive in the concentration range of 3.0 × 10−5
to 1.4 × 10−2 M, with a detection limit of 10 ␮M. This novel urea In 1980, Caras and Janata described the first pH-ion-selective,
sensor has been applied toward the measurement of urea in field effect transistor (pH-ISFET)-based, enzyme-modified biosen-
urine. sor sensitive to penicillin [126]. ISFET-based biosensors have many
Kirstein et al. [121] described a urea biosensor in which amper- advantages, such as miniaturization, high sensitivity, high level of
ometric pH sensing was realized by pH dependence of hydrazine activity, low cost and multi-detection potential, which is especially
electrochemical oxidation in the Tafel region. The sensor showed important for the fabrication of multi-biosensors. As the enzy-
a linear range from 0.8 to 35 mM. Disadvantages lie in its low matic hydrolysis of urea causes an increase in pH of the system,
sensitivity and use of the hazardous compound hydrazine. Piz- ISFETs have been widely used for this purpose. Various methods
zariello et al. [124] have made use of a pH-sensitive redox-active for urease immobilization on the surface of ISFET transducers have
mediator, Hematin, for the development of an amperometric urea been reported, such as inclusion in a gel [127,87], confinement
biosensor. Urea biosensors are constructed with urease immobi- [128], reticulation with a bifunctional agent [129], covalent bond-
lized on the surface of platinum and graphite composite electrodes ing on an active support [130,131], and Langmuir–Blodgett films
and amperometric measurements are done using 0.5 mM hematin. [132]. Bovine serum albumin (BSA) has been widely used for the
The amperometric sensor reported by Yoneyama was combined immobilization of urease on a transducer surface via chemical
with a pH-stat method and flow injection analysis. This system cross-linking [133,134]. Apart from all these methods, silaniza-
recorded a linear relation between current and urea concentration tion is most commonly used [135–143]. This procedure enables
within a narrow range of 0–600 ␮M urea solutions [119]. Stedansky the introduction of chemically reactive groups. In the case of
[125] tried to exploit the redox activity of certain mediators (e.g., silanization performed with aminosilanes, reactive groups are of
hematein, lauryl gallate, and adsorbed methylene blue) for the con- the amine type (NH2 ). These groups are used to form covalent
struction of cheap and reliable urea biosensors. Under optimum bonds between enzyme amino groups and glutaraldehyde. Sev-
conditions, the sensors showed a detection limit of 2 ␮M, linear- eral authors have used this technique for the fabrication of urea
ity from 0.002 to 0.75 mM, dynamic range of 4 mM and sensitivity biosensors [144–149].
of 15.2 Ma/Mcm. However, attention should be paid to the selec- The various detection techniques and matrices for the fabrica-
tion of working conditions, such as initial pH, operating potential, tion of urea biosensors are listed in Table 1.
buffer type and capacity. The authors have concluded that further
improvements could be expected with a more systematic search 8. Demands, challenges and commercialization
for pH-sensitive, redox probe molecules.
It is suggested that a large number of pH sensitive compounds According to the National Kidney Foundation, USA, more than
(e.g., quinines, polyphenols, antioxidants, and organic dyes) are 20 million Americans (one in nine adults) have chronic kidney dis-
redox active and could serve as probe molecules for monitoring ease. More than 20 million more are at increased risk for developing
urea. Conversely, from a practical point of view, a suitable pH- kidney disease and most do not even know it [150]. The major cause
sensitive redox compound should have physical stability and work for kidney failure has been attributed to diabetes. In 2003, diabetes
at a potential near 0 mV, to avoid electrochemical interference from accounted for about 44% of new cases and about 36% of all kid-
real clinical samples. ney failure cases in the US [151]. Uncontrolled or poorly controlled
Therefore, the use of mixed enzyme systems with cost- high blood pressure is the second leading cause of kidney failure in
effective matrices may be suitable for the development of urea the US, followed by glomerulonephritis, an inflammatory disease
biosensors. of the kidneys, and polycystic kidney disease. Kidney-related dis-

Table 1

Matrix Transducer Response time Sensitivity Detection limit Stability Ref.

Poly(vinylferrocium) Amperometry 60 s 1 ␮M 29 days [29]


Polyethylenimine Amperometry 15–30 s 4 weeks [30]
Polyvinyl alcohol Potentiometry 2–4 min 10−4 M 1 month [32]
Polyvinyl chloride–palmitic acid Potentiometry 1–2 min 54.3 ± 0.4 mV/Purea 60 days [34]
PVC–NH2 Potentiometry < 10s 48 ± 5 mV/dl 3 × 10−5 M 1 month [35]
Chitosan Potentiometry 30 s–2 min 56 mV/dl 2 month [42]
PVC–nonactin Optical 16–20 s 0.05 absorption unit in 0.08 mM [63]
0.01–2 mM urea
Polypyrrole Potentiometry 31.8 mV/dl [74]
Polypyrrole Amperometry 60 ug/l 2 weeks [75]
Poly pyrrole Amperometry 3 ppm [77]
Poly(N-3-aminopropylpyrrole-co-pyrrole) film Potentiometry 25–50 s 27.5 mV/dl 2 months [79]
Tetraphenyl doped polyaniline Potentiometry 86 mV/pH 20 ␮M 2 months [80]
Tetramethyl ortho silicate Conductometric 204 ␮S/mM 30 ␮M 25 days [86]
TMOS derived sol–gel silicate matrix Glass–pH electrode 5 min dpH/dp(C) = 2.4 52 ␮g/ml; 25 days at 4 ◦ C [89]
Poly(N-vinylcarbazole) and stearic acid Ammonium 10 mV/mM 5 mM 5 weeks at 4 ◦ C [91]
ion-selective electrode
(potentiometry)
Urease onto Si/SiO2 Struct. CV measurements 22 mV/Purea 1 mM Few days [93]
Urease/PVP nanocomposite nanofibers Quantum dots [95]
Optical 0.01 mM [102]
Carbon paste electrode Amperometric 5.0 × 10−6 M 15 days [106]
Zn3 Al–urease layered double hydroxides Impedometric 21 mV/Purea [110]
␥-Al2 O3 Potentiometric 10 ␮M [120]
50 G. Dhawan et al. / Biochemical Engineering Journal 44 (2009) 42–52

orders rank third amongst all life threatening diseases in India as nanostructure synthesis, has great promise in future nanotechnolo-
well, after cancer and cardiac ailments [152]. gies.
There is increasing demand for a more reliable, accurate urea
monitoring system. Urea biosensors face many problems in differ- Acknowledgements
ent terms. For very low concentrations, there is always a chance of
losing ammonia, as it may not be obtained in the dissolved form and We are grateful to Dr. Vikram Kumar, Director, National Phys-
may be liberated into the atmosphere yielding results with errors ical Laboratory, New Delhi, India for his interest in this work.
in terms of reduced urea levels. Another important problem that Gunjan Dhawan is thankful to the Council of Scientific and
needs to be rectified while fabricating urea biosensors, especially Industrial Research (CSIR) and University Grants Commission
potentiometric biosensors, is the effect of interferants. Sodium and (UGC), India, for a Junior Research Fellowship. We thank Dr.
potassium ions present in blood, serum or urine samples interfere Kavita Aora for interesting discussions during the preparation
in urea analyses and hamper accurate monitoring. Hence, there is of this manuscript. We sincerely acknowledge financial sup-
an urgent need for more and better analyzers that can serve the port from DST-sponsored projects: DST/CLP041332 CSIR funded
desired purpose of monitoring urea in clinical samples, be cost- CMM-11, DBT/PR7667/MEP/14/1057/2006, DST/INT/JAP/P-21/17,
effective and also be free of any disadvantages usually occurring in DST/TST/TE/2007/06 and DST/SEN/SI/03.
the present day biosensors.
A commercially available urea biosensor is the i-STAT Portable
References
Clinical Analyzer, which analyzes different electrolytes, including
urea, nitrogen, glucose, sodium ions, potassium ions, chloride ions [1] A.J. Taylor, P. Vadgama, Ann. Clin. Biochem. 29 (1992) 245–264.
and hematocrit [153]. Recently, Christoph Ritter [154] introduced [2] L. Della Ciana, G. Caputo, Clin. Chem. 42 (1996) 1079–1085.
two commercial products from Roche Diagnostics, in which one [3] J. Carlsson, J. Bergström, Acta Vet. Scand. 35 (1994) (1994) 67–77.
[4] J.L.F.C. Lima, C. Delerue-Matos, M.C.V.F. Vaz, J. Agric. Food Chem. 46 (1998)
sensor was capable of detecting urea, glucose, and lactate, while the 1386–1389.
other was used for urea and creatinine determinations. Although [5] D. Lefier, Bull. Int. Dairy Fed. 315 (1996) 35–38.
these biosensors were compact and commercially available, they [6] N.M. Price, P.J. Harrison, Mar. Biol. 94 (1987) 307–319.
[7] L. Goeyens, N. Kindermans, M.A. Yusuf, M. Elskens, Estuarine Coastal Shelf Sci.
suffered the drawbacks of high cost, complicated construction,
47 (1998) 415–418.
and requirement of extra electrodes to compensate for interfer- [8] C.J. Patton, S.R. Crouch, Anal. Chem. 49 (1977) 464.
ences. Therefore, the development of cheap and disposable array [9] A. Ramsing, J. Ruzicka, E.H. Hensen, Anal. Chim. Acta 114 (1980) 165.
[10] F. Roch-Ramel, Anal. Biochem. 21 (1967) 372.
biosensors for the simultaneous detection of clinically important
[11] D. Bansi, Malhotra, S.P. Asha Chaubey, Singh, Anal. Chim. Acta 578 (2006)
metabolites and rapid screening of renal-related diseases is still 59–74.
needed. [12] K.Y. Park, S.B. Choi, M. Lee, B.K. Sohn, S.Y. Choi, Sens. Actuators B 83 (2002) 90.
The search for new advanced materials is an important area [13] P. Caillat, D. David, M. Belleville, F. Clerc, C. Massit, F. Revol-Cavalier, P. Peltié,
T. Livache, G. Bidan, A. Roget, E. Crapez, Sens. Actuators B 61 (1999) 154–162.
of contemporary research in numerous disciplines of science and [14] P-.G. Su, C.-S. Wang, Sens. Actuators B 124 (2007) 303–308.
the development of many new technologies. Great attention has [15] R. Wang, D. Zhang, W. Sun, Z. Han, C. Liu, J. Mol. Struct. THEOCHEM 806 (2007)
been paid in recent years to nanostructured materials of differ- 93–97.
[16] L.C. Clark, C. Lyons, Ann. N.Y. Acad. Sci. 102 (1962) 29.
ent chemical composition, produced as nanoparticles, nanowires [17] G.G. Guilbault, J. Montalvo, A urea specific enzyme electrode, J. Am. Chem.
and nanotubes. Electrochemical nanobiosensors offer without Soc. 91 (1969) 2164.
doubt an important step toward the development of selective, [18] G.G. Guilbault, J. Montalvo, Anal. Lett. 2 (1969) 283.
[19] G.G. Guilbault, E. Hrabankova, Anal. Chem. 42 (1970) 1779.
sensitive biorecognition devices with diagnostic applications. Elec- [20] G.G. Guilbault, J.G. Montalvo Jr., J. Am. Chem. Soc. 92 (1970) 2533.
trochemical nanobiosensors, consisting of single carbon nanotubes, [21] G.G. Guilbault, G. Nagy, Improved urea electrode, Anal. Chem. 45 (1973) 417.
quantum dots and nanorods, are another future path of biosen- [22] G.G. Guilbault, F.R. Shu, Enzyme electrodes based on the use of carbon dioxide
sensor, Anal. Chem. 44 (1972) 2161.
sor development. It is expected that such devices will develop
[23] G.G. Guilbault, M. Mascini, Anal. Chem. 49 (1977) 795.
reliable point-of-care diagnostics, biochip fabrication, and in vivo [24] C. Eggenstein, M. Borchardt, C. Diekmann, B. Gründig, C. Dumschat, K. Cam-
sensing, as well as multi-analyte sensors with increased cost- mann, M. Knoll, F. Spener, Biosens. Bioelectron. 14 (1999) 33.
[25] I. Walcerz, S. Glab, R. Koncki, Anal. Chim. Acta 369 (1998) 129.
effectiveness.
[26] T. Anfält, A. Graneli, D. Jagner, Anal. Lett. 6 (1973) 969–976.
[27] J. Muñoz, C. Jimenez, A. Bratov, J. Bartrolí, S. Alegret, C. Dominguez, Biosens.
Bioelectron. 12 (1997) 577.
9. Conclusions [28] A. Senillou, N. Jaffrezic-Renault, S. Cosnier, Talanta 50 (1999) 219.
[29] M. Mascini, Potentiometry: enzyme electrodes, in: A. Townsheand (Ed.), Ency-
clopedia of Analytical Science, Academic, London, 1995, p. 4112.
Attempts have been made to bring out the salient features of [30] B. Adhikari, S. Majumdar, Prog. Polym. Sci. 29 (2004) 699.
various urea biosensors as reported in the literature to date. Urea [31] S. Kwabena, P.H. Helen, H. William, Enzyme Microb. Technol. 17 (1995) 804.
[32] S. Ciurli, C. Marzadori, S. Benini, S. delana, C. Gessa, Soil Biol. Biochem. 28
sensors have the potential to be a billion-dollar market and the (1996) 81l–817.
technology needs improvement in biological stability, signal trans- [33] C. Puig-Lleixà, C. Jiménez, J. Alonso, J. Bartrolí, Anal. Chem. Acta 389 (1999)
duction, precision and cost effectiveness. The role of matrices for 179–188.
[34] T. Godjevargova, A. Dimov, J. Membr. Sci. 135 (1997) 93–98.
biosensing and the characteristics of various biosensors, in terms of [35] C. Marzadori, S. Miletti, C. Gessa, S. Ciurli, Soil Biol. Biochem. 30 (1998)
detection limit, response time, and linear range, have been delin- 1485–1490.
eated and it is suggested that microfabrication technology and [36] A.M. Kayastha, N. Das, Biochem. Educ. 27 (1999) 114–117.
[37] A.P. Soldatkin, V. Volotovsky, A.V. El’skaya, N. Jaffrezic-Renault, C. Martelet,
nano-engineering does indeed have an enormous and profitable
Anal. Chim. Acta 403 (2000) 25–29.
impact on the urea biosensor market. The development of cheap [38] F. Kuralay, H. Özyörük, A. Yýldýz, Sens. Actuators B 109 (2005) 194–199.
and disposable array biosensors for the simultaneous detection [39] F. Kuralay, H. Özyörük, A. Yýldýz, Sens. Actuators B 114 (2006) 500–506.
[40] B. Lakard, G. Herlem, S. Lakard, A. Antoniou, B. Fahys, Biosens. Bioelectron. 19
of clinically important metabolites and rapid screening of renal-
(2004) 1641–1647.
related diseases is still needed. The use of biomolecules to grow [41] A. Sehitogullari, A.H. Uslan, Talanta 57 (2002) 1039–1044.
NPs and nanowires has great promise in the future of biosens- [42] M. Gutíerrez, S. Alegret, M.D. Valle, Biosens. Bioelectron. 22 (2007) 2171–2178.
ing and design of bioelectronic systems. We believe that the use [43] K.S. Pekyardýmcý, E. Kýlýç, Artif. Cells Blood Substit. Biotechnol. 33 (2005)
329–341.
of biocatalytic nanostructure growth, using dip-pen nanolithog- [44] N. Tinkilic, O. Cubuk, I. Isildak, Anal. Chim. Acta 452 (2002) 29–34.
raphy as a patterning method and biomolecules as templates for [45] J.M.C.S. Magalhães, A.A.S.C. Machado, Talanta 47 (1998) 183–191.
G. Dhawan et al. / Biochemical Engineering Journal 44 (2009) 42–52 51

[46] S. Hirano, Production and application of chitin and chitosan in Japan, in: G. [98] S.K. Arya, P.R. Solanki, S.P. Singh, K. Kaneto, M.K. Pandey, M. Datta, B.D. Mal-
Skja-Braek, T. Anthonsen, P. Sandford (Eds.), Chitin and Chitosan, Sources, hotra, Biosens. Bioelectron. 22 (2007) 2516–2524.
Chemistry, Biochemistry, Physical Properties and Applications, Elsevier, Ams- [99] Ó.A. Loaiza, S. Campuzano, M. Pedrero, M. José, Pingarrón, Talanta 73 (2007)
terdam, 1988, p. 37. 838–844.
[47] M. Nakajima, K. Atsumi, K. Kifune, Development of absorbable sutures from [100] K.C. Grabar, R.G. Freeman, M.B. Hommer, M.J. Natan, Anal. Chem. 67 (1995)
chitin, in: J.P. Zikakis (Ed.), Chitin, Chitosan and Related Enzymes, Academic 735–743.
Press, New York, 1984, pp. 407–410. [101] M.-C. Daniel, D. Astruc, Chem. Rev. 104 (2004) 293–346.
[48] J. Ruzicka, E.H. Hansen, A.K. Ghose, H.A. Mottola, Anal. Chem. 51 (1979) [102] A.L. Crumbliss, S.C. Perine, J. Stonehuerner, K.R. Tubergen, J. Zhao, R.W.
199–203. Henkens, J.P. O’Daly, Biotechnol. Bioeng. 40 (1992) 483–490.
[49] P.M. Vadgama, K.G.M.M. Alberti, A.K. Covington, Anal. Chim. Acta 136 (1982) [103] Y. Yang, Z. Wang, M. Yang, M. Guoa, Z. Wu, G. Shen, R. Yu, Sens. Actuators B
403–406. 114 (2006) 1–8.
[50] I. Karube, E. Tamiya, J.M. Dicks, Anal. Chim. Acta 185 (1986) 195–200. [104] P.I. Gouma, Nanostructured polymorphic oxides for advanced chemosensors,
[51] G.G. Guilbault, J.G. Montalvo, J. Am. Chem. Soc. 91 (1969) 2164–2165. Rev. Adv. Mater. Sci. 5 (2003) 123.
[52] L. Campanella, F. Mazzei, M.P. Sammartino, M. Tomassetti, Bioelectrochem. [105] C.P. Huang, Y.K. Li, T.M. Chen, Biosens. Bioelectron. 22 (2007) 1835–1838.
Bioenerg. 23 (1990) 195–202. [106] J.H. Jin, S.H. Paek, C.W. Lee, N.K. Min, S.I. Hong, J. Korean Phys. Soc. 42 (2003)
[53] C. Eggenstein, M. Borchardt, D. Christoph, B. Gründig, C. Dumschat, K. Cam- S735–S738.
mann, M. Knoll, F. Spener, Biosens. Bioelectron. 14 (1999) 33–41. [107] D.H. Yun, M.J.J. Song, S.I. Hong, M.S. Kang, N.K. Min, J. Korean Phys. Soc. 47
[54] B.A. Petersson, Anal. Chim. Acta 209 (1988) 239–248. (2005) S445–S449.
[55] S.B. Butt, K. Cammann, Anal. Lett. 25 (1992) 1597–1615. [108] D.P. Nikolelis, C.O. Siontorou, Anal. Chem. 67 (1995) 936–944.
[56] J. Saurina, S. Hernandez-Cassou, E. Fabregas, S. Alegret, Anal. Chim. Acta 371 [109] J.K. Yang, K.S. Ha, H.S. Baek, S.S. Lee, M.L. Seo, Bull. Korean Chem. Soc. 25 (2004)
(1998) 49–56. 1499.
[57] S. Alegret, E. Martínez-Fábregas, Biosensors 4 (1989) 287–297. [110] J. Chen, S. Chiu, Enzyme Microb. Technol. 26 (2000) 359–367.
[58] S.A. Rosario, G.S. Cha, M.E. Meyerhoff, M. Trojanowicz, Anal. Chem. 62 (1990) [111] P.K. Srivastava, A.M. Kayastha, Srinivasan, Biotechnol. Appl. Biochem. 34
2418–2424. (2001) 55–62.
[59] S. Glab, R. Koncki, E. Kopczewska, I. Walcerz, A. Hulanicki, Talanta 41 (1994) [112] J.V. de Melo, S. Cosnier, C. Mousty, C. Martelet, N.J. Renault, Anal. Chem. 74
1201–1205. (2002) 4037–4043.
[60] M. Holmberg, M. Eriksson, C. Krantz-Rülcker, T. Artursson, F. Winquist, A. [113] H. Barhoumi, A. Maaref, M. Rammah, C. Martelet, N. Jaffrezic-Renault, C.
Lloyd-Spetz, I. Lundström, Sens. Actuators B 101 (2004) 213–223. Mousty, S. Cosnier, E. Perez, I. Rico-Lattes, Biosens. Bioelectron. 20 (2005)
[61] T. Krawczynski vel Krawczyk, M. Trojanowicz, A. Lewenstam, Talanta 41 (1994) 2318–2323.
1229–1236. [114] H.H. Weetall, N.B. Havewala, W.H. Pitcher, C.C. Detar, W.P. Vann, S. Yaverbaum,
[62] I. Walcerz, R. Koncki, E. Leszczynska, S. Glab, Anal. Chim. Acta 315 (1995) Biotechnol. Bioeng. 16 (1974) 295–313.
289–296. [115] Y.Y. Lee, A.R. Fratzke, K. Wun, G.T. Tsao, Biotechnol. Bioeng. 18 (1976) 389–413.
[63] B. Kovács, G. Nagy, R. Dombi, K. Tóth, Biosens. Bioelectron. 18 (2003) 111–118. [116] D.B. Johnson, D. Thornton, P.D. Ryan, Biochem. Soc. Trans. 2 (1974) 494.
[64] K. Yasuda, H. Miyagi, Y. Yamata, Y. Takata, Analyst 109 (1984) 61–64. [117] D. Thornton, A. Flynn, D.B. Johnson, P.D. Ryan, Biotechnol. Bioeng. 17 (1975)
[65] R.J. Forster, F. Regan, D. Diamond, Anal. Chem. 63 (1991) 876–882. 1679–1693.
[66] F. Sales, M.P. Callao, F.X. Rius, Analyst 124 (1999) 1045–1051. [118] D. Thornton, M.J. Byrne, A. Flynn, D.B. Johnson, Biochem. Soc. Trans. 2 (1974)
[67] N.C. Foulds, C.R. Lowe, J. Chem. Soc., Faraday Trans. 182 (1986) 1259–1264. 1360.
[68] G. Fortier, E. Brassard, D. Belanger, Biosens. Bioelectron. 5 (1990) 473–490. [119] K. Yoneyama, Y. Fujino, T. Osaka, I. Satoh, Sens. Actuators B 76 (2001) 152–157.
[69] S.E. Wolowacz, B.F.Y. Y-Hin, C.R. Lowe, Anal. Chem. 64 (1992) 1541. [120] A. Seki, S. Ikeda, I. Kubo, I. Karube, Anal. Chim. Acta 373 (1998) 9–13.
[70] B.F.Y. Y-Hin, M. Smolander, T. Cromptan, C.R. Lowe, Anal. Chem. 65 (1993) [121] D. Kirstein, F. Scheller, B. Olsson, G. Johansson, Anal. Chim. Acta 171 (1985)
2067–2071. 345–350.
[71] T. Kaku, H.I. Karan, Y. Okamoto, Anal. Chem. 66 (1994) 1231–1235. [122] M.L. Seo, J.S. Kim, S.S. Lee, Z.U. Bae, H.L. Lee, T.M. Park, J. Korean Chem. Soc. 37
[72] L. C-Guerente, S. Cosnier, C. Innocent, P. Mailley, Anal. Chim. Acta 311 (1995) (1993) 937.
23–30. [123] B. Liu, R. Hub, J. Dengay, Anal. Chim. Acta 341 (1997) 161–169.
[73] S. de Marcos, R. Hortigiiela, J. Galbfin, J.R. Castillo, O.S. Wolfbeis, Mikrochim. [124] A. Pizzariello, M. Stredanský, S. Stredanská, S. Miertus, Talanta 54 (2001)
Acta 130 (1999) 267–272. 763–772.
[74] S. Komaba, M. Seyama, T. Momma, T. Osaka, Electrochim. Acta 42 (1997) [125] M. Stred’anský, A. Pizzariello, S. Stred’anská, S. Miertuš, Anal. Chim. Acta 415
383–388. (2000) 151–157.
[75] S.B. Adeloju, S.J. Shaw, G.G. Wallace, Anal. Chim. Acta 341 (1997) 155–160. [126] S. Caras, J. Janata, Anal. Chem. 52 (1980) 1935–1937.
[76] S.B. Adeloju, S.J. Shaw, G.G. Wallace, Anal. Chim. Acta 281 (1993) 611–620. [127] W.Y. Lee, K.S. Lee, T.H. Kim, M.C. Shin, J.K. Park, Electroanalysis 12 (2000)
[77] S.B. Adeloju, S.J. Shaw, G.G. Wallace, Anal. Chim. Acta 323 (1996) 107–113. 78–82.
[78] S.B. Adeloju, S.J. Shaw, G.G. Wallace, Anal. Chim. Acta 281 (1993) 621–627. [128] S. Komaba, M. Seyama, K. Tanabe, T. Osaka, Denki Kagaku 64 (1996)
[79] Rajesh, V. Bisht, W. Takashima, K. Kaneto, Surf. Coat. Technol. 198 (2005) 1228–1233.
231–236. [129] D.G. Pijanowska, W. Torbicz, Sens. Actuators B 44 (1997) 370–376.
[80] Komaba, et al., Denki Kagaku 64 (1996) 1228–1233. [130] A. Senillou, N.J. Renault, C. Martelet, S. Cosnier, Anal. Chim. Acta 401 (1999)
[81] T. Osaka, S. Komaba, M. Seyama, K. Tanabe, Sens. Actuators B 35–36 (1996) 117–124.
463–469. [131] A.B. Kharitonov, M. Zayats, A. Lichtenstien, E. Katz, I. Willner, Sens. Actuators
[82] P.C. Pandey, G. Singh, Talanta 55 (2001) 773–782. B 70 (2000) 222–231.
[83] M.M.C. Ortega, D.E. Rodriguez, J.C. Encinas, M. Plascenia, F.A. Mendez Velarde, [132] S. Arisawa, R. Yamamoto, Thin Solid Films 210 (1992) 443–445.
R. Olayo, Sens. Actuators B 85 (2002) 19–25. [133] S. Poyard, D. Gorchkov, A. Jdanova, N.J. Renault, C. Martelet, A.P. Soldatkin, A.
[84] W.J. Cho, H.J. Huang, Anal. Chem. 70 (1998) 3946–3951. El’skaya, Biochim. Biochem. 319 (1996) 257–262.
[85] Y.C. Luo, J.S. Do, Biosens. Bioelectron. 20 (2004) 15–23. [134] G.A. Zhylyak, S.V. Dzyadevich, Y.I. Korpan, A.P. Soldatkin, A.V. El’skaya, Sens.
[86] A.S. Jdanova, S. Poyard, A.P. Soldatkin, N. Jaffrezic-Renault, C. Marteiett, Anal. Actuators B 24 (1995) 145–148.
Chim. Acta 321 (1996) 35–40. [135] T. Kuriyama, J. Kimura, Y. Kawana, Chem. Econ. Eng. Rev. 17 (1985) 22.
[87] W.Y. Lee, S.R. Kim, T.H. Kim, K.S. Lee, M.C. Shin, J.K. Park, Anal. Chim. Acta 404 [136] Y. Miyahara, T. Moriizumi, K. Ichimura, Sens. Actuators B 7 (1985) 1–10.
(2000) 195–203. [137] T. Kuriyama, S. Nakamoto, Y. Kawana, J. Kimura, New fabrication methods of
[88] H.C. Tsai, R.A. Doong, H.C. Chiang, K.T. Chen, Anal. Chim. Acta 481 (2003) enzyme immobilized membrane for ENFET, in: Proc. 2nd Int. Meet. Chemical
75–84. Sensors, Bordeaux, France, 7–10 July, 1986, pp. 568–571.
[89] H.C. Tsai, R.A. Doong, Biosens. Bioelectron. 20 (2005) 1796–1804. [138] S. Nakamoto, N. Ito, T. Kuriyama, J. Kimura, Sens. Actuators B 13 (1987)
[90] R. Sahney, S. Ananda, B.K. Puri, A.K. Srivastava, Anal. Chim. Acta 578 (2006) 165–169.
156–161. [139] J. Kimura, Y. Kawana, T. Kuriyama, Biosensors 4 (1988) 41–45.
[91] P.K. Jha, P.I. Gouma, Sensors, in: Proceedings of IEEE, vol. 24, 2004, pp. [140] Y. Miyahara, F. Matsu, T. Moriizumi, H. Matsuoka, I. Karube and S. Suzuki,
1199–1201. Micro-enzyme sensors using semiconductor and enzyme-immobilization
[92] R. Singhal, A. Gambhir, M.K. Pandey, S. Annapoorni, B.D. Malhotra, Biosens. techniques, Proc. Int. Meet. Chemical Sensors, Fukuoka, Japan, 19–22 Septem-
Bioelectron. 17 (2002) 697–703. ber 1983, Anal. Chem. Symp. Ser., vol. 17, 1983, pp. 501–506.
[93] A.P. Girard-Egrod, R.M. Morélis, P.R. Coulet, Thin Solid Films 292 (1997) [141] I. Karube, E. Tamiya, J.M. Dicks, M. Gotoh, Anal. Chim. Acta 185 (1986) 195–200.
282–289. [142] Y. Hanazato, M. Nakako, S. Shiono, IEEE Trans. Electr. Dev. ED 33 (1986) 47–51.
[94] A. Zhang, Y. Hou, N. Jaffrezic-Renault, J. Wan, A. Soldatkin, J.-M. Chovelon, [143] B.H. van der Schoot, P. Bergveld, Anal. Chim. Acta 199 (1987) 157–160.
Bioelectrochemistry 56 (2002) 157. [144] J.C. Chen, J.C. Chou, T.P. Sun, S.K. Hsiung, Sens. Actuators B 91 (2003) 180–186.
[95] Y. Hou, N.J. Renault, A. Zhang, J. Wan, A. Errachid, J.M. Chovelon, Sens. Actuators [145] N. Kazanskaya, A. Kukhtin, M. Manenkova, N. Reshetilov, Biosens. Bioelectron.
B 86 (2002) 143–149. 11 (1996) 253–261.
[96] P. Pandey, S.P. Singh, S.K. Arya, V. Gupta, M. Datta, S. Singh, B.D. Malhotra, [146] D.V. Gorchkov, S. Poyard, A.P. Soldatkin, N. Jaffrezic-Renault, C. Martelet,
Langmuir 23 (2007) 3333–3337. Mater. Sci. Eng. C 5 (1997) 29–34.
[97] K. Sawickaa, P. Goumab, S. Simonc, Sens. Actuators B 108 (2005) 585–588. [147] D.G. Pijaowska, W. Torbicz, Sens. Actuators B 44 (1997) 370–376.
52 G. Dhawan et al. / Biochemical Engineering Journal 44 (2009) 42–52

[148] O.A. Boubriak, A.P. Soldatkin, N.F. Starodub, A.K. Sandrovsky, A.K. El’skaya, [152] National Kidney Foundation India (http://www.nkfi.in/).
Sens. Actuators B 27 (1995) 429–431. [153] E. Jacobs, E. Vadasdi, L. Sarkozi, N. Colman, Clin. Chem. 39 (1993)
[149] A.P. Soldatkin, J. Montoriol, W. Sant, C. Martelet, N. Jaffrezic-Renault, Biosens. 1069–1074.
Bioelectron. 19 (2003) 131–135. [154] C. Ritter, F. Heike, K. Herbert, K.F. Josef, L. Susanne, N. Christian, O. Helmut, P.
[150] NKF-K/DOQI Clinical Practice Guidelines for Chronic Kidney Disease: Evalua- Gabriele, S. Bernhard, S. Marieluise, S. Wolfgang, S. Gregor, Sens. Actuators B
tion, Classification and Stratification. 76 (2001) 220.
[151] Kidney and Urologic Diseases Statistics for the United States, from
the National Kidney and Urologic Diseases Information Clearinghouse
(http://www.niddk.nih.gov/).

View publication stats

You might also like