Anal Bioanal Chem (2013) 405:3753–3771
DOI 10.1007/s00216-012-6578-2
REVIEW
New trends in the electrochemical sensing of dopamine
Krystyna Jackowska & Pawel Krysinski
Received: 25 September 2012 / Revised: 12 November 2012 / Accepted: 13 November 2012 / Published online: 16 December 2012
# The Author(s) 2012 This article is published with open access at Springerlink.com
Abstract Since the early 70s electrochemistry has been used
as a powerful analytical technique for monitoring electroactive species in living organisms. In particular, after extremely
rapid evolution of new micro and nanotechnology it has been
established as an invaluable technique ranging from experiments in vivo to measurement of exocytosis during communication between cells under in vitro conditions. This review
highlights recent advances in the development of electrochemical sensors for selective sensing of one of the most
important neurotransmitters—dopamine. Dopamine is an
electroactive catecholamine neurotransmitter, abundant in
the mammalian central nervous system, affecting both cognitive and behavioral functions of living organisms. We have
not attempted to cover a large time-span nor to be comprehensive in presenting the vast literature devoted to electrochemical dopamine sensing. Instead, we have focused on the
last five years, describing recent progress as well as showing
some problems and directions for future development.
Keywords Dopamine . Biosensors . Sensors .
In vivo detection . Implantable sensors
Introduction
Dopamine (DA), which belongs to the catecholamine family
of neurotransmitters, is crucially important in humans. It is
produced in adrenal glands and several areas of the brain,
Published in the topical collection Bioelectroanalysis with guest
editors Nicolas Plumeré, Magdalena Gebala, and Wolfgang
Schuhmann.
K. Jackowska : P. Krysinski (*)
Faculty of Chemistry, University of Warsaw,
02-093 Warsaw, Poland
e-mail:
[email protected]
and is the most abundant of the catecholamines involved in
brain–body integration (Fig. 1).
Dopamine is formed by decarboxylation of DOPA and is
a precursor of two other neurotransmitters—adrenaline and
noradrenalin (Scheme 1).
Under physiological conditions in the nervous tissues and
body fluids, dopamine occurs in the form of large organic
cations.
As a potent neuromodulator, it affects many aspects of
brain circuitry, neuronal plasticity, and the organization and
control of stress responses; it is also of crucial importance in
the attention span, learning, and memory [1–4]. Dopamine
also affects the cardiovascular and renal systems [5–7] and a
variety of motivated behavior (e.g., perceiving rewards and
pleasure). Low levels or practically complete depletion of
dopamine in the central nervous system is implicated as a
major cause of several neurological diseases, for example
schizophrenia, Parkinson’s disease, and ADHD/ADD
[8–11]. In the neurotransmission process, dopamine acts as
an extrasynaptic messenger molecule via volume transmission
across the synaptic cleft to bind to extrasynaptic receptors (G
protein-coupled receptors) and transporters [12–16] (Fig. 2).
Because all drugs of abuse affect the dopaminergic pathways,
much study of addiction focuses on dopamine [17].
Given the wide range of physiological and pathophysiological implications, the development of a sensor for precise
and selective measurement of dopamine (and/or its metabolites) at the low levels characteristic of living system (26–
40 nmolL−1 [18, 19] and below) can make a great contribution to disease diagnosis.
Many recently published papers have been devoted to
detection of dopamine. Sampling and separation techniques
for cell or tissue homogenates (microdialysis, low-flow push–
pull perfusion) coupled to analytical techniques such as enzyme assays, liquid chromatography, mass spectrometry, capillary electrophoresis, with optical or electrochemical
3754
K. Jackowska, P. Krysinski
Fig. 1 Integrated view of
brain–body interactions in
response to stressful stimuli.
Acute stressors change the
release of different
neurotransmitters in limbic
areas of the brain (large
rectangle), which are
functionally interconnected,
leading to behavioral processes
for coping with the stress. The
hypothalamic activation
initiates the release of different
body hormones and peptides,
which cross the blood–brain
barrier to feed back to the brain.
Among these, corticosterone is
important in modulating release
of different neurotransmitters in
the limbic areas, affecting, in
turn, coping behavior (from
Ref. [1], with permission)
detection (e.g., biosensors) are the main techniques used for
measurement of dopamine (and neurotransmitters in general).
However, the most straightforward, rapid, and cost-effective is
the use of electrochemical sensing techniques for detection of
electroactive neurotransmitters, including dopamine.
Recent developments in electrochemical sensors focus on
improvement of their sensitivity, selectivity, and biocompatibility. Miniaturization of such sensors has also been attempted, with the objective of achieving better spatial and
temporal resolution, and effective multi-analyte detection.
It is beyond the scope of this brief review to cover all these
topics. The interested reader is referred to the paper by
Kennedy [20], and references cited therein. In this paper
we focus solely on amperometric sensors for dopamine.
The 2e/2H+ redox reactions of dopamine under physiological conditions constitute the basis of electrochemical detection (e.g., amperometric, differential pulse voltammetry,
cyclic voltammetry) of this neurotransmitter. Currents generated in these reactions may be linearly dependent on the
concentration of the electroactive molecules in the extracellular medium, enabling quantification of these compounds.
DOPA
Dopamine
Noradrenaline
Adrenaline
Scheme 1 Sequence of enzymatic reactions generating other catecholamines from L-DOPA
Appropriate design of a sensor system should enable numerous limitations of direct electrochemical detection of catecholamines to be overcome. Electrochemical methods have been
widely used for direct measurement for dopamine, with several advantages, for example as rapid response with high
accuracy and relatively easy operation. However, early work
by Hubbard and coworkers revealed the effects of adsorption,
orientation of adsorbed molecules, concentration, supporting
electrolyte, etc., on the electrochemistry of dopamine [21, 22].
Other limitations include the relatively large oxidation potential of direct oxidative detection on the electrode and formation of phenoxy radicals with subsequent coupling and
formation of passivating polymeric film on the electrode
[23–25], consecutive reactions of dopaminoquinone (DOQ),
a product of electrochemical DA oxidation, resulting in aminochrome formation and electrode passivation [26]. The other
problem which must be solved in the electrochemical detection of dopamine is the co-existence of many interfering
compounds in biological systems. Among these interfering
compounds, ascorbic acid (AA), epinephrine (EP), and uric
acid (UA) are particularly important because they can all be
oxidized at similar potentials resulting in overlap of voltammetric responses [27–29]. It is, therefore, important to develop
an electrochemical method for selective determination of dopamine with high sensitivity and a low detection limit (LOD).
In recent years, numerous attempts have been made to
develop such a method. Among these, different electrode
materials have been tested, including metal, carbon, and
boron-doped diamond electrodes [30, 31], and different electrolytes, including ionic liquids [32]. Modified electrodes with
self assembled monolayers [33, 34], covalent modification
[31, 35], nanotubes, nanowires, nanoparticles [36–42], permselective membranes [43], and conducting polymer films have
New trends in the electrochemical sensing of dopamine
3755
Fig. 2 Synaptic region
coupling two cells, separated by
a synaptic cleft (SC); arrows
show dopamine trafficking.
ME, metabolic enzymes; DR,
DU, neuronal dopamine release
and uptake, respectively; PR, R,
pre-synaptic and post-synaptic
receptors, respectively; V, DA
storage vesicles; SO, synaptic
overflow
been reported to have favorable effects on detection of DA in
the presence of AA, UA, or both AA and UA [38, 44–47].
Conducting polymers have several advantages in the sensor
preparation for DA detection. Some of these have pendant
groups (amine or carboxylic) which interact (e.g., electrostatically) with the analyte or interferent molecules, which results
in better separation of the redox signals of DA, AA, and UA
[44, 46, 48, 49]. Another approach is to modify the electrode
surface for immobilization of enzymes (e.g., tyrosinase), selective toward DA molecules [49–55].
Electrochemical sensors
Electrochemical sensors can be broadly divided in two
classes—biosensors and chemical sensors. According to
the IUPAC definition [56]: “the electrochemical biosensor is a self-contained integrated device, which is capable of providing specific quantitative or semi
quantitative analytical information using a biological
recognition element (biochemical receptor) which is
retained in direct spatial contact with an electrochemical transduction element”. In the literature, however,
many authors use a wider definition of biosensors, focused not on the indispensability of a biological recognition element but on the type of analyte. Throughout
this review, we will use the term “biosensor” according
to the IUPAC convention.
In contrast with biosensors, chemical sensors contain
non-biologically active elements, improving their sensitivity and selectivity in analyte detection. In electrochemical sensors, modified electrodes are widely used
as sensing elements. These modified electrodes can be
formed on the basis of different inorganic or organic
materials characterized by good electrical conductivity
and catalytic properties.
Biosensors in dopamine detection
In recent years, much effort has been devoted to improving the
catalytic properties, sensitivity, and selectivity of electrochemical sensors, either by application of nanoscale materials (nanoparticles, nanowires, nanotubes) [57–62] or by use of advanced
membrane materials (sol–gel composites, hydrogels, lipid
membranes) for biomolecule (enzyme) immobilization [63].
Many of the electrochemical biosensors tested for dopamine
detection utilize tyrosinase (Tyr), the enzyme also known as
polyphenol oxidase (PPO), as recognition element. Tyrosinase
is a multifunctional copper-containing oxidoreductase with two
types of catalytic activity in the presence of oxygen:
1. hydroxylation of monophenols to o-diphenols by cresolase activity; and
2. oxidation of diphenols to o-quinones by catecholase
activity [64, 65].
The resulting quinones can be further reduced electrochemically on the electrode without any mediator, reforming
the original o-diphenol. This reaction constitutes the basis of
amperometric detection at negative potentials and quantification of phenolic compounds. By the same principle, dopamine, a catechol-like phenolic compound, can be detected
by amperometry with an electrode modified with tyrosinase.
Application of tyrosinase to electrochemical sensing of dopamine eliminates several problems connected with consecutive reactions of DA oxidation products resulting in the
formation of aminochrome [26]. Application of tyrosinase,
which is sensitive to some classes of reagent only, can also
reduce the problem of many interfering compounds in biological systems, for example ascorbic acid (AA), uric acid
(UA), and other neurotransmitters.
The crucial issue in the development of electrochemical
biosensors is immobilization of the enzyme on a conductive
3756
surface (electrochemical transducer). Some prerequisites
should be fulfilled by the enzyme after immobilization:
&
&
&
&
it must be efficiently immobilized on the surface, so loss
of enzyme does not occur during measurement;
it must retain its biological activity;
it should be compatible with and chemically inert toward the host structure; and
it should be accessible to the analyte;
There are also requirements of the matrices used for
enzyme immobilization:
&
&
&
&
the procedure for matrix formation should be simple,
reproducible, and inexpensive;
the matrix should be chemically stable in the analytical
environment and inert to the enzyme and detected
analytes;
if the matrix works as a mediator in an electron-transfer
process, its properties should be reversible; and
the conductivity of the matrix should be high, especially
when used in an electrochemical biosensor.
The procedures used for immobilization of enzymes on
different substrates, and their advantages and disadvantages,
have been widely discussed in many reviews of nanoscale
materials [57, 58, 62] and conducting polymers [66–69]. A
review has been devoted to immobilization of laccase (Lac)
and tyrosinase (Tyr) on different substrates [70]. Data were
reported for application of biosensors to detection of phenolic compounds; there were no data on dopamine detection, however.
Methods of enzyme immobilization can be divided into
two main groups—entrapment in the bulk of the matrix and
entrapment on the matrix surface. Entrapment in the bulk is
expected when the active layer of the biosensor is formed by
mixing the enzyme with the compound used for matrix
formation, e.g., carbon paste, or monomers for chemical or
electrochemical polymerization of polymers. When sensor
active layers are formed by the sol–gel or layer-by-layer
techniques, the bulk entrapment is, again, expected to be
dominant. Of course, some of the enzyme is also bonded to
the matrix surface, but most is trapped within the bulk. In
surface-entrapment processes, physical or chemical adsorption are used, or the enzymes are immobilized by covalent
bonding to appropriately functionalized matrices. The localization and distribution of the enzyme within the matrix is
crucial and can determine the characteristics and performance of the biosensor. For instance, diffusion of analyte
to enzyme trapped in the bulk of the matrix is limited;
physical adsorption, however, the weakness of the physical
interaction and the susceptibility of the binding energy to the
environment (pH, ionic strength, type of ions) can result in
K. Jackowska, P. Krysinski
enzyme losses and instability during practical application. In
comparison, covalent linking of the enzyme to the matrix
seems to be the most effective means of immobilization and
formation of stable biosensors. However, when surface attachment is used for biosensor formation, the main problem
is insufficient enzyme (a monolayer on a surface matrix)
resulting in reduced biosensor sensitivity. Another limitation
is proper orientation of covalently bound enzyme; this also
should be taken into account for biosensors based on the socalled direct electron transfer (DET) mechanism. It is beyond the scope of this review to discuss this type of sensor,
which has already been thoroughly discussed in an excellent
review by Armstrong [71].
Many efforts have been made to improve the properties of
the active layer of biosensors, for example accessibility to
dopamine and loading capacity, by use of conducting porous
and mesoporous materials. Similarly, increased surface active
site density for enzyme immobilization can be achieved by
constructing the matrix with nano-scale materials, for example
metal or oxide nanoparticles, carbon nanowires , nanotubes,
etc. These not only increase the amount of enzyme attached to
the matrix but can also catalyze the enzymatic reaction, affecting kinetic data, sensitivity, and response time of biosensors. Such an approach has also been used for formation of Tyr
biosensors for dopamine detection [49–55, 72–78]. Information about different types of matrix for tyrosinase immobilization, type of entrapment, analytical data (sensitivity,
detection limit, linear range), and interferences are collected
in Table 1. All the experimental results presented in this table
were recalculated to present them in the same units for comparison purposes.
The best characteristics for dopamine biosensors were
obtained by Njagi et al. [55] and Wang et al. [78]. They
achieved a single-figure nanomole level detection limit for
dopamine. However, it is very difficult to compare the
results, because tyrosinase activity is very rarely determined
after its entrapment in the matrix (in the bulk or attached to
the surface). Also the enzyme loading is generally unknown
and uncontrolled. Such data are given in a few publications
only [55, 76, 77].
Response time is another important analytical property
describing Tyr biosensor performance [50, 51, 54, 55, 75,
76]. The lowest response time (<5 s) for DA detection has
been obtained with biosensors containing SWNT-PPy (single-wall carbon nanotubes–polypyrrole) as matrix with tyrosinase covalently bound to nanotubes functionalized with
carboxyl groups [54]. The effect of matrix composition on
response time was tested by Njagi et al. [55]. They obtained
the lowest response times (<8 s) for biosensors with a
relatively complex matrix consisting of Tyr-Chit/CF and
Tyr/TiO2/CeO2/Chit/CF, where CF and Chit denote carbon
fiber and chitosan, respectively. Replacement of chitosan in
such a matrix by a sol–gel system resulted in a three or
New trends in the electrochemical sensing of dopamine
3757
Table 1 Analytical characteristics for electrochemical detection of dopamine by use of different tyrosinase biosensors
Layer description (immobilization matrix)
Interferents
Linearity (molL−1)
C paste, b.entrp
Polyacrylamide gel, b.entrp
AA, DOPA, DOPAC
Without AA
AA
AA
1.5×10−5–2.5×10−4
5×10−6–1×10−4
3×10−8–1×10−6
5×10−8–8×10−5
AA
Up to 2×10−4
2×10−6–1×10−5
5×10−6– 1.2×10−4
5×10−6–2.3×10−5
5×10−7–2.5×10−4
Polypyrrole modified, b.entrp
Poly(vinyl alcohol) ferrocene-Pd-sol-gel, b.entrp
Polythiophene, b.entrp
Biocomposite gel (agar, guar gum), b.entrp
Amine-terminated BDD, surf., cov. bond.
MWNT/Nafion, b. surf. entrp
Eggshell membrane, GA, surf., phys. ads
Chit/GC
TiO2/CeO2/Chit/GC, b. surf. entrp
SWNT/Ppy, surf cov. bond
Fe3O4/Chit, b. surf. entrp
TiO2/CeO2/Chit/CF
Chit/CF
TiO2/CeO2/sol-gel/CF
sol-gel/CF
PIn5COOH surf. cov. bond
AA
AA
AA
AA ,UA, DOPA, DOPAC,
SE, EP, NorEP
AA <1×10−4 molL−1
UA <1×10−4 molL−1
AA/UA 12.5
5×10−6–5×10−5
2×10−8–7.5×10−5
1×10−8–2.2×10−4
5×10−8–1.2×10−4
1×10−7–1.8×10−4
2×10−7–2.6×10−4
5×10−7–2×10−5
Sensitivity
(mAmol−1 L)
6.3
59
0.125
133
68.6
12
9.7
14.9
467
46
14.2
7.3
33.4
6.4
2,200
6,200
1,290
Detection limit
(molL−1)
Ref.
1.5×10 −5
5×10−6
3×10−8
5×10−8
5×10−5
1×10−7
9×10−7
[72]
[73]
5.2×10−7
2.5×10−6
1.9×10−6
2.4×10−8
5×10−6
6×10−9
1.1×10−9
1.4×10−7
4×10−8
8×10−8
1×10−7
[74]
[75]
[76]
[77]
[50]
[51]
[52]
[53]
[54]
[78]
[55]
[49]
b.entrp., entrapment of tyrosinase in matrix; surf. entrp., entrapment of tyrosinase on matrix surface; surf., cov. bond, entrapment of tyrosinase on
surface, covalent bonding; surf., phys. ads, entrapment of tyrosinase on surface, physical adsorption; Chit, chitosan; CF, carbon fiber; PIn5COOH,
poly(indolo-5-carboxylic acid); SWNT, single walled carbon nanotube; MWNT, multi walled carbon nanotube; BDD, boron-doped diamond; GC,
glassy carbon; GA, glutaraldehyde; AA, ascorbic acid; UA, uric acid; DOPA, 3,4-dihydroxy-L-phenylalanine; DOPAC, dihydroxyphenylacetic
acid; SE, serotonin; EP, epinephrine; NorEP, norepinephrine
fourfold increase of response time. This was ascribed to the
diffusion barrier imposed by this matrix.
The important features for practical use of biosensors are:
1. sensor stability;
2. reproducibility of sensor readout, and
3. life-time.
These properties depend very much on biosensor construction and its working environment (pH, ions, interferents, inhibitors). The reproducibility, determined in several
cases [54, 55, 78], has been found to be in the range 2.2–
2.5 %. The stability varied from a week to a few months [52,
55, 75, 77, 78]. The biosensors usually retained approximately 90–95 % of their original response after one week
under the same experimental conditions. In the long-term
the stability gradually decreased to 70 % or less. The best
results were obtained by Pandey et al. [75]. They found their
biosensor was stable over three months, retaining 95 %
reproducibility when stored at 4 °C. However, the procedure
used to evaluate stability was not described precisely. The
most complex measurements of stability and reproducibility
were made by Njagi et al. [55] for tyrosinase biosensors
containing mixed oxides (TiO2, CeO2). It was shown that
the presence of oxides improved biosensor stability; this can
be attributed to strong adsorption of the enzyme on the
metal oxides.
For practical purposes dopamine biosensors should be
tested using real biological systems, and several papers
report testing of Tyr biosensors in real biological (in vivo)
measurements [55, 74, 77]. In another strategy, detection of
dopamine was performed in the presence of different interferents, usually ascorbic acid (AA) [49, 50, 53–55, 72, 73]
but also uric acid [49, 55], L-DOPA, DOPAC [55, 72], and
even the presence of another catecholamines—serotonin
(SE), norepinephrine (NorEP), and epinephrine (EP) [55].
It is generally accepted that Tyr biosensors can be used for
selective detection of dopamine in the presence of AA, UA,
and SE.
As already mentioned, in practical use all of the properties of dopamine biosensors discussed above are very important. However, to understand the mechanism of the
response of the biosensor, why interferents affects its analytical performance, analytical performance, for example
3758
detection limit and sensitivity (assuming similar values of
enzyme activity and loading) depend substantially on the
type of matrix and on immobilization procedures, one
should also determine physicochemical data for the enzyme,
namely the values of KM, Vmax, and kcat. These data are
crucial for proper characterization of enzymatic reactions.
KM, the so-called Michaelis–Menten constant is defined as
the substrate concentration with which half of the maximum
reaction rate is achieved, or half of the active enzyme sites
are filled. Vmax is the maximum rate of the enzymatic
reaction (maximum turnover rate), and kcat reflects the turnover number of the enzyme, i.e. the number of molecules of
substrate converted to product per unit time when the enzyme is working at its maximum efficiency.
A popular definition of KM (but only valid when there is
a appropriate relationship between rate constants—k1 >k2
for the enzymatic reaction, where k1 and k2 are the reaction
rate constants for complex formation and dissociation, respectively) is that it is a measure of the strength of enzyme–
substrate [ES] complexes or a measure of the affinity of
formation of such complexes. There are many publications
on different biosensors, and on detection of different substrates (analytes); surprisingly enough, however, kinetic data for such sensors are very limited. Values of KM, Vmax
(or Imax), with the active layer and type of Tyr entrapment in
biosensors for dopamine detection are listed in Table 2. The
values of KM and Vmax for dopamine in solution containing
tyrosinase were determined spectrophotometrically. They
are 2.0×10−4 molL−1 and 5.2×10−5 molL−1 min−1, respectively [77]. By comparing this KM value with those obtained
for immobilized tyrosinase one can conclude that entrapment of Tyr in carbon paste has not affected the catalytic
step of the reaction. The higher KM values obtained for
chitosan, from eggshell membranes, results from some diffusion limitation for dopamine. The lowest value of KM was
obtained for the enzyme covalently bound to the carboxyl
groups of the polymer. However, for proper characterization
of biosensors we must know not only kinetic data but also
the enzyme load and its activity after immobilization. Unfortunately, such complex investigations have not been conducted. There is also no specific information about the effect
of interferents, for example AA and UA, on kinetic data for
the enzymatic reaction.
Chemical sensors in dopamine detection
As remarked above, biosensors based on the enzyme tyrosinase have been successfully used for selective determination of dopamine in the presence of ascorbic and uric acids
in a biological environment. The main problem in the practical application of such biosensors is their poor long-term
stability and reproducibility. One strategy used to overcome
K. Jackowska, P. Krysinski
problems with DA detection is the development chemical
sensors, usually chemically modified electrodes. In the last
decade many of different materials have been tested to
improve catalytic affinity, to enlarge the reaction space,
and to eliminate the effect of interferents. A huge amount
of research has resulted in more than hundred publications
in the last three years. In 2010, Compton et al. [79] reviewed
the literature on amperometric detection of DA in the presence of interferents, for example AA, UA, and serotonin .
They provided a brief description of modifying layers: polymers, carbon nanotubes, and other types of surface modification. In this section we focus on some new materials
used for electrode modification, for example graphene
[80–90], ionic liquids [91–98], and nanoparticles [99–119].
We mostly discuss articles published within the last three
years. The linear range of DA detection, sensitivity, detection limit, and short descriptions of electrode composition
are presented in Tables 3–6.
Graphene
Data for selective determination of DA by use of electrodes
modified with graphene (GR) are listed in Table 3 [81–89].
An excellent, short review on the applications of graphene
for electrochemical sensing and biosensing was recently
published by Pumera et al. [80]. Graphene is a very promising material for electrochemical applications because of its
very large electrical conductivity, large surface area, and
relatively low cost of preparation. Electrodes can be modified with graphene suspension [81], powder [87], and graphene composites [86]. Graphene sheets [88] can also be
used as electrodes. Use of reduced graphene oxide (rGO) for
DA detection has recently been studied [82–86, 88]. It is
accepted that the reduced GO enhanced the electrochemical
response as a result of the presence of oxygen-containing
groups on the reduced GO surface [80]. In many articles the
rGO is also called reduced graphene (rGR), but this is not
correct. The DA sensors with the best detection limit (1–
10 nmolL−1) and linear range were obtained by use of
electrodes modified with graphene prepared by GO reduction [82–86, 88]. It is difficult to compare the results because of lack of information about the surface properties of
he graphene-modified electrodes. Some FTIR and Raman
spectra can be found for graphene electrodes modified with
organic compounds [82, 84, 88], and XRD and TEM have
been used for their characterization [89]. There are no kinetic data for DA oxidation. The stability and reproducibility of electrodes modified with quercetin (Qu)/GR [82] and
PPy/GR [84] have been investigated. The relative standard
deviation was found to be lower than 2 % [82] or 3 % [84]
and stability was approximately two months for Qu/GR.
Graphene-based electrodes have been used to monitor DA
in biological samples, for example rat striatum [82], mouse
New trends in the electrochemical sensing of dopamine
Table 2 Kinetic data for tyrosinase biosensors for dopamine
3759
Layer description (immobilization)
KM (molL−1)
Vmax or Imax
Ref.
C paste, b.entrp
Biocomposite gel (agar, guar gum), b.entrp
MWNT/Nafion, b, surf. entrp
Chitosan, b. surf. entrp, TiO2/CeO2/Chit/GC,
b. surf. entrp
Eggshell membrane, GA, surf, phys. ads
PIn5COOH surf. attach, cov. bond
1.1–2.3×10−4
5×10−6
6×10−5
7.99×10−4
0.342 μA
–
–
–
[72]
[77]
[51]
[53]
The abbreviations are explained
in the footnote to Table 1
hippocampus tissues [83], and human serum and human
urine [84, 85].They have also been used for simultaneous
determination of DA, AA, and UA [90].
Ionic liquid composite electrodes
Wai and Ivaska [96] recently reviewed the applications of
ionic liquids (IL) in electrochemical sensors. The ionic
liquids have unique properties—they are salts with melting
points near room temperature, they have a high ionic conductivity and large electrochemical windows, and can accelerate electron transfer in several reactions. Different ionic
liquids have been used to form composite IL electrodes—
OPPF 6–N-octylpyridium hexafluorophosphate [91, 92],
BMIMPF6–1-butyl-3-methlimidazoliumhexafluorophosphate [93, 95], and PMIMCH 3 COO–1-(3-chlorine-2hydroxypropyl)-3-methylimidazolium acetate [94]. The
worst detection limit (LOD) was obtained for carbon paste
containing PMIMCH3COO [94]. The low LOD of this
electrode could, however, result not only from the type of
IL but also from application of Ni/Al hydroxide layers. The
kinetic data for the sensor developed in this work, for
example the electron-transfer rate constant, kct, and the
6.7×10−4
1.2×10−6
1.33×10−6
(AA, UA>100 μmolL−1)
0.08 mmolL−1 min−1
5.5 μAcm–2
16.5 μAcm–2
[52]
[49]
charge transfer coefficient, α, were 1.66 s−1 and 0.43, respectively. These data are somewhat lower than those
obtained for functionalized ordered mesoporous carbon/ionic liquid gel/glassy carbon (f-OMC/ILgel/GC) electrodes
[95], for which the detection limit was 4.1×10−9 molL−1,
ks 02.5 s−1, and α00.55. For the f-OMC/GC electrode, but
without IL, these authors obtained the lower value of ks 0
0.97 s−1, indicating that the presence of IL facilitates electron transfer between dopamine and the electrode surface
[95]. The stability and reproducibility of such electrodes
modified with IL were also examined. The RSD was in
the range 1.5–3 % [91, 92, 94] and stability was ca 1 month
for the Ni/Al-layered double hydroxide/carbon ionic liquid
electrode (Ni/Al-LDH/CILE, ca 10 % response decrease).
The molecular sieve/carbon ionic liquid electrode (MS/
CILE) was used for monitoring DA levels in human serum
[91]. Analytical data for electrodes containing ionic liquids
(IL) are listed in Table 4.
Metal and semiconductor nanoparticles
Substantial growth of applications of metal and semiconductor nanoparticles for electroanalysis has recently been
Table 3 Analytical data for electrooxidation of dopamine at graphene-modified electrodes
Electrode
Method
Interferents
Linearity
(molL−1)
PPy/GR/GC
Qu/GR/GC
GR/CFE
PPy/rGO/GC
TCPP/GR/GC
EDTA/GR/Nafion/GC silanized
GR/GC
β-CD/GR/GC sheet
GR-Chit/GC
CV
DPV
CV
DPV
DPV
DPV
DPV
CV
CV
AA
AA,
AA,
AA,
AA,
AA
AA
AA
AA,
5×10−7–1×10−5
3×10−8–4×10−4
1×10−8–1×10−4
1×10−7–1.5×10−4
1×10−7–1×10−6
2×10−7–2.5×10−6
4×10−6–1×10−4
9×10−9–1.3×10−5
5×10−6–2×10−4
UA, others
UA
UA
UA
UA
Sensitivity
(mAmol−1 L)
Detection limit
(molL−1)
Ref.
94
–
–
–
–
–
–
–
–
1×10−7
1×10−8
1×10−8
2.3×10−8
2.2×10−8
1×10−8
2.64×10−6
5×10−9
–
[81]
[82]
[83]
[84]
[85]
[86]
[87]
[88]
[89]
GR, graphene; rGO, reduced graphene oxide; GC, glassy carbon; CF, carbon fiber; PPy, polypyrrole; β-CD, β-cyclodextrin; Chit, chitosan; Qu,
quercetin; TCPP, meso-tetra(4-N,N,N,N-trimethlanilinium)porphyrin
3760
K. Jackowska, P. Krysinski
Table 4 Analytical data for electrooxidation of dopamine on ionic liquid-modified electrodes
Electrode
Method
Interferents
Linearity (mol L−1)
Detection limit (mol L−1)
Ref.
MS/CILE
Pglycine/CILE
SCP/CILE
Ni/Al-LDH/CILE
f-OMC/ILgel/GC
SWV
SWV
SWV
CV
DPV
AA
AA
AA
AA, UA, others
AA, UA
5×10−8–8×10−4
1×10−7–3×10−4
to 1×10−4
1×10−5–1.1×10−3
1×10−7–5×10−4
1×10−8
5×10−9
2.6×10−7
5×10−6
4.1×10−9
[91]
[92]
[93]
[94]
[95]
MS, molecular sieve; CILE, carbon ionic liquid electrode; Pglycine, polyglycine; SCP, screen printing; Ni/Al-LDH, Ni/Al, layered double
hydroxide; f-OMC/ILgel, functionalized ordered mesoporous carbon/ionic liquid gel
observed [62, 118, 119]. These nanoparticles perform several
important functions, including an increase of the sensing
surface and facilitation of electron transfer. They also facilitate
electrical contact between the redox center of a biomolecule
and the electrode surface. They can also be modified further.
Analytical data obtained from use of electrodes modified with
nanoparticles to detect DA in the presence of AA and UA are
listed in Table 5. Data for simultaneous determination of DA,
AA, and UA are listed in Table 6
Comparing of detection limits showed the best result for
selective determination of DA in the presence of AA was
obtained by use of a sensor containing conducting polymer
(PEDOT) and inorganic Cu crystals of appropriate size [110].
It was demonstrated that optimization of the thickness of
PEDOT, surface coverage of PEDOT by Cu nanocrystals,
and their diameter, resulted in a sensor operating in the nanomolar concentration range. This high sensitivity was related to
formation of a Cu(II)–o-quinolate complex and its promotion
of subsequent DA oxidation [110]. There is insufficient data
describing the properties of the surface of electrodes modified
with nano-scale materials and used for DA monitoring. Some
information on surface coverage, morphology, and loading
can be found for PNPy/Pd(nanoclusters) [112] and Au(nanoparticles)/PANI [111] composites. We were impressed by the
publication describing an electrode with methylene blue
adsorbed on a phosphorylated zirconia–silica (MB)SZP for
which surface properties and electrode composition, for example amount of zirconia, phosphorus content, surface area,
the type of surface bond, were determined [116]. Analytical
data for electrodes used for simultaneous determination of DA
and other analytes show the best results for simultaneous
determination of DA and UA were obtained by use of an
electrode modified with polypyrrole and Cu nanoparticles,
CuNP/Ppy [113]; for detection of DA in the presence of AA,
Table 5 Analytical data for electrooxidation of dopamine on modified electrode with different nano-scale materials (selective detection of
dopamine)
Electrode
Method
Interferents
Linearity (molL−1)
Sensitivity
(mAmol−1 L)
Detection limit
(molL−1)
Ref.
CD-MWCNT Plu/AuNP/GC
Pt-DEN/GC
AuNP
Au/Pt/Pd/TiO2NT/Ti
AuNP/TiO2NT/Ti
DHBPD/TiO2NP/CP
CNP/f-silicate particles/ITO
Au nanostructured
a) CuONP/CP
b) CuONP/CP
nanoZnO/GC
Fe2O3-SWCNT/PGE
Cu(CR)-PEDOT/Pt
DPV
FIA
UV–visible
DPV
CV
DPV
DPV
DPV
DPV
AA, UA, other
UA
AA, DOPA, other
AA, UA
AA
AA
AA
–
–
–
–
–
–
–
139
–
–
–
–
9,000
313
1.9×10−7
19 ppb
3.6×10−7
3×10−8
–
3.14×10−8
3.6×10−7
5×10−6
1.8×10−7
5.5×10−8
6×10−8
3.7×10−7
4×10−9
[99]
[100]
[101]
[102]
[103]
[104]
[105]
[106]
[107]
DPV
SWV
DPV
1×10−6–5×10−5
–
5.4×10−7–5.4×10−6
5×10−8–3×10−5
5×10−4–2.5×10−3
8×10−8–2×10−5
3×10−7–1.8×10−5
1×10−6–1×10−4
3×10−7
2×10−5
1.1×10−7–8×10−4
3.2×10−6–3.1×10−5
1.2×10−8–6.9×10−8
3×10−7–2×10−6
AA, UA, NADH, other
AA
AA
[108]
[109]
[110]
β-CD, β-cyclodextrin; MWCNT, multi-walled carbon nanotubes; SWCNT, single-walled carbon nanotubes; Plu, poly(luminal); NP, nanoparticles;
NT, nanotubes; DEN, dendrimers; f, functionalized; CP, carbon paste; PGE, pyrolytic graphite electrode; DHBPD, N,N-(2,3-dixydroxybenzylidine)-1,4-phenylenediamine; CR, crystals; PEDOT, poly(3,4-ethylenedioxythiophene); a) rod-shaped CuO nanoparticles; b) flake-shaped CuO
nanoparticles
New trends in the electrochemical sensing of dopamine
Table 6 Analytical data for
electrooxidation of dopamine on
electrodes modified with nanoscale materials (simultaneous
detection)
Electrode
Method Compound Linearity (molL−1)
Chit/GR/GC
DPV
MWNT/IL/GC
DPV
CILE/CP
DPV
AuNP-PANI/GC
DPV
PdNC/PMPy/Pt
DPV
CuNP/PPy/GC
Chit, chitosan; GR, graphene;
CP, carbon paste; CNF, carbon
nanofibers; MWNT, multiwalled carbon nano tubes; IL,
ionic liquid; CILE, carbon ionic
liquid electrode; NP, nano particles; NC, nanoclusters;
HBNBH, 2,2′-[1,7-hepthandiylbis(nitriloethylidine))-bis-hydroquinone; PANI, polyaniline;
Ppy, polypyrrole; PMPy, poly
(N-methylpyrrole); Pfu, polyfuran; MB, methylene blue; SZP,
silica/zirkonia/phosphate
3761
DPV
HBNBH/ TiO2NP/ CP SWV
PdNC/PFu/Pt
DPV
(MB)SZP/C
DPV
PdNP/CNF/C
DPV
and UA the best results were obtained by use of electrodes
containing poly(N-methylpyrrole) and Pd clusters [112]. Interesting results were also obtained for simultaneous detection
of DA, UA, and AC on a PdNP/PFu/Pt electrode [115]. It is,
however, very difficult to compare results obtained with nanoparticles and clusters because the amount used is usually
uncontrolled and, therefore, unknown. There are also insufficient data to enable comparison of kinetic data, for example
electron-transfer rate constant for DA electrooxidation. Some
data are given for nanoparticles of CuO/CP and Fe2O3SWCNT/PGE electrodes [107, 109], and it was pointed out
that particle shape (CuO) and electrode composition (Fe2O3SWCNT) affect ks values. What can be compared for NPmodified electrodes, however, is separation of anodic peaks
for the AA–DA and DA–UA pairs. The best separation results
were obtained by use of Pd nanoparticle-modified electrodes
—PdNP/CNF (0.244 V, 0.148 V) and PdNC/PMPy/Pt
(0.205 V, 0.201 V) [112, 117]. The reproducibility and stability of such electrodes has also been monitored by several
workers. RSD, when determined, was found to be in the range
2–5 %. The electrodes retained 90–92 % of their initial response after storage for 7–14 days [99, 105, 112, 113]. The
electrodes were tested with real biological samples (human
Sensitivity
Detection limit Ref.
(mAmol−1 L) (molL−1)
DA
UA
AA
DA
SE
DA
UA
AA
DA
UA
DA
UA
1×10−6–2.4×10−5
2×10−6–4.5×10−5
5×10−5–1.2×10−3
1×10−7–1.2×10−5
2×10−8–7×10−6
2×10−6–1.5×10−3
2×10−6–2.2×10−4
5×10−5–7.4×10−3
7×10−6–1.4×10−4
2.9×10−5–7.2×10−4
1×10−7–1×10−5
710
5×10−7–2×10−5
280
AA
DA
UA
DA
UA
DA
AA
ACOP
DA
UA
AA
DA
UA
AA
5×10−5–1×10−3
1×10−9–1×10−7
1×10−9–1×10−5
8×10−6–1.4×10−3
1×10−4–6×10−4
5×10−7–1×10−4
5×10−5–1×10−3
5×10−7–1×10−4
6×10−6–1×10−4
2.2×10−5–3.5×10−4
1×10−4–1.6×10−3
5×10−7–1.6×10−4
2×10−6–2×10−4
5×10−5–4×10−3
5.6
43.9
42
478.4
21.3
263.7
[90]
6×10−8
8×10−9
1×10−6
1×10−6
2×10−5
3×10−6
2×10−5
1.2×10−8
2.7×10−8
7×10−6
8.5×10−10
8×10−10
8.4×10−7
4.82×10−8
7.13×10−6
7.64×10−8
1.7×10−6
3.7×10−6
8.3×10−6
2×10−7
7×10−7
1.5×10−5
[97]
[98]
[111]
[112]
[113]
[114]
[115]
[116]
[117]
urine or human serum) [102, 112, 113, 117] and with pharmacological dopamine injections [99, 104, 108, 109, 114].
Taking all these data into account it is clearly apparent
that the best results for selective, simultaneous determination of DA can be obtained by use of electrodes modified
with conducting polymers and nanoclusters, and/or nanoparticles of selected metals (Pd, Cu). However, the properties of the surface active layer should be optimized. In our
opinion, there is no need to prepare sophisticated composites or materials that are not carefully chosen and properly
characterized; such an approach will not improve sensor
performance and is more expensive.
Finally, comparison of biosensors with chemical sensors
reveals it is possible to obtain comparable detection limits.
However, the advantage of chemical sensors is the possibility of simultaneous detection of DA, AA, and UA. Most
result also show that the long-term stability of chemical
sensors is higher than that of biosensors. However, their
stability also decreases over longer time periods, most probably as a result of adsorption of oxidation products, formed
during successive measurements, on the electrode surface.
Investigation of surface properties after such measurements
has not been conducted.
3762
In-vivo and in-vitro sensing—miniaturization
Use of the sensors and biosensors described above for real time
in vivo and/or in vitro monitoring of clinically relevant physiological analytes, for example dopamine and other neurotransmitters has, with few exceptions, been restricted to laboratory
use. This is primarily because they suffer from poor selectivity
and sensitivity when used in a biological or biomimicking
environment. Other factors that restrict their application include:
1. most of the sensors developed in laboratories on the
bench are too large to be used for implantation because
of extensive tissue or cell damage; the same restriction
applies to in-vitro extracellular measurements;
2. bio-incompatibility of materials used in their development;
3. long term stability;
4. frequent fouling compromising their sensitivity; and/or
5. concern about selectivity to the analyte of interest.
Recent developments in nano and microtechnology have
enabled the manufacture of electrochemical sensors
(electrodes) in the range 1–30 μm in diameter. These are now
available with different geometry (disc, rod, band, etc.) and
materials (C, Au, Pt, Ag) [120, 121] and are expected to take
full advantage of the results described in the previous sections
of this brief review. With use of appropriate electrochemical
techniques (e.g., fast scan cyclic voltammetry, FSCV), microelectrochemistry has enabled the real-time, in-vivo and/or invitro measurements with high spatial and temporal resolution
of fluctuations of the neurotransmitter concentration because of
its transient release and uptake by living cells. The advantages
of miniaturized electrodes (whether implantable or not) for
electrochemical sensing of redox neurotransmitters, for example dopamine, can be briefly summarized as follows:
1. they have improved signal-to-noise ratios, because analytical Faraday currents are substantially increased by
the higher rates of mass transfer of hemispherical diffusion of the electroactive compounds;
2. their response times are much faster because of the
small double-layer capacitance charging currents and
time constants, resulting in the capability of high temporal resolution of neurotransmitter fluctuations; and
3. the so-called iR drop is of less concern because the total
analytical currents measured by such electrodes are much
smaller than those measured with typical large-scale
electrodes.
Nevertheless, for such a small dimensions, the major
issue, relatively easily resolved for laboratory bench sensors, is the design and fabrication of the sensor–solution
interface and its effect on sensitivity and selectivity toward
the target dopamine and interfering agents.
K. Jackowska, P. Krysinski
Relatively few papers have addressed design of the sensor–
solution interface for miniaturized implantable electrodes;
most research has focused on development of the electrochemical methods. However, for clarity it is necessary to state
that not all state-of-the-art electrochemical techniques, for
example scanning electrochemical microscopy, SECM, can
be applied to in-vivo measurements for awake, moving animals, for obvious “geometric” reasons. Until now SECM has
been restricted to cell cultures or (at most) to small anaesthetized animals [122, 147]. Nevertheless, use of micro or ultramicroelectrodes (UMEs) and microfabricated electrode arrays
(MEAs) is not restricted by any particular electrochemical
technique and can be used both in vivo and in vitro. Here,
we will summarize the state-of-the-art of electrochemical
methods used for in-vivo detection of dopamine, mostly for
awake, mobile animals. This will be followed by an overview
of surface modification of implantable electrodes to improve
biocompatibility and selectivity for dopamine, with the proviso that although such modifications can be effective at reducing interferences, they may also reduce the efficiency of the
electron transfer kinetics, reducing sensor sensitivity.
Electrochemical techniques for detection of dopamine
with implantable electrodes
Since the early work of Adams and colleagues [123, 124]
which introduced electrochemistry to the neurosciences, numerous electrochemical techniques and electrode materials
have been used to identify and resolve catecholamines. In direct
electrochemical detection of in vivo and in vitro dopamine,
potentiostats with a three or two-electrode configurations have
been used. The two-electrode configuration, consisting of a
working electrode (microelectrode or UME) and a reference
electrode, is usually preferred, because the measured currents
are sufficiently small to preclude polarization of the reference
electrode at ca 150 mmolL−1 chloride concentrations in physiological electrolytes. The reference electrode is typically a
micrometer diameter silver wire coated with a silver chloride
layer, positioned next to the working electrode. The techniques
most commonly used for direct detection of dopamine (or other
electroactive neurotransmitters) are constant-potential amperometry (DC amperometry), differential-pulse voltammetry
(DPV), and fast-scan cyclic voltammetry (FSCV), the last
being a so-called “dynamic” technique.
In DC amperometry, a constant potential is applied which
is sufficient to oxidize dopamine (or reduce dopaquinone)
and the current, related to the amount of dopamine by Faraday’s law, is recorded as a function of time. With current
sampling rates in the kHz range, this technique can resolve
signals on time scales below milliseconds. This technique
has been successfully used for studies of catecholamine
concentrations in the brain and in brain slices [125, 126],
New trends in the electrochemical sensing of dopamine
exocytosis of the small synaptic vesicles [127], neuroblastoma, and other cells [128, 129]; it has the best temporal
resolution because of sampling rates down to 1 ms. However, the disadvantage of DC amperometry is that it is
essentially nonselective, because all electroactive compounds that oxidize (or reduce) at the applied potential will
produce a faradaic response at the electrode. Moreover,
much amplification is required so the technique is susceptible to noise artifacts arising as a result of animal movement.
In a variant of DC amperometry, chronoamperometry, the
potential is stepped from its initial value, where no redox
reaction occurs, toward a potential at which an oxidation or
reduction of the analyte molecules proceeds. Then, after a
given holding time, the potential is stepped back to its initial
value. This variant has been applied to real-time measurement of dopamine concentrations in the cerebral fluid of the
brain [130, 131]. Yet again, it suffers from limited chemical
selectivity. Nevertheless, both techniques can be successfully used to measure dopamine dynamics evoked by an electric stimulus (for example deep-brain stimulus, DBS) [132]
or reward-induced firing of dopaminergic neurons [133].
Differential pulse voltammetry (DPV) is more sensitive and
selective. It uses square potential pulses of constant height
superimposed over a linearly increasing potential from the initial
to a final value. The potential pulse of constant frequency is
typically ca 25 mV. Response currents are sampled before the
potential pulse and at the end of the potential pulse. The difference between these current responses is plotted against the
potential of the linear ramp. Because DPV is a differential
technique, the response is a current peak of amplitude proportional to analyte (dopamine) concentration and its position
shows the half-wave potential (for reversible systems). By excluding the first few milliseconds of each pulse (usually lasting
tens of milliseconds), the effect of the capacitative current arising
as a result of charging and/or discharging of the electrical double
layer at the electrode interface is typically minimized. Although
differential pulse voltammetry is more sensitive than the previous techniques [133–135], it has limited time resolution, because one full scan can take longer than seconds.
Finally, perhaps the most frequently used technique for in
vivo detection of neurotransmitter trafficking is fast-scan cyclic
voltammetry (FSCV). This technique provides high temporal
resolution without compromising very good selectivity. Since
its early application by Whitman’s group [136, 137], this
technique has proved effective for multi-analyte detection of
catecholamines with chemical selectivity [2]. In FSCV, the
voltage applied to the microelectrode is rapidly cycled in a
triangular fashion at a rate higher than 100 Vs−1. The limits of
the triangular wave are chosen so that oxidation and reduction
of, e.g., dopamine, lie within this potential window and the
current (both faradaic and capacitive) resulting from the process is monitored. FSCV is usually performed with bare electrodes, for which rates of electron transfer are much faster than
3763
for modified electrodes. When used with appropriate ultramicroelectrodes (UMEs), FSCV can provide high spatial resolution [2, 138] and time resolution. This fast scanning limits
diffusion distance to the electrode and minimizes electrode
fouling [20, 139, 140]. To eliminate the capacitive contribution
that scales linearly with the scan rate (high-frequency cycling),
careful and precise background subtraction is necessary to
eliminate the signal from capacitance build-up at the electrode
interface [139]. This is usually achieved by analog background
subtraction, use of the Hilbert transform, or by principalcomponents analysis; it is beyond the scope of this work to
analyze the advantages and drawbacks of each method. The
interested reader is referred to Refs. [20, 139, 140] and references cited therein for more detailed description.
The experimental techniques and approach used for in-vivo
experiments with awake Japanese monkeys (Macaca fuscata)
during cue–reward trials are summarized in Figs. 3 and 4 [133].
Lee et al. recently introduced pair–pulse voltammetry (PPV)
in which a selected binary waveform with a defined time lapse
is applied to an electrode which is held at a negative potential
between each of two pulses [141]. This enables recording of
two simultaneous, yet different, voltammograms, and enables
discrimination of analytes on the basis of their adsorption
behavior on the electrode [141]. More detailed information
about electrochemical techniques as bioanalytical tools for invivo and in-vitro experiments can be found in several excellent
recent reviews [142–148]. Here we briefly summarize the three
techniques most commonly used for monitoring catecholamines in vivo by use of implanted (ultra)microelectrodes.
Electrode materials, surface functionalization and coatings
Electrode materials and the components of electrodes are
selected carefully to improve sensor sensitivity, selectivity,
reproducibility, robustness, and long-term stability and/or
resistance to fouling in the biological environment. As already stated, a variety of noble metals have been used for
microelectrode fabrication, because of the relative ease
preparation and subsequent modifications, yet most research
on dopamine release and uptake in vivo cited in this review
relies on fiber microelectrodes fabricated from carbon, because of the apparent biocompatibility of this material.
However, irrespective of electrode material, relatively few
papers address the design of the sensor–solution interface
for miniaturized implantable electrodes. Below we discuss
some recent approaches to this problem.
Carbon and carbon-derived ultramicroelectrodes
Carbon fiber (CF) microelectrodes are the basis of a vast
amount of experimental work in neuroscience. Electrochemical characterization of dopamine oxidation on these
3764
K. Jackowska, P. Krysinski
Fig. 3 Applied waveforms and current recordings for fast-scan cyclic
voltammetry (FSCV; top row), square-wave differential pulse voltammetry (SW; middle row), and constant-potential amperometry (bottom
row). The left column shows the potential applied to the Ag/AgCl
reference electrode. The middle column shows the actual current
measured for each waveform at the carbon fiber. For square-wave,
only one pulse is indicated. The right column shows the current
measured at the diamond microelectrode. Typical recorded currents
in PBS (black) and in high (10 and 100 nmolL−1) concentrations of
dopamine (green). (Adapted, with permission, from Ref. [133])
electrodes has been well described and optimized [2, 140,
146]. These microelectrodes have been shown to be relatively
unsusceptible to fouling by products of dopamine electrooxidation and much superior to other types of material for invivo electrochemical experiments [120]. It has been suggested
that electrochemical detection of dopamine and other cationic
neurotransmitters depends on their adsorption by the CF microelectrode. Therefore, to improve sensitivity and selectivity,
carbon fiber microelectrodes have undergone many treatments, for example overoxidation to develop the surface area
and/or to control the surface chemistry [149–152]. Other
approaches include modification of CF electrodes with carbon
nanotubes, either adsorbed [152, 153] or self-assembled on
the surface of a functionalized microelectrode [154]. Also, by
use of appropriate surface polar functional groups to modify
the CNTs used on CNT-modified carbon-based electrodes, the
New trends in the electrochemical sensing of dopamine
3765
Fig. 4 Experimental methods. (a) Side view of the tips of the carbon
fiber (Cf) and diamond microelectrode (BDD). Scale bar0500 μm. (b)
Scanning electron microgram of the tip of the diamond microelectrode.
(c) Circuit of potentiostat. (d) Electrode positions for recordings in
mouse brains. (e) Left: elongated diamond microelectrode and guide
cannula. Right: microelectrode positions on the monkey head. POT,
potentiostat. (Adapted, with permission, from Ref. [133])
electron-transfer kinetics have been adjusted to enhance the
sensitivity to cationic neurotransmitters (dopamine and serotonin). However, an unexpected result was obtained for negatively charged CNTs (carboxyl groups)—an equivalent
signal increase for negatively charged ascorbic acid, a well
known interferent, was also observed, suggesting that for this
particular molecule, electrostatic interactions are of secondary
importance [152]. A procedure for renewing the surface of a
carbon microelectrode, by regenerating its electrochemically
active surface while simultaneously reducing or removing
fouling problems and improving electron-transfer kinetics,
has recently been reported [155].
To summarize, use of carbon fiber microelectrodes seems
to enable control of surface functional groups to satisfy most
of the criteria for ideal implantable electrodes for dopamine
detection: they are easily and reproducibly fabricated and
small enough to be located in specific areas with minimal
tissue damage. It has also been suggested that carbon electrodes modified with graphene perform better than multiwalled carbon nanotube-modified electrodes in selective
detection of DA because of π–π stacking interactions between dopamine and the graphene surface. Apparently,
these interactions accelerate electron transfer for DA but
hinder oxidation of ascorbic acid on the graphenemodified electrode, completely eliminating the interfering
signal from ascorbic acid [89]. Another material recently
considered for in-vivo electrochemical detection of dopamine is boron doped diamond (BDD). BDD is known to be
a robust, durable, biocompatible electrode material with low
charging (capacitive) background currents and a wide range
of potential [156]. It is also highly resistant to biological
fouling of its surface [133, 147]. It has been found that
extensive anodic polarization before in-vivo use improved
the selectivity of such electrodes for dopamine, most probably because of the formation of a carboxyl or hydroxylterminated surface. Such electrodes were found to be less
sensitive to dopamine than CF electrodes, however.
Metal electrodes
Metal electrodes, usually Pt and Au, enable both versatile
surface modification and microfabrication of electrode
arrays for multi-analyte monitoring and multiplexing, particularly for “lab-on-a-chip” (LOC) monitoring, differentiation, and reporting of catecholamines [121, 157–160]. Gold
has the beneficial properties of adsorbing and preconcentrating catecholamines on its surface, thus increasing the sensitivity of gold electrodes to these compounds. The gold surface
can, moreover, be easily modified by covalent bonding with
thiol groups, enabling formation of self-assembled monolayers (SAMs) with different functionality facing the external
solution [161]. Miniature gold electrodes prepared in this way
have been used to characterize a variety of neurotransmitters
(including dopamine) and interfering agents, and the results
have been compared with analogous results obtained by use of
carbon fiber microelectrodes [120].
Miniaturized enzymatic electrodes; biosensors
As already mentioned, there are relatively few reports of
fabrication and use of miniaturized enzymatic biosensors for
in-vivo experiments. Early reports describe ceramic-based
lithographically prepared electrode arrays for electrochemical measurement of neurotransmitters [162]. These arrays
were coated with glutamate oxidase crosslinked with bovine
serum albumin via glutaraldehyde treatment. Their
3766
voltammetric performance was characterized in the presence
of dopamine and hydrogen peroxide yet they were not used
for in-vivo dopamine measurements. A polyphenol oxidase
has also been used to prepare electrochemical microsensors
for detection of dopamine and glutamate [74]. However, this
enzyme was not selective for dopamine. Catalytic complexes that mimic the active site of catecholases have also
been used in sensors of phenolic substrates [163–165].
Despite the innovative technology used in the nanosensing devices discussed above, as far as we are aware only one
in-vivo enzymatic microbiosensor for dopamine detection
has been reported [55]. This biosensor was fabricated by
immobilizing tyrosinase in a matrix consisting of chitosan
and ceria-based metal oxides (CeO2 and TiO2) deposited on
a carbon-fiber microelectrode. The chitosan coating resulted
in a biocompatible sensor and, after optimization, the biosensor was reported to have an extremely large linear range,
between 10 nmolL−1 and 220 μmolL−1 dopamine, and
sensitivity of ca 14 nAμmol−1 L. Sensor selectivity against
ascorbic acid (AA), uric acid (UA), serotonin, norepinephrine, epinephrine, and L-DOPA was also tested. Lack of
responsive to AA and UA was attributed to the low operating potential of the sensor, and insensitivity/low sensitivity
to serotonin, L-DOPA, norepinephrine, and epinephrine was
attributed to the bulkier structure of these compounds,
which cannot reach the active site of tyrosinase [55].
Conclusions and perspectives
In this brief review we have discussed recent achievements in
the development of dopamine sensing. Many advances have
been made in the construction of sensors in laboratories, on
the bench, leading to a wealth of information on the production of DA electrochemical sensors with detection limits in the
picomolar range, selective and specific toward DA, yet easy to
operate and reusable or single-use but cost-effective. The
biosensors described above are based on tyrosinase attached
to different types of immobilizing and signal-transducing
substrates, which enable electroanalytical detection. However,
the analytical performance of some of these biosensors, for
example low detection limit and low linear range of sensor
response, clearly indicate that, in the presence of dominant
interferents, determination of dopamine at the nanomolar levels characteristic of living system is rarely possible. The
situation is very similar for the chemical sensors used for
dopamine detection. Many new materials, for example carbon
materials (carbon fibers, carbon nanotubes, graphene), nanoparticles, clusters (Au, Ag, Pd, Cu, semiconducting oxides,
magnetic oxides, among others), conducting polymers, and
ionic liquids, have been used for sensor construction to improve their performance. However, the desired nanomolar
limit of detection has been obtained in a few cases only, when
K. Jackowska, P. Krysinski
electrodes modified with conducting polymers and metal (Pd,
Cu) nanoclusters and/or nanoparticles have been used for
selective, simultaneous determination of DA in the presence
of a variety of interfering agents.
Rapid progress in the use of these new materials for
sensor construction has not been accompanied by appropriate characterization of the sensors. For biosensors, tyrosinase activity is very rarely determined after its entrapment in
the matrix (in the bulk or attached to the surface). Enzyme
loading also is usually unknown and uncontrolled. The
situation is similar for chemical sensors also—surface properties and nanoparticle load are rarely controlled.
To our surprise, few kinetic data have been reported for
enzymatic reactions of tyrosinase in the presence of dopamine, so it is impossible to compare kinetic data for DA
electrooxidation, for example electron-transfer rate constant,
or to relate these to the surface properties and analytical
performance of such sensors.
To summarize, proper characterization of sensors, i.e.
determination of kinetics and analytical performance, then
correlation of these with material properties is more important than relentless application of increasingly sophisticated
composites or new materials to sensor construction.
Currently available methodology for miniaturization
could take advantage of developments in real-time mapping
of neurotransmitter release zones on cell surfaces, with high
spatial and temporal resolution both in vitro and in vivo,
e.g., for monitoring of dopamine synaptic trafficking. Although technology, including sensor arrays, is available for
combined monitoring of multiple catecholamines, design of
sensor–solution interfaces for determination of the selectivity and sensitivity of sensors awaits further improvement.
Acknowledgments This work was supported by Project PSPB-079/
2010 supported by a grant from Switzerland through the Swiss Contribution to the enlarged European Union.
Open Access This article is distributed under the terms of the Creative
Commons Attribution License which permits any use, distribution, and
reproduction in any medium, provided the original author(s) and the
source are credited.
References
1. Mora F, Segovia G, del Arco A, de Blas M, Garrido P (2012)
Stress, neurotransmitters, corticosterone and body-brain integration. Brain Res. doi:10.1016/j.brainres.2011.12.049
2. Robinson DL, Venton BJ, Heien MLAV, Wightman RM (2003)
Detecting subsecond dopamine release with fast-scan cyclic voltammetry in vivo. Clin Chem 49:1763–1773
3. Hefco V, Yamada K, Hefco A, Hrticu L, Tiron A, Nabeshima T
(2003) Role of the mesotelencephalic dopamine system in learning and memory processes in the rat. Eur J Pharmacol 475:55–60
4. Swamy BEK, Venton BJ (2007) Carbon nanotube-modified
microelectrodes for simultaneous detection of dopamine and
serotonin in vivo. Analyst 132:876–884
New trends in the electrochemical sensing of dopamine
5. Lupu S, Lete C, Marin M, Totir N, Balaure PC (2009)
Electrochemical sensors based on platinum electrodes modified with hybrid inorganic–organic coatings for determination
of 4-nitrophenol and dopamine. Electrochim Acta 54:1932–
1938
6. Zhao Y, Gao Y, Zhan D, Liu H, Zhao Q, Kou Y, Shao Y, Li M,
Zhuang Q, Zhu Z (2005) Selective detection of dopamine in the
presence of ascorbic acid and uric acid by a carbon nanotubesionic liquid gel modified electrode. Talanta 66:51–57
7. Chernyshov DV, Shvedene NV, Antiova ER, Pletnev IV (2008)
Ionic liquid-based miniature electrochemical sensors for the voltammetric determination of catecholamines. Anal Chim Acta
621:178–184
8. Galvan A, Wichmann T (2008) Pathophysiology of parkinsonism. Clin Neurophysiol 119:1459–1474
9. Swanson JM, Volkov ND (2009) Psychopharmacology: concepts
and opinions about the use of stimulant medications. J Child
Psychol Psychiatry 50:180–193
10. Kollins SH, March JS (2007) Advances in the pharmacotherapy of
attention-deficit/hyperactivity disorder. Biol Psychiatry 62:951–953
11. Dalley JW, Roiser JP (2012) Dopamine, serotonin and impulsivity. Neuroscience 215:42–58
12. Smeets WJ, Gonzalez A (2000) Catecholamine systems in the
brain of vertebrates: new persectives through a comparative approach. Brain Res Brain Res Rev 33:308–379
13. Miller MG (2011) The emerging role of trace amine-associated
receptor 1 in the functional regulation of monoamine transporters
and dopaminergic activity. J Neurochem 116:164–176
14. Neve KA, Seamans JK, Trantham-Davidson H (2004) Dopamine
receptor signaling. J Recept Signal Transduct 24:165–205
15. Neve KA, Jeremy K, Seamans JK, Trantham-Davidson H (2004)
Dopamine receptor signaling. J Recept Signal Transduct 24:165–205
16. Wightman RM, May LJ, Michael AC (1988) Detection of dopamine dynamics. Anal Chem 80:769A–776A
17. Sofuoglu M, Sewell RA (2009) Norepinephrine and stimulant
addiction. Addict Biol 14:119–129
18. Shou M, Ferrario CR, Schultz KN, Robinson TE, Kennedy RT
(2006) Monitoring dopamine in vivo by microdialysis sampling
and on-line CE-laser induced fluorescence. Anal Chem 78:6717–
6725
19. Wang D, Zhu W, An Y, Zheng J, Zhang W, Jin L, Gao H, Lin L
(2008) LC with novel electrochemical detection for analysis of
monoamine neurotransmitters in rat brain after administration of
(R)-salsolinol and (R)-N-methylsalsolinol. Chromatographia
67:369–374
20. Perry M, Li Q, Kennedy RT (2009) Review of recent advances in
analytical techniques for the determination of neurotransmitters.
Anal Chim Acta 653:1–22
21. Hubbard AT, Stickney JI, Soriaga MP, Chia VKF, Rosasco SD,
Schardt BS, Solomun T, Song D, White JH, Wieckowski A
(1984) Electrochemical processes at well-defined surfaces. J
Electroanal Chem 168:43–66
22. Salaita GN, Laguren-Davidson L, Lu F, Walton N, Wellner
E, Stern DA, Batina N, Frank DG, Lin C-H, Benton CS,
Hubbard AT (1988) Electrochemical reactivity of 2,2′,5,5′tetrahydroxybiphenyl and related compounds adsorbed at Pt
(111) surfaces; studies by EELS, LEED, Auger spectroscopy
and cyclic voltammetry. J Electroanal Chem Interf
Electrochem 245:253–273
23. Tenreiro AM, Nabais C, Correia JP, Fernandes FMSS, Romero
JR, Abrantes LM (2007) Progress in the understanding of tyramine electropolymerisation mechanism. J Solid State
Electrochem 11:1059–1069
24. Cooper SE, Venton BJ (2009) Fast-scan cyclic voltammetry for
the detection of tyramine and octopamine. Anal Bioanal Chem
394:329–336
3767
25. Sanchez-Cortes S, Francioso O, Garcia-Ramos JV, Ciavatta C,
Gessa C (2001) Catechol polymerization in the presence of silver
surface. Colloids Surf A 176:177–184
26. Majewska UE, Chmurski K, Biesaga K, Olszyna AR, Bilewicz R
(2006) Dopamine oxidation at per(6-deoxy-6-thio)-α-cyclodextrin
monolayer modified gold electrodes. Electroanalysis 18:1463–1470
27. Ernst H, Knoll M (2001) Electrochemical characterisation of uric
acid and ascorbic acid at a platinum electrode. Anal Chim Acta
449:129–134
28. Zhu L, Tian C, Zhai J, Yang R (2007) Sol–gel derived carbon
nanotubes ceramic composite electrodes for electrochemical sensing. Sensors Actuators B Chem 125:254–261
29. Huffman ML, Venton BJ (2008) Electrochemical properties of
different carbon-fiber microelectrodes using fast-scan cyclic voltammetry. Electroanalysis 20:2422–2428
30. Fujishima A, Rao TN, Popa E, Sarada BV, Yagi I, Tryk DA
(1999) Electroanalysis of dopamine and NADH at conductive
diamond electrodes. J Electroanal Chem 473:179–185
31. Roy PR, Saha MS, Okajima T, Park S-G, Fujishima A, Ohsaka T
(2004) Selective detection of dopamine and its metabolite, DOPAC,
in the presence of ascorbic acid using diamond electrode modified
by the polymer film. Electroanalysis 16:1777–1784
32. Barnes EO, O’Mahony AM, Aldous L, Hardacre C, Compton RG
(2010) The electrochemical oxidation of catechol and dopamine
on platinum in 1-ethyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide ([C2mim][NTf2]) and 1-butyl-3-methylimidazolium tetrafluoroborate ([C4mim][BF4]): adsorption effects in
ionic liquid voltammetry. J Electroanal Chem 646:11–17
33. Raj CR, Okajima T, Ohsaka T (2003) Gold nanoparticle arrays for
the voltammetric sensing of dopamine. J Electroanal Chem
543:127–133
34. Cabrita JF, Abrantes LM, Viana AS (2005) N-Hydroxysuccinimideterminated self-assembled monolayers on gold for biomolecules
immobilization. Electrochim Acta 50:2117–2124
35. Zhao H, Zhang YZ, Yuan ZB (2002) Determination of dopamine in the presence of ascorbic acid using poly(hippuric
acid) modified glassy carbon electrode. Electroanalysis
14:1031–1034
36. Liu A, Wei M, Honma I, Zhou H (2006) Biosensing properties of
titanate nanotube films: selective detection of dopamine in the
presence of ascorbate and uric acid. Adv Funct Mater 16:371–376
37. Yan W, Feng X, Chen X, Li X, Zhu J-J (2008) A selective
dopamine biosensor based on AgCl@polyaniline core–shell
nanocomposites. Bioelectrochemistry 72:21–27
38. Lange U, Roznyatouskaya NV, Mirsky VM (2008)
Conducting polymers in chemical sensors and arrays. Anal
Chim Acta 8:1–26
39. Zheng M, Zhou Y, Chen Y, Tang Y, Lu T (2010) Electrochemical
behavior of dopamine in the presence of phosphonate and the
determination of dopamine at phosphonate modified zirconia
films electrode with highly antifouling capability. Electrochim
Acta 55:4789–4798
40. Tyagi P, Postetter D, Saragnese DL, Randall CL, Mirski MA,
Gracias DH (2009) Patternable nanowire sensors for electrochemical recording of dopamine. Anal Chem 81:9979–9984
41. Oztekin Y, Tok M, Bilici E, Mikoliunaite L, Yazicigil Z,
Ramanaviciene A, Ramanavicius A (2012) Copper nanoparticle
modified carbon electrode for determination of dopamine.
Electrochim Acta 76:201–207
42. Ly SY (2006) Detection of dopamine in the pharmacy with a carbon
nanotube paste electrode using voltammetry. Bioelectrochemistry
68:227–231
43. Jeong H, Jeon S (2008) Determination of dopamine in the presence of ascorbic acid by nafion and single-walled carbon nanotube film modified on carbon fiber microelectrode. Sensors
8:6924–6935
3768
44. Li NB, Ren W, Luo HQ (2008) Simultaneous voltammetric
measurement of ascorbic acid and dopamine on poly(caffeic
acid)-modified glassy carbon electrode. J Solid State
Electrochem 12:693–699
45. Kumar SS, Mathiyarasu J, Phani KLN, Yegnaraman V (2006)
Simultaneous determination of dopamine and ascorbic acid on
poly (3,4-ethylenedioxythiophene) modified glassy carbon electrode. J Solid State Electrochem 10:905–913
46. Balamurugan A, Chen SM (2007) Poly(3,4-ethylenedioxythiophene-co-(5-amino-2-naphthalenesulfonic acid)) (PEDOTPANS) film modified glassy carbon electrode for selective detection of dopamine in the presence of ascorbic acid and uric acid.
Anal Chim Acta 596:92–98
47. Vasantha VS, Chen SM (2006) Electrocatalysis and simultaneous
detection of dopamine and ascorbic acid using poly(3,4-ethylenedioxy)thiophene film modified electrodes. J Electroanal Chem
592:77–87
48. Lin X, Zhang Y, Chen W, Wu P (2007) Electrocatalytic oxidation
and determination of dopamine in the presence of ascorbic acid
and uric acid at a poly (p-nitrobenzenazo resorcinol) modified
glassy carbon electrode. Sens Actuators B 122:309–314
49. Maciejewska J, Pisarek K, Bartosiewicz I, Krysinski P,
Jackowska K, Biegunski AT (2011) Selective detection of dopamine on poly(indole-5-carboxylic acid)/tyrosinase electrode.
Electrochim Acta 56:3700–3706
50. Zhou YL, Tian RH, Zhi JF (2007) Amperometric biosensor based
on tyrosinase immobilized on a boron-doped diamond electrode.
Biosens Bioelectron 22:822–828
51. Yu-Chen T, Chiu C-C (2007) Amperometric biosemsors based on
multiwalled carbon nanotube-Nafion-tyrosinase nanobiocomposites for the determination of phenolic compounds. Sens
Actuators B 125:10–16
52. Tembe S, Kubal BS, Karve M, Souza SF (2008) Glutaradelhyde
activated eggshell membrane for immobilization of tyrosinase from
Amorphophallus companulatus: application in construction of electrochemical biosensor for dopamine. Anal Chim Acta 612:212–217
53. Njagi J, Ispas C, Andreescu S (2008) Mixed ceria-based metal
oxides biosensor for operation in oxygen restictive enviroments.
Anal Chem 80:7266–7244
54. Min K, Yoo YJ (2009) Amperometric detection of dopamine based on
tyrosinase-SWNTs-Ppy composite electrode. Talanta 80:1007–1011
55. Njagi J, Chernov MM, Leiter JC, Andreescu S (2010)
Amperometric detection of dopamine in vivo with an enzyme
based carbon fiber microbiosensor. Anal Chem 82:989–996
56. Thevenot DR, Toth K, Durst RA, Wilson GS (2001)
Electrochemical biosensors: recommended definitions and classification. Biosens Bioelectron 16:121–131
57. Pumera M, Sanchez S, Ichinose I, Tang J (2007) Electrochemical
nanobiosensors. Sens Actuators B 123:1195–1205
58. Yogeswaran U, Chen S-M (2008) A review on the electrochemical sensors and biosensors composed of nanowires as sensing
materials. Sensors 8:290–313
59. Wang J, Lin Y (2008) Functionalized carbon nanotubes and
nanofibers for biosensing applications. Trends Anal Chem
27:619–626
60. Balasubramanian K, Burghard M (2006) Biosensors based on
carbon nanotubes. Anal Bioanal Chem 385:452–468
61. Lu Y, Yang M, Qu F, Shen G, Yu R (2007) Enzymefunctionalized gold nanowires for the fabrication of biosensors.
Bioelectrochemistry 71:211–216
62. Pingarron JM, Yanez-Sadeno P, Gonzales-Cortes A (2008) Gold
nanoparticle-based electrochemical biosensors. Electrochim Acta
53:5848–5866
63. Xu Z, Chen X, Dong S (2006) Electrochemical biosensors based
on advanced bioimmobilization matrices. Trends Anal Chem
9:899–907
K. Jackowska, P. Krysinski
64. Solomon EI, Sundaram UM, Machonkin TE (1996) Multicopper
oxidases and oxygenases. Chem Rev 96:2563–2605
65. Shleev SS, Tkac J, Christenson A, Ruzgas T, Yaropolov AI,
Whittaker JW, Gorton L (2005) Direct electron transfer between
copper-containing proteins and electrodes. Biosens Bioelectron
20:2517–2554
66. Cosnier S (1999) Biomolecule immobilization on electrode surfaces by entrapment or attachment to electrochemically polymerized films. A review. Biosens Bioelectron 14:443–456
67. Gerard M, Chaubey A, Malhorta RD (2002) Application of conducting polymers to biosensors. Biosens Bioelectron 17:345–359
68. Cosnier S (2003) Biosensors based on electropolymerized films:
new trends. Anal Bioanal Chem 377:507–520
69. Ahuja T, Mir IA, Kumar D, Rajesh (2007) Biomolecular immobilization on conducting polymers for biosensing applications.
Biomaterials 28:791–805
70. Duran N, Rosa MA, D'Annibale A, Gianfreda L (2002) Application
of laccase and tyrosinases (phenoloxidases) immobilized on different supports: a review. Enzym Microb Technol 31:907–931
71. Cracknell JA, Vincent KA, Armstrong FA (2008) Enzymes as
working or inspirational electrocatalysts for fuel cells and electrolysis. Chem Rev 108:2439–2461
72. Forzani ES, Rivas GA, Solis VM (1995) Amperometric detection
of dopamine on an enzymatically modiefied carbon paste electrodes. J Electroanal Chem 382:33–40
73. Hasebe Y, Hirano T, Uchiyama S (1995) Determination of catecholamines and uric acid in biological fluids without preatreatment, using chemically amplified biosensors. Sens Actuators B
24–25:94–97
74. Cosnier S, Innocent C, Allien L, Poitry S, Tsacopoulos M (1997)
An electrochemical method for making enzyme microsensors.
Application of the detection to the detection of dopamine and
glutamate. Anal Chem 89:968–971
75. Pandey PC, Upadhay S, Tiwari I, Singh G, Tripathi VS (2001) A
novel ferrocene encapsulated palladium-linked ormosil-based
electrocatalytic dopamine biosensor. Sens Actuators B 75:48–55
76. Vedrine C, Fabiano S, Tran-Minh C (2003) Amperometric tyrosinase based biosensor using an electrogenerated polythiophene
film as an entrapment support. Talanta 59:535–544
77. Tembe S, Karve M, Inamdar S, Haram S, Melo J, D’Souza SF
(2006) Development of electrochemical biosensor based on tyrosinase immobilized in composite biopolimeric film. Anal
Biochem 349:72–77
78. Wang Y, Zhang X, Chen Y, Xu H, Tan Y, Wang S (2010)
Detection of dopamine based tyrosinase-Fe3O4 nanoparticleschitosan nanocomposite biosensor. Am J Biomed Sci 2:209–216
79. Henstridge MC, Dickinson EJF, Aslanoglu M, Batchelor-McAuley C,
Compton RG (2010) Voltammetric selectivity conferred by the modification of electrodes using conductive porous layers or films:
the oxidation of dopamine on glassy carbon electrodes modified with multiwalled carbon nanotubes. Sensors Actuators
B Chem 145:417–427
80. Pumera M, Ambrosi A, Bonnani A, Chng EL, Poh HL (2010)
Graphene for electrochemical sensing and biosensing. Trends
Anal Chem 29:954–965
81. Zhuang Z, Li J, Xu R, Xiao D (2011) Electrochemical detection
of dopamine in the presence of ascorbic acid using overoxidized
polypyrole/graphene modified electrode. Int J Electrochem Sci
6:2149–2161
82. Wang Z, Xia J, Zhu L, Chen X, Zhang F, Yao S, Li Y, Xia Y
(2011) A selective voltammetric method for detecting dopamine
at quercetin modified electrode incorporating graphene.
Electroanalysis 23:2463–2471
83. Zhu M, Zeng C, Ye J (2011) Graphene-modified carbon fiber
microelectrode for the detection of dopamine in mice hippocampus tissue. Electroanalysis 23:907–914
New trends in the electrochemical sensing of dopamine
84. Si P, Chen H, Kannan P, Kim D-H (2011) Selective and sensitive
determination of dopamine by composites of polypyrrole and
graphene modified electrodes. Analyst 136:5134–5138
85. Wu L, Feng L, Ren J, Qu X (2012) Electrochemical detection of
dopamine using porphyrin-functionalized graphene. Biosens
Bioelectron 34:57–62
86. Hon S, Kasner ML, Su S, Patel K, Cuellari R (2010) Highly
sensitive and selective dopamine biosensor fabricated with silanized graphene. J Phys Chem C 114:14915–14921
87. Kim Y-R, Bong S, Kang Y-J, Yang Y, Mahajan RK, Kim JS, Kim
H (2010) Electrochemical detection of dopamine in the presence
of ascorbic acid using graphene modified electrodes. Biosens
Bioelectron 25:2366–2369
88. Tan L, Zhou K-G, Zhang Y-H, Wang H-X, Wang X-D, Guo Y-F,
Zhang H-L (2010) Nanomolar detection of dopamine in the
presence of ascorbic acid at β-cyclodextrin/graphene nanocomposite platform. Electrochem Commun 12:557–560
89. Wang Y, Li Y, Tang L, Lu J, Li J (2009) Application of graphenemodified electrode for selective detection of dopamine.
Electrochem Commun 11:889–892
90. Han D, Han T, Shan C, Ivaska A, Niu L (2010) Simultaneous
determination of ascorbic acid, dopamine, uric acid with chitosangraphene modified electrode. Electroanalysis 22:2001–2008
91. Li Y, Liu X, Zeng X, Liu X, Kong B, Wei W, Luo S (2011)
Selective and sensitive detection of dopamine in the presence of
ascorbic acid by molecular sieve/ionic liquids composite electrode. Electrochim Acta 56:2730–2734
92. Li Y, Liu X, Wei W (2011) Square wave voltammetry for selective detection of dopamine using polyglycine modified carbon
ionic liquid electrode. Electroanalysis 23:2832–2838
93. Chang J-L, Wei G-T, Zen J-M (2011) Screen –printed ionic
liquid/preanodized carbon electrode: effective detection of dopamine in the presence of high concentration of ascorbic acid.
Electrochem Commun 13:174–177
94. Zhu Z, Qu L, Guo Y, Zeng Y, Sun W, Huang X (2011)
Electrochemical detection of dopamine on a Ni/Al layered double
hydroxide modified carbon ionic liquid electrode. Sens Actuators B
151:146–152
95. Dong J, Hu Y, Zhu S, Xu J, Xu Y (2010) A highly selective and
sensitive dopamine and uric acid biosensor fabricated with functionalized ordered mesoporous carbon and hydrophobic ionic
liquid. Anal Bioanal Chem 396:1755–1762
96. Wai G, Ivaska A (2008) Application of ionic liquids in electrochemical sensors. Anal Chim Acta 607:126–135
97. Sun Y, Fei J, Hou J, Qin Z, Liu Y, Hu B (2009) Simultaneous
determination of dopamine and serotonin using a carbon
nanotubes-ionic liquid gel modified glassy carbon electrode.
Microchim Acta 165:373–379
98. Safavi A, Maleki N, Moradlou O, Tajabadi F (2006) Simultanous
determination of dopamine, ascorbic acid and uric acid using
carbon ionic liquids. Anal Biochem 359:224–229
99. Jia D, Dai J, Yuan H, Lei L, Xiao D (2011) Selective detection of
dopamine in the presence of uric acid using a gold nanoparticles –
poly(luminol) hybrid film and multi-walled carbon nanotubes
with incorporated β-cyclodextrin modified glassy carbon electrode. Talanta 85:2344–2351
100. Garcia MG, Armendariz GME, Godinez LA, Torres J, SepulvedaGuzman S, Bustos E (2011) Detection of dopamine in non-treated
urine samples using glassy carbon electrodes modified with
PAMAM dendrimer-Pt composites. Electrochim Acta 56:7712–
7717
101. Zheng Y, Wang Y, Yang X (2011) Aptamer-based colorimetric
biosensing of dopamine using unmodified nanoparticles. Sens
Actuators B 156:95–99
102. Mahshid S, Li C, Mahshid SS, Askari M, Dolati A, Yang L, Luo
S, Cai Q (2011) Sensitive determination of dopamine in the
3769
presence TiO2 nanotubes modiefied with Pd, Pt and Au nanoparticles. Analyst 136:2322–2329
103. Hosseini M, Momeni MM, Faraji M (2010) An innovative approach to electro-oxidationof dopamine on titanium dioxide
nanotubes electrode modified by gold particles. J Appl
Electrochem 40:1421–1427
104. Ardakani-Mozloum M, Rajabi H, Beitollahi H, Mirjalili BBF,
Akbari A, Taghavinia N (2010) Voltammetric determination of
dopamine at the surface of TiO2 nanoparticles modified carbon
paste electrode. Int J Electrochem Sci 5:147–157
105. Celebańska A, Tomaszewska D, Lesniewski A, Opallo M (2011)
Film electrode prepared from oppositely charged silicate submicroparticles and carbon nanoparticles for selective dopamine
sensing. Biosens Bioelectron 26:4417–4422
106. Plowman BJ, Mahajan M, O’Mullane AP, Bhargava SK (2010)
Electrochemical detection of dopamine and cytochrome c at a
nanostructured gold electrode. Electrochim Acta 55:8953–8959
107. Reddy S, Swamy BEK, Jayadevappa H (2012) CuO nanoparticle
sensor for the electrochemical determination of dopamine.
Electrochim Acta 61:78–86
108. Xia C, Ning W, Long W, Lin G (2010) Synthesis of nanochainassembled ZnO flowers and their application to dopamine sensing. Sens Actuators B 147:629–634
109. Adekunle AS, Agbola BO, Pillay J, Ozoemena KJ (2010)
Electrocatalytical detection of dopamine at single-walled carbon
nanotubes-iron(III)oxide nanoparticles platform. Sens Actuators
B 148:93–102
110. Stoyanova A, Tsakova V (2010) Copper-modified poly (3,4ethylenedioxythiophene) layers for selective determination of
dopamine in the presence of ascorbic acid: II role of the
characteristics of the metal deposition. J Solid State
Electrochem 14:1957–1965
111. Stoyanova A, Ivanov S, Tsakova V, Bund A (2011) Au nanoparticle–polyaniline nanocomposite layers obtained through
layer-by-layer adsorption for the simultaneous determination of
dopamine and uric acid. Electrochim Acta 56:3693–3699
112. Atta NF, El-Kady MF, Galal A (2010) Simulaneous determination
of catecholamines, uric acid and ascorbic acid at physiological
levels using poly(N-methylpyrrole)/Pd clusters sensor. Anal
Biochem 400:78–88
113. Ulubay S, Dursun Z (2010) Cu nanoparticles incorporated polypyrrole modified GCE for sensitive simultaneous determination
of dopamine and uric acid. Talanta 80:1467–1466
114. Ardakani MM, Talebi A, Nacimi H, Barzoky MN, Taghavinia N
(2009) Fabrication of modified TiO2 nanoparticle carbon paste
electrode for simultaneous determination of dopamine, uric acid
and L-cysteine. J Solid State Electrochem 13:1433–1440
115. Atta NF, El-Kady MF, Galal A (2009) Palladium nanoclasterscoated polyfuran as a novel sensor for catecholamine neurotransmitters and paracetamol. Sens Actuators B 141:566–574
116. Arguello J, Leidens VL, Magosso HA, Ramos RR, Gushikem Y
(2008) Simultaneous voltametric determination of ascorbic acid,
dopamine and uric acid by methylene blue adsorbed on phosphorylated zirconia–silica composite electrode. Electrochim Acta
54:560–565
117. Huang J, Liu Y, Hou H, You T (2008) Simultaneous electrochemical determination of dopamine, uric acid and ascorbic acid using
palladium nanoparticle-loaded carbon nanofibers modified electrode. Biosens Bioelectron 24:632–637
118. Katz E, Willner I, Wang J (2003) Electroanalytical and bioelectroanalytical system based on metal and semiconductor nanoparticles. Electroanalysis 16:19–44
119. Hernandez-Santos D, Gonzalez-Garcia MB, Garcia CA (2002) Metalnanoparticles based electroanalysis. Electroanalysis 14:1225–1235
120. Zachek MK, Hermans A, Wightman RM, McCarty GS (2007)
Electrochemical dopamine detection: comparing gold and carbon
3770
fiber microelectrodes using background subtracted fast scan cyclic voltammetry. J Electroanal Chem 614:113–120
121. Zhao Y, Li S-H, Chu J, Chen Y-P, Li W-W, Yu H-Q, Liu G, Tian
Y-C, Xiong Y (2012) A nano-sized Au electrode fabricated using
lithographic technology for electrochemical detection of dopamine. Biosens Bioelectron 35:115–122
122. Zachek MK, Park J, Takmakov P, Wightman RM, McCarty GS
(2010) Microfabricated FSCV-compatible microelectrode array
for real-time monitoring of heterogeneous dopamine release.
Analyst 135:1556–1563
123. Blank CL, Adams RN, Murrill E (1972) Central nervous-system
effects of 6-aminodopamine and 6-hydroxydopamine. Brain Res
45:635–637
124. Adams RN, Murrill E, McCreery R, Blank L, Karolczak M
(1972) 6-Hydroxydopamine, a new oxidation mechanism. Eur J
Pharmacol 17:287–292
125. Troyer KP, Heien MLAV, Venton BJ, Wightman RM (2002)
Neurochemistry and electroanalytical probes. Curr Opin Chem
Biol 6:696–703
126. Venton BJ, Troyer KP, Wightman RM (2002) Response times of
carbon microelectrodes to dynamic changes in catecholamine
concentration. Anal Chem 74:539–546
127. Chen P, Xu B, Tokranova N, Feng X, Castracane J, Gillis KD
(2003) Amperometric detection of quantal catecholamine secretion from individual cells on micromachined silicon chips. Anal
Chem 75:518–524
128. Dong Y, Helen ML, Maxson MM, Ewing AG (2008)
Amperometric measurements of catecholamine release from single vesicles in MN9D cells. J Neurochem 107:1589–1595
129. Li Z-Y, Zhou W, Wu Z-X, Zhang R-Y, Xu T (2009) Fabrication of
size-controllable ultrasmall-disk electrode: monitoring single vesicle release kinetics at tiny structures with high spatio-temporal
resolution. Biosens Bioelectron 24:1358–1364
130. Miller AD, Forster GL, Yeomans JS, Blaha CD (2005) Midbrain
muscarinic receptors modulate morphine-induced accumbal and
striatal dopamine efflux in the rat. Neuroscience 136:531–538
131. Unger EL, Eve DJ, Perez XA, Reichenbach DK, Xu YQ, Lee MK,
Andrews AM (2006) Locomotor hyperactivity and alterations in dopamine neurotransmission are associated with overexpression of A53T
mutant human alpha-synuclein in mice. Neurobiol Dis 21:431–443
132. Shon YM, Lee KH, Goerss SJ, Kim IY, Kimble C, Van Gompel
JJ, Bennet K, Blaha CD, Chang SY (2010) High frequency
stimulation of the subthalamic nucleus evokes striatal dopamine
release in a large animal model of human DBS neurosurgery.
Neurosci Lett 475:136–140
133. Yoshimi K, Naye Y, Mitani N, Kato T, Inpue M, Natori S,
Takahashi T, Weitemeier A, Nishikawa N, McHugh T, Einaga
Y, Kitazawa S (2011) Phasic reward responses in the monkey
striatum as detected by voltammetry with diamond microelectrodes. Neurosci Res 71:49–62
134. Oyama G, Yoshimi K, Natori S, Chikaoka Y, Ren Y-R, Funayama
M, Shimo Y, Takahashi R, Nakazato T, Kitazawa S, Hattori N
(2010) Impaired in vivo dopamine release inparkin knockout
mice. Brain Res 1352:214–222
135. Suzuki A, Ivandini TA, Yoshimi K, Fujishima A, Oyama G,
Nakazato T, Hattori N, Kitazawa S, Einaga Y (2007) Fabrication,
characterization and application of boron-doped diamond microelectrodes for in vivo dopamine detection. Anal Chem 79:8608–8615
136. Howell JO, Kuhr WG, Ensman RF, Wightman RM (1986)
Background subtraction for rapid scan voltammetry. J
Electroanal Chem 209:77–90
137. Dayton MA, Ewing AG, Wightman RM (1983) Diffusional processes measured at microvoltammetric electrodes in brain tissue.
J Electroanal Chem 146:189–200
138. Strand AM, Venton BJ (2008) Flame etching enhances the sensitivity of carbon-fiber microelectrodes. Anal Chem 80:3708–3715
K. Jackowska, P. Krysinski
139. Anastassiou CA, Patel BA, Arundell M, Yeoman MS, Parker KH,
O’Hare D (2006) Subsecond voltammetric separation between
dopamine and serotonin in the presence of ascorbate. Anal
Chem 78:6990–6998
140. Keithley RB, Takmakov P, Bucher ES, Belle AM, OwessonWhite CA, Park J, Wightman RM (2011) Higher sensitivity
dopamine measurements with faster-scan cyclic voltammetry.
Anal Chem 83:3563–3571
141. Jang DP, Kim I, Chang SY, Min HK, Arora K, Marsh MP, Hwang
SC, Kimble CJ, Bennet KE, Lee KH (2012) Paired pulse volatmmetry for differentiating complex analytes. Analyst 137:1428–
1435
142. Lin Y, Trouillon R, Safina G, Ewing AG (2011) Chemical analysis of single cells. Anal Chem 83:4369–4392
143. Amatore C, Arbault S, Guille M, Lemaitre F (2008)
Electrochemical monitoring of single cell secretion: vesicular
exocytosis and oxidative stress. Chem Rev 108:2585–2621
144. Adams KL, Puchades M, Ewing AG (2008) In vitro electrochemistry of biological systems. Annu Rev Anal Chem 1:329–355
145. Wang W, Zhang S-H, Li L-M, Wang Z-L, Cheng J-K, Huang W-H
(2009) Monitoring of vesicular exocytosis from single cells using
micrometer and nanometer-sized electrochemical sensors. Anal
Bioanal Chem 394:17–32
146. Robinson DL, Hermans A, Seipel AT, Wightman RM (2008)
Monitoring rapid communication in the Brain. Chem Rev
108:2554–2584
147. Trouillon R, Svensson MI, Berglund EC, Cans A-S, Ewing AG (2012)
Highlights of selected recent electrochemical measurements in living
systems. Electrochim Acta. doi:10.1016/j.electacta.2012.04.164
148. Schulte A, Schuhmann W (2007) Single-cell microelectrochemistry. Angew Chem Int Ed 46:8760–8777
149. Gotch AJ, Kelly RS, Kuwana T (2003) Characterization and
modeling of the nonfaradaic response of ultrahigh surface area
carbon fibers by electrochemical flow injection analysis. J Phys
Chem B 107:935–941
150. Duvall SH, McCreery RL (2000) Self-catalysis by catechols and
quinones during heterogeneous electron transfer at carbon electrodes. J Am Chem Soc 122:6759–6764
151. Bath BD, Martin HB, Wightman RM, Anderson MR (2001)
Dopamine adsorption at surface modified carbon-fiber electrodes.
Langmuir 17:7032–7039
152. Jacobs CB, Vickerey TL, Venton BJ (2011) Functional groups
modulate the sensitivity and electron transfer kinetics of neurochemicals at carbon nanotube modified microelectrodes. Analyst
136:3557–3565
153. Peairs MJ, Ross AE, Venton BJ (2011) Comparison of Nafionand overoxidized polypyrrole-carbon nanotubes electrodes for
neurotransmitter detection. Anal Chem 3:2379–2386
154. Xiao N, Venton BJ (2012) Rapid, sensitive detection of neurotransmitters at microelectrodes modified with self assembled
SWCNT forests. Anal Chem. doi:10.1021/ac301445
155. Takmakov P, Zachek MK, Keithley RB, Walsh PL, Donley C,
McCarty GS, Wightman RM (2010) Carbon microelectrodes with
a renewable surface. Anal Chem 82:2020–2028
156. Hutton L, Newton ME, Unwin PR, Macpherson JV (2009)
Amperometric oxygen sensor based on a platinum nanoparticlemodified polycrystalline boron doped diamond disk electrode.
Anal Chem 81:1023–1032
157. Johnson MD, Franklin RK, Gibson MD, Brown RB, Kipke DR
(2008) Implantable microelectrode arrays for simultaneous electrophysiological and neurochemical recordings. J Neurosci
Methods 174:62–70
158. Day BK, Pomerleau F, Burmeister JJ, Huettl P, Gerhardt GA
(2006) J Neurochem 96:1626–1635
159. Van Gompel JJ, Chang S-Y, Goers SJ, Kim IY, Kimble C, Bennet
KE, Lee KH (2010) Development of intraoperative electrochemical
New trends in the electrochemical sensing of dopamine
detection: wireless instantaneous neurochemical concentration sensor for deep brain stimulation feedback. Neurosurg Focus 29:E6
160. Chan F-L, Chang W-Y, Kuo L-M, Lin C-H, Wang S-W, Yang Y-S,
Lu MS-C (2008) An electrochemical dopamine sensor with a
CMOS detection circuit. J Micromech Microeng 18:075028
161. Tsai T-C, Guo C-X, Han H-Z, Li Y-T, Huang Y-Z, Li C-M, Chen
J-J J (2012) Microelectrodes with gold nanoparticles and selfassembled monolayers for in vivo recording of striatal dopamine.
Analyst 137:2813–2820
162. Burmeister JJ, Moxon K, Gerhardt GA (2000) Ceramic-based
multisite microelectrodes for electrochemical recordings. Anal
Chem 72:187–192
3771
163. Fernandes SC, Osorio RE-HMB, dos Anjos A, Neves A, Micke
GA, Vieira IC (2008) Determination of catechin in green tea
using catechol oxidase biomimetic sensor. J Braz Chem Soc
19:1215–1223
164. Oliveita IRWZ, Osorio RE-HMB, Neves A, Vieira IC (2007)
Biomimetic sensor based on a novel copper complexes for the
determination of hydroquinone in cosmetics. Sens Actuators B
122:89–94
165. Fernandes SC, Vieira IC, Peralta RA, Neves A (2010)
Development of a biomimetic chitosan film-coated gold electrode for determination of dopamine in the presence of ascorbic
acid and uric acid. Electrochim Acta 55:7152–7157