Recent advances in electrochemical sensing for hydrogen peroxide: a review

Wei Chen , Shu Cai , Qiong-Qiong Ren , Wei Wen and Yuan-Di Zhao *
Britton Chance Center for Biomedical Photonics, Wuhan National Laboratory for Optoelectronics—Huazhong University of Science and Technology, Wuhan, Hubei 430074, P. R. China. E-mail: zydi@mail.hust.edu.cn; Fax: +86 27 87792202; Tel: +86 27 87792235

Received 12th August 2011 , Accepted 16th October 2011

First published on 14th November 2011


Abstract

Due to the significance of hydrogen peroxide (H2O2) in biological systems and its practical applications, the development of efficient electrochemical H2O2 sensors holds a special attraction for researchers. Various materials such as Prussian blue (PB), heme proteins, carbon nanotubes (CNTs) and transition metals have been applied to the construction of H2O2 sensors. In this article, the electrocatalytic H2O2 determinations are mainly focused on because they can provide a superior sensing performance over non-electrocatalytic ones. The synergetic effect between nanotechnology and electrochemical H2O2 determination is also highlighted in various aspects. In addition, some recent progress for in vivo H2O2 measurements is also presented. Finally, the future prospects for more efficient H2O2 sensing are discussed.


1. Introduction

Hydrogen peroxide (H2O2) is a very simple compound in nature but with great importance in pharmaceutical, clinical, environmental, mining, textile and food manufacturing applications.1 In living organisms, besides its well-known cytotoxic effects, H2O2 also plays an essential role as a signalling molecule in regulating diverse biological processes such as immune cell activation, vascular remodelling, apoptosis, stomatal closure and root growth.2–4 H2O2 is also a side product generated from some classic biochemical reactions catalyzed by enzymes such as glucose oxidase (GOx), alcohol oxidase (AlOx), lactate oxidase (LOx), urate oxidase (UOx), cholesterol oxidase (ChoOx), D-amino acid oxidase (DAAO), glutamate oxidase (GlOx), lysine oxidase (LyOx), oxalate oxidase (OxaOx), etc. Therefore, the study on H2O2 detection is of practical significance for both academic and industrial purposes. Conventional techniques for hydrogen peroxide determination such as fluorimetry,5 chemiluminescence,6 fluorescence7 and spectrophotometry8 are complex, costly and time consuming. In comparison, electrochemistry can offer simple, rapid, sensitive, and cost effective means since H2O2 is an electroactive molecule.9

In electrochemistry, H2O2 can be either oxidized or reduced directly at ordinary solid electrodes. However, these processes in analytical applications are limited by slow electrode kinetics and high overpotential which will downgrade the sensing performance and may incur large interferences from other existing electroactive species in real samples such as ascorbate, urate, bilirubin, etc. Thus, the current research on H2O2 detection is mainly focused on electrode modifications in order to decrease the overpotential and increase the electron transfer kinetics. For these considerations, a large range of materials such as redox proteins, dyes, transition metals, metal oxides, metal phthalocyanines, metal porphyrins, redox polymers, and carbon nanotubes have been employed to conduct electrocatalytic H2O2 detection. On the other hand, in recent years, nanomaterials have attracted tremendous research interest because of their desirable chemical, physical and electronic properties that are different from those of bulk materials. Furthermore, the size and structure of nanomaterials can be tailored for designing a novel sensing platform and enhancing sensing performance.10 Herein, some of the above-mentioned materials tailored into or combined with nanomaterials have shown distinct advantages over conventional materials for H2O2 sensing.11–13 The aim of this review is to summarize the recent advances since 2000 and gain an insight into the materials used in the electrocatalytic sensing of H2O2.

2. Materials used for electrocatalytic hydrogen peroxide sensing

2.1. Metal hexacyanoferrates

2.1.1. Ferric hexacyanoferrate. Ferric hexacyanoferrate or Prussian blue (PB) has been denoted as an “artificial peroxidase” because the reduced form of PB—Prussian white—is capable of catalyzing the reduction of H2O2 at low potentials (−50 mV (vs. Ag/AgCl)) like peroxidases.14,15 PB also owns good catalytic specificity to H2O2 due to the polycrystal structure of PB which may allow penetration of only small molecules into its lattice while larger molecules such as ascorbic acid (AA), uric acid (UA), and para-acetylaminophenol (APAP) are excluded.16 Therefore, PB or PB-based composites have been intensively studied and widely used in biosensor constructions.14,17,18 The most straightforward use of PB in H2O2 sensing is to electrochemically coat the working electrode surface with a PB layer followed by overlaying additional layers which may be used for loading enzyme, stabilizing PB, improving selectivity, etc. Various oxidases including GOx,19–21 choline oxidase (ChOx),22,23 ChoOx,24 AlOx,25 galactose oxidase (GaOx),26 GlOx,27 LOx,28 LyOx,29,30 OxaOx31 and xanthine oxidase (XOx)32 have been combined with PB for biosensor construction. Screen printing technology was also employed to fabricate PB based sensors. O'Halloran et al.33 made the bulk modification of the carbon ink by PB microparticles (<38 μm) and Mattos et al.34 modified Au and Pt screen-printed electrodes with electrochemically deposited PB. Ricci et al.23 modified the graphite ink based screen printed electrode surfaces with in situ chemically synthesized PB and found the enhanced life-time and pH stability of resulted sensors.

With the attractive advantages from nanotechnologies, PB based H2O2 sensing has been developed with nanomaterials. Li et al.35 constructed an amperometric biosensor via grafting PB nanoparticles on the polymeric matrix of multiwalled carbon nanotubes (MWCNTs) and poly(4-vinylpyridine) (PVP). The MWCNT/PVP/PB composite films were fabricated by casting films of MWCNTs wrapped with PVP on gold electrodes followed by electrochemical deposition of PB on the MWCNT/PVP matrix. This modified electrode shows largely enhanced sensitivity of 1.3 μA μM−1 cm−2 and a detection limit of 25 nM due to the remarkable synergistic effect of MWCNTs and PB. Li et al.36 deposited PB on the MWCNT modified pyrolytic graphite electrode and found improved electrochemical stability over a wide pH range and larger response to the reduction of hydrogen peroxide compared to a PB modified pyrolytic graphite electrode. One-step electroless depositions of PB nanoparticles on CNTs were carried out by dispersing CNTs in a mixture of Fe3+, [Fe(CN)6]3 and KCl.37,38 The driving forces for CNT assisted PB nanoparticle synthesis are due to the difference in the redox potentials between CNTs and Fe3+/Fe2+ and between CNTs and [Fe(CN)6]3/[Fe(CN)6]4. Later, Du et al.39 fabricated a porous PB–MWCNT film by the electrochemical co-deposition method. The characterization showed that this PB–MWCNT composite film gave a larger response current to the reduction of H2O2 with a sensitivity of 856 μA mM−1 cm−2 and a detection limit of 23 nM. PB has also been electrodeposited on a graphene oxide modified glassy carbon electrode for fabricating a glucose sensor and this graphene oxide/PB hybrid based sensor showed a largely enhanced response to H2O2.40,41 PB itself can also be tailored into different nanostructures for sensor constructions. Liu et al.42 synthesized PB nanoparticles (ca. 50 nm) self-assembled onto a gold electrode using cysteine as a bridge between the gold surface and the nanoparticles. Zhang et al.43 constructed multi-PB nanocluster/enzyme-immobilized poly(toluidine blue) layers by depositing PB and polymerizing toluidine blue alternately from an acidic solution of ferricyanide. Later, Zhao et al.44 used a layer by layer assembly method to construct multi-PB nanoparticle/enzyme-immobilized polymer layers. A bilayer of poly(diallydimethylammonium chloride) (PDDA) and poly(sodium 4-styrenesulfonate) (PSS) is consecutively adsorbed on 3-mercapto-1-propanesulfonic acid modified Au electrode surfaces, forming stable, ultrathin multilayer films. Subsequently, PDDA protected PB nanoparticles and negatively charged glucose oxidase are consecutively adsorbed onto the PSS-terminated bilayer. This process allows the fabrication of sensing membranes with controlled thickness, structural morphology and biocatalyst loading. The template synthesized PB nanostructures were also achieved by depositing PB in lyotropic liquid crystalline templates45 and highly ordered porous anodic alumina (PAA) membrane,46 respectively. Both results demonstrated enhanced sensing performances in comparison with conventional PB based electrodes.

The major drawback of PB in sensing applications is the lack of operational stability in neutral and alkaline solutions because the reduced form of PB, Prussian white, can be dissolved by hydroxide ions.16,47 Therefore, it is necessary to improve the stability of PB at relatively high pH. Fortunately, some studies demonstrated that surfactants including cetyltrimethylammonium bromide (CTAB), polyvinyl alcohol, polyvinyl pyrrolidone, polyallylamine hydrochloride, polydiallyldimethyldiammonium chloride, tetrabutylammonium toluene-4-sulfonate and polystyrene sulfonate can effectively enhance the electrochemical stability of metal hexacyanoferrate.27,48–56

2.1.2. Other metal hexacyanoferrates. After the discovery of the electroactivity of PB, other metal hexacyanoferrates including copper,57–61 nickel,58,62–65 cobalt,58,66–69 chromium,70,71 vanadium,1 ruthenium72,73 and manganese74 hexacyanoferrates have been also investigated for hydrogen peroxide sensing. In comparison with PB based electrodes, these metal hexacyanoferrate based sensors have a similar or lower capability of electrocatalytic reduction for H2O2, but with more electrochemical stabilities over a wide range of pH.18,57,74 This advantage is taken into account because it permits the better sensing operation at physiological pHs which are favoured by enzymes. Another advantage is that metal hexacyanoferrates can perform well in solution containing not only potassium but also other alkali metal cations such as lithium, sodium, rubidium or caesium, contrary to PB that only shows good electroactivity in K+ containing electrolytes.75–78

2.2. Heme proteins

Heme proteins such as horseradish peroxidase (HRP), catalase (CAT), cytochrome c (Cyt c), hemoglobin (Hb), microperoxidase (MP) and myoglobin (Mb) are a category of metalloproteins containing iron centered porphyrin as their prosthetic groups. This iron in the heme can easily undergo oxidation and reduction over a wide range of potentials which are varied by the protein environment around heme groups.79,80 Due to the redox capability of heme proteins, they have the great potential to be used in bioelectrochemical applications, especially biosensors. The most efficient practice for fabricating redox protein based electrochemical biosensors is to establish direct electron transfer between the protein and electrode (the third generation biosensor). The biosensors based on this configuration can offer better selectivity since they are able to operate in a potential range closer to the redox potential of the protein itself, thus giving less exposure to interfering reactions, unlike the redox mediator based ones (the second generation biosensor) that facilitate not only the electron transfer between the electrode and enzyme but also various interfering reactions.81,82 The main challenge to construct the third generation biosensor is to optimize the electron transfer between the heme protein and electrode, because the prosthetic group of heme protein is shielded by the polypeptides, which make the electron transfer distance so long that the tunnelling mechanism rarely happens.82 Thus, a careful design of the biosensor architectures is essential for facilitating direct electron transfer in predefined pathways interconnecting the active centre and the electrode surface.83 Various strategies such as silica sol–gel,84 conducting polymer,85,86 ionic liquid,87–89 self-assembly monolayer90 and layer-by-layer assembly91,92 have been successfully proved effective in building the third generation hydrogen peroxide sensor. Furthermore, with the attractive advantages, more and more nanomaterials have been integrated with heme proteins to construct direct electron transfer based hydrogen peroxide sensors. Table 1 shows the part of recently published heme proteins and nanomaterials based third generation biosensors and their performances.
Table 1 Examples of heme proteins and nanomaterials based third generation biosensors for H2O2
Heme proteins Nanomaterials Linear range/μM Detection limit/μM References
HRP Au nanoparticles 41–6.3 × 102 5.9 93
0.5–5.2 × 103 0.1 94
15–1.1 × 103 9.0 95
Au nanowire array 0.74–1.5 × 104 0.42 96
TiO2 nanotube 0.5–10 and 50–1.0 × 103 0.1 97
Graphene 3.5–3.29 × 102 1.17 98
0.33–14.0 0.11 99
CNTs 0.6–1.8 × 103 0.3 100
1.07–48.4 0.357 101
5.0–1.05 × 103 0.5 102
86–1.0 × 104 86 103
CAT CNTs Up to 0.1 10 104
10–1 × 102 1 105
Au nanoparticles 0.3–6 × 102 0.05 106
1 × 103–5 × 103 107
NiO nanoparticles 1–1 × 103 108
Cyt c Au nanoparticles 8.5 × 102–1.3 × 104 9.8 109
50–1.15 × 103 3 110
10–1.2 × 104 6.3 111
2–3.0 × 102 0.5 112
ZnO nanosheets 1–1 × 103 0.8 113
Hb CNTs 23.6–1.34 × 102 7.87 101
Au nanoparticles 7.4 × 102–1.3 × 104 6.4 109
6 × 102–1.7 × 103 and 2 × 103–2.2 × 104 114
Pt nanoparticles 0.44–44 2.8 × 10−2 115
CdTe nanoparticles 7.44–6.95 × 102 2.23 116
TiO2 nanorods 0–1.02 × 102 0.72 117
Mb CNTs 24.2–1.67 × 102 8.07 101
Au nanoparticles 1.3 × 103–1.3 × 104 1.8 109
0.1–2.35 × 102 118
Nanoporous ZnO 4.8–2.0 × 102 2.0 119


Though HRP, Hb, CAT, Mb and Cyt c have been used to successfully construct the third generation biosensors, respectively, their differences in direct electrochemistry still remain. Lotzbeyer et al.120 made early investigations on the direct electron transfer capability of HRP, Cyt c, Mb, microperoxidase MP-11 and haemin all of which can catalyze the reduction of hydrogen peroxide. These heme proteins were covalently tethered to self-assembled monolayers on the gold surface. As direct electron transfer processes are predominantly limited by the distance between the active site of the protein and the electrode surface, the highest electrocatalytic efficiency was observed for the smallest heme proteins (e.g. microperoxidase MP-11, haemin). Although haemin exhibits a 3300 fold lower catalytic activity for hydrogen peroxide reduction in solution in comparison with HRP, it showed a more than 10 fold higher electrocatalytic activity when immobilized at the monolayer. This tremendous difference could be attributed to a higher surface concentration for the smaller proteins, the improved access for the substrate to their active sites and the increased electron-transfer rate due to the decrease of the distance between the redox site and electrode surface. Wu et al.121 studied the catalysis of Mb, Hb, HRP and CAT in the silk fibroin film on graphite electrodes with direct electrochemistry. At a potential of −0.2 V (vs. SCE), at which the reduction of heme Fe(III) does not interfere, the current was monitored with consecutive H2O2 injection. The result showed that the HRP based electrode possesses the highest sensitivity followed by the CAT modified electrode. Li et al.122 investigated the direct electron transfer reaction of HRP and CAT immobilized by methyl cellulose and found that the HRP based electrode showed about 5 fold higher sensitivity to H2O2 than the CAT based electrode. Peng et al.123 fabricated a magnetic core–shell Fe3O4@Al2O3 nanoparticle attached magnetic glassy carbon electrode for the immobilization of heme proteins including Hb, Mb and HRP. The direct electrochemistry of Hb, Mb and HRP was observed and their apparent Michaelis–Menten constants (KappM) for hydrogen peroxide reduction were calculated as 0.12 mM, 0.105 mM and 0.083 mM, respectively. This result suggested that the HRP based electrode has the highest affinity to H2O2 among the three heme protein based electrodes. Wang et al.124 investigated the effects of 1-butyl-3-methylimidazolium tetrafluoroborate ([bmim][BF4]) on direct electrochemistry and bioelectrocatalysis of Hb, Mb and CAT entrapped in agarose hydrogel films. The KappM of CAT, Hb, and Mb based electrodes for H2O2 reduction was 0.317 mM, 0.755 mM and 0.593 mM, respectively, and the slope of the calibration curves of CAT, Hb, and Mb based electrodes for H2O2 detection was 2.17 mA M−1, 0.24 mA M−1 and 0.996 mA M−1. All the constants and the slopes indicated that the CAT based electrode has the highest affinity and best sensitivity over Hb and Mb based electrodes for H2O2 determination. Later, Wang et al.101 studied direct electrochemistry and electrocatalysis of heme proteins including Hb, Mb and HRP immobilized in SWCNT–CTAB nanocomposite film modified electrodes. The electrocatalytic reduction of H2O2 on these heme protein based electrodes was investigated by amperometry with an applied potential of −0.4 V. The result showed the slope of the calibration curve of HRP, Hb and Mb was 0.133 μA μM−1, 0.112 μA μM−1 and 0.094 μA μM−1, respectively, which suggested that the HRP modified electrode gave the best performance over Hb and Mb based electrodes on H2O2 determination. The same results were also achieved on HRP, Hb and Mb immobilized Au nanoparticle-bacteria cellulose nanofiber modified electrodes by Wang et al.125

2.3. CNTs and graphene

2.3.1. CNTs. CNTs have been widely used for chemical and biological sensing applications due to their high surface area, high electric conductivity, ability to accumulate analyte, alleviation of surface fouling, capability to make surface functionalization and electrocatalytic activity.126,127 Studies have shown that CNTs can electrocatalyze both the oxidation and reduction of H2O2. Wang's group fabricated CNT/Nafion modified electrodes128 and CNT/Teflon modified electrodes129 based on the dispersion of MWCNTs within Nafion and Teflon as binders, respectively. Significant oxidation and reduction currents around +0.20 V (vs. Ag/AgCl) for H2O2 were observed on both sensor constructions. The performance of the MWCNT/Teflon/GOx electrodes operated at the potentials of +0.6 V and +0.1 V (vs. Ag/AgCl) was studied. It was found that low-potential operation results in a highly linear response (over 2–20 mM range), high selectivity and a slower response time (∼1 min vs. 25 s at +0.6 V) (vs. Ag/AgCl). Despite the absence of permselective coating, the sensor response operated at +0.1 V (vs. Ag/AgCl) was not affected by the addition of APAP and UA.129 Manesh et al.130 reported a nanofibrous glucose electrode fabricated by the immobilization of GOx into an electrospun composite membrane consisting of polymethylmethacrylate (PMMA) dispersed with multiwall carbon nanotubes wrapped by a cationic polymer PDDA. This modified electrode recorded significant electrocatalysis for the oxidation and reduction of H2O2 around 0 mV. The operation potential of this electrode was also at +0.1 V (vs. Ag/AgCl) for H2O2 oxidation. Nafion was necessary to eliminate the interferences from UA and AA in this study. Kachoosangi et al.131 fabricated CNT/ionic liquid n-octylpyridinum hexafluorophosphate (OPFP) composite based sensors and found that this configuration showed overwhelming performance over the CNT/mineral oil composite based electrode due to the excellent charge transfer properties and extremely low capacitance of this ion liquid. CNTs dispersed in mineral oil,132 polyaniline (PANI),133,134 polypyrrole (PPy),135 poly(vinyl alcohol) (PVA),136 poly(pyrocatechol violet) (poly-PCV),137 poly(3,4-ethylenedioxythiophene) (PEDOT),138 chitosan139–142 and nano-Fe3O4143 (for magnetic loading) have also shown good electrocatalysis to H2O2. Interestingly, Gomathi et al.144 reported that MWCNT/chitosan nanocomposite modified electrodes were operated based on the oxidation of H2O2 at 0.34 V (vs. Ag/AgCl). Without a permselective membrane at such a potential, the sensor exhibited no interferences from AA, UA and APAP. Rivas' group studied the electrochemical behaviour of MWCNTs dispersed in polyethylenimine (PEI).145 This PEI/CNT modified electrode showed an excellent electrocatalytic activity toward H2O2 because it was found that the overpotentials for the oxidation and reduction of hydrogen peroxide were decreased by 350 mV and 450 mV, respectively, and the currents were also higher than those obtained with other dispersant agents like Nafion, concentrated acids or chitosan. Besides the traditional dispersion, Lin et al.146 developed glucose biosensors based on a CNT nanoelectrode array grown from Ni nanoparticles on a Cr-coated Si substrate. This sensor array was operated at −0.2 V (vs. Ag/AgCl) with good selectivity against the interferences from AA, UA, and AC.

To get further insight into the electrocatalysis of CNT to H2O2, Xu et al.147 compared the nitrogen-doped carbon nanotubes (NCNTs) with MWCNTs. CNTs doped with N can yield a large number of defective sites on nanotube surfaces and have been proved to have largely improved electrocatalysis.148 In this study, the NCNTs exhibited greatly enhanced electrocatalytic activity toward the oxidation and reduction of H2O2 at +0.25 V and −0.1 V (vs. Ag/AgCl), respectively. Compton's group reported a very interesting finding on the electrocatalytic activity of MWCNTs.149 By comparing the electrochemical behaviour of Fe(III) oxide modified basal plane pyrolytic graphite (BPPG) electrode and CNT modified BPPG electrode in H2O2 solution, they found that the electrocatalysis was due to iron oxide particles arising from the chemical vapor deposition nanotube fabrication process rather than due to intrinsic catalysis attributable to the carbon nanotubes arising, for example, from edge plane-like sites/defects.

2.3.2. Graphene. Graphene has attracted considerable attention from both scientific and technological communities due to its unique physicochemical properties and various potential applications in recent years.150 In comparison with CNTs, graphene shows competitive advantages of low cost, ease of processing and safety.151 Graphene is also an ideal platform for electrochemical research since it is free from the contamination of transition metals which are apt to exist in CNTs. Zhou et al.152 characterized a chemically reduced graphene oxide (CR-GO) modified electrode which shows the much lower onset potential of H2O2 oxidation/reduction started at 0.2/0.1 V (vs. Ag/AgCl) than those of the graphite electrode and bare electrode. The superior electrocatalytic activity of CR-GO is most likely attributed to the high density of edge-plane-like defective sites on CR-GO.153–155 Niu's group characterized a PVP-protected graphene/polyethylenimine functionalized ionic liquid (PFIL) electrode for H2O2 electrocatalysis.156 The more positive reduction potential (onset potential at ∼0 V) and more obvious reduction current indicated that the graphene-based modified electrode had much better electrocatalysis toward H2O2 than the PFIL modified electrode. Later, the same group combined both graphene and gold nanoparticles (AuNPs) dispersed in chitosan to make a glucose biosensor.157 This modified electrode exhibited the highest electrocatalytic activity toward H2O2 than the graphene/chitosan, AuNPs/chitosan, and chitosan modified electrodes. This electrocatalysis enhancement might be attributed to the synergistic effect of graphene and AuNPs.158

2.4. Metals and metal oxides

2.4.1. Metals. Transition metals and their compounds are known as good catalysts either because of their ability to adopt multiple oxidation states or, in the case of the metals, to adsorb other substances onto their surface and activate them in the process. On the other hand, nano-sized metals can display unique advantages of enhanced mass transport, high effective surface area, size controlled electrical, chemical and optical properties and effective utilization of expensive materials.13 Thus, transition metal nanoparticles can be made excellent catalysts due to their high ratio of surface atoms with free valences to the cluster of total atoms.159 A wide range of transition metals including platinum (Pt),160 palladium (Pd),160 copper (Cu),161–163 rhodium (Rh),79,80 iridium (Ir)79,163–165 and ferrum (Fe)166 have been successfully used for electrocatalyzed H2O2 determination. As one of the most intensively studied transition metals, various Au nanomaterials with different shapes and structures such as Au nanowire assembling architecture,167 nanoporous Au,168 Au nanocages11 and Au nanoparticles169 have been investigated for electrocatalytic H2O2 reduction. It seems that Au nanocage11 and nanoporous Au168 slightly outperformed others.

To achieve high loading and reserving high catalytic activities for transition metal nanoparticles, carbon based nanomaterials especially CNTs were extensively utilized as supporting substrates due to their high surface area, high conductivity and high electrocatalytic activity. After the first report of Pt nanoparticle/CNT as a sensing platform,159 numerous studies have been conducted based on this strategy for H2O2 sensing though the approaches may slightly vary.170–175 Other transition metal nanoparticles such as Pd nanoparticles176 and Ag nanoparticles177,178 also have been utilized with CNTs for electrocatalytic H2O2 sensing. Furthermore, besides the most used CNTs, graphene,103,179,180 graphite nanoplatelet,13 carbon nanofiber,181,182 silicon nanowire,183 and Au nanowire array184 were also proved to be suitable nanostructural substrates for various transition metal nanoparticles.

In recent years, bimetallic alloy nanoparticles have attracted considerable attention because they have both the favourable catalytic properties over monometallic counterparts and unique advantages from nanomaterials. Xiao et al.185 fabricated AuPt nanoparticles on chitosan–ionic liquid (trihexyltetradecylphosphonium bis(trifluoromethylsulfonyl)imide ([P(C6)3C14][Tf2N])) film by using an ultrasonic electrodeposition method. Chitosan acted as an adsorbent for the metal ions, and [P(C6)3C14][Tf2N] played a dual role of matrix and stabilizer in the formation of the nanoparticles. By comparing the electrochemical response of different modified electrodes toward H2O2 reduction, the AuPt-chitosan-[P(C6)3C14][Tf2N] modified electrode exhibits the lowest overpotential. This synergistic effect can be attributed to the presence of Au that can reduce the strength of Pt–OH formation.186,187 Furthermore, this sensor was operated at a low working potential (−0.05 V vs. Ag/AgCl) with a sensitivity of 3.98 mA mM−1 cm−2, a detection limit of 0.3 nM in a linear range of 5–355 nM and a very high selectivity of less than 5% decrease of the 20 nM H2O2 signal with the existence of a bunch of foreign species. Later, Safavi and Farjami188 used the same strategy to construct a cholesterol biosensor based on a AuPt-Ch-IL modified electrode and also achieved good performance for H2O2 determination. Li et al.189 utilized AuPt alloy nanoparticles deposited on MWCNTs to fabricate an H2O2 based amperometric immunosensor in which AuPt alloy nanoparticles not only could be used to assemble biomolecules with well maintained bioactivity, but also could facilitate the shuttle of electrons. Liu et al.190 reported an amorphous ternary FeNiPt nanomaterial with tunable length to construct an electrochemical sensing platform. The FeNiPt nanorods with large axial ratio exhibit enhanced electrocatalytic activity towards both the oxidation and reduction of H2O2. H2O2 is cathodically determined on FeNiPt-nanorod modified glassy carbon electrodes with relatively high selectivity at the appropriate potential of 0 V (vs. Ag/AgCl). The anodic H2O2 detection based on the same electrode showed a sensitivity of 2.45 mA mM−1 cm−2, which is 4-fold higher than that of the cathodic detection and those Pt nanoparticle-based H2O2 determination, and the detection limit of 40 nM with a dynamic linear range from 100 nM to 30 mM, which is wider than Pt nanoparticles and binary Pt alloys based detection.

2.4.2. Metal oxides. Some transition metal oxides such as manganese oxide,163,191–198 cobalt oxide,199 titanium dioxide,200 copper oxide201,202 and iridium oxide203 have been reported to show electrocatalytic activity to H2O2. However, many of them were based on the electrocatalytic oxidation of H2O2 in which high potentials were applied on the working electrode. For example, MnO2 nanoparticle and dihexadecyl hydrogen phosphate composite film modified electrode was used for H2O2 determination at an applied potential of +0.65 V (vs. saturated calomel electrode (SCE))192 and a TiO2/MWCNT modified electrode was employed to detect H2O2 at an applied potential of +0.4 V (vs. Ag/AgCl).200 From the application point of view, the operation potentials of these sensors are too high to be applied to real biological samples. Xu et al.198 reported the MnO2/MWCNT modified electrode for H2O2 determination which showed high anti-interference property without any permselective membranes against AA, citric acid and UA in spite of the high operation potential of 0.45 V (vs. Ag/AgCl). Bai et al.204 found that MnO2 nanoparticle modified electrodes show bi-direction electrocatalytic ability toward the reduction/oxidation of H2O2. When this modified electrode was operated at 0 V (vs. Ag/AgCl) in the H2O2 reduction pathway, the interference from AA was greatly depressed. Copper oxide seems a more suitable material for H2O2 sensing. Luque et al.202 reported that the CuO/carbon paste electrode showed an excellent electrocatalytic activity towards the oxidation and reduction of H2O2. This modified electrode was operated at a potential of −0.1 V (vs. Ag/AgCl) for glucose detection. Miao et al.201 constructed a CuO nanoparticle/Nafion modified electrode for H2O2 detection. In 0.1 M NaOH, this sensor exhibited a detection limit of 0.06 μM in the range of 0.15 μM–9.00 mM at an applied potential of −0.3 V (vs. SCE). Magnetite has also been reported to own good catalytic activity towards the reduction of H2O2. Lin and Len205 constructed a Fe3O4/chitosan modified glassy carbon rotating disk electrode for a H2O2 sensor with unique features of electrocatalytic reduction and interference elimination. This magnetite based H2O2 sensor exhibited the advantages of low applied potential (−0.2 V vs. Ag/AgCl), low background current, rapid response and long half-life (9 months at room temperature). Comba et al.206 utilized electrosynthesized Fe3O4 nanoparticles within a carbon paste electrode to fabricate a glucose sensor. Due to the low operation potential at −0.1 V (vs. Ag/AgCl), this sensor showed no response to 0.1 μM AA and 0.4 μM UA with the average sensitivity of 32 ± 4 μA M−1 and detection limit of 3.0 × 10−4 M.

2.5. Other electrocatalytic compounds for hydrogen peroxide sensing

Besides above-mentioned materials, some compounds with reversible redox capabilities have been reported in electrocatalytic H2O2 sensing. Metallophthalocyanines and metalloporphyrins such as cobalt phthalocyanine,207,208 cobalt tetraruthenated porphyrin,209 ether-linked cobalt phthalocyanine–cobalt tetraphenylporphyrin,210 and iron phthalocyanine211 are frequently used due to their excellent catalytic properties and high chemical stabilities. Perovskite-type oxides containing transition metals,212,213 redox polymers,162,214 redox dyes215 and ironsulfur protein216 have also been proved effective in electrocatalytic H2O2 sensing.

3. In vivo applications

When the H2O2 sensor is applied to in vivo measurements, more challenges will arise. The sensor should be able to provide appropriate spatial and temporal resolution within acceptable sensitivity and limits of detection. Due to the complex environment of in vivo measurements, sensing selectivity becomes more critical. Sometimes, a trade-off between selectivity and other parameters such as response time is needed. Furthermore, the sensor should be stable enough to conduct the whole experiment. All of these considerations may be balanced to achieve optimum performance according to the specific requirements and conditions of in vivo measurements.

Salazar et al.217 constructed PB modified carbon fiber microelectrodes for H2O2 detection in brain extracellular fluid. To improve stability and selectivity, several polymeric films including Nafion, poly(o-phenylenediamine) (PPD), and a hybrid configuration of these two polymers were investigated. The PPD coating was selected due to its excellent anti-interference properties and stabilization capability for PB against solubilisation at high pH though the sensitivity was about 50% loss due to diffusion phenomena across the polymeric film. The result showed that the PPD film significantly enhanced the selectivity against the main endogenous brain interference species, expressed as the ratio of the sensitivity slopes, which was close to 600 for all interference molecules studied. Later, Salazar et al.218 adopted the same sensor configuration based on PB for in vivo glucose detection in extracellular fluid of the prefrontal cortex. Due to the addition of GOx, Nafion, PEI and PPD were employed to attenuate oxygen dependence, to stabilize the enzyme and to shield the sensor surface from interferences, respectively. The sensor showed the sufficient sensitivity and stability to monitor multi-phasic and reversible changes in brain extracellular fluid.

Kulagina and Michael219 reported a cross-linked redox polymer containing a HRP modified carbon fibre electrode which was used for H2O2 measurements in the brain of anesthetized rats. When implanted in the striatal region of the rat brain, a biphasic response was observed upon electrical stimulation of the dopaminergic pathway that innervates the striatal tissue, while no response was observed at sensors containing no HRP. This study confirmed that amperometric sensors based on carbon fibre microelectrodes can be used for monitoring the kinetic neurochemical activity in the brain. Rui et al.113 constructed Cyt c adsorbed hexagonal ZnO nanosheets based sensors for the extracellular released H2O2 from living cancer cells. The direct electron transfer of Cyt c was realized on the nanostructured ZnO. The result showed that under the optimized potential of 0 V (vs. Ag/AgCl) the electrochemical determination of H2O2 is free from both anodic interferences and cathodic interference of O2.

A strategy to realize in vivo selective detection without sacrifice of response time was reported by Sanford et al.220 The electrochemical technique, background-substrated, fast-scan cyclic voltammetry (FSCV), provides chemical selectivity in addition to high temporal resolution and sensitivity.221 With this approach, a cyclic voltammogram is generated to serve as a chemical signature for the analyte of interest, allowing discrimination from other electroactive species in the brain.161 In this study, the first voltammetric characterization of rapid H2O2 fluctuations at an uncoated carbon fibre microelectrode was demonstrated with unprecedented chemical and spatial resolution. The carbon fibre was electrochemically conditioned on the anodic scan and the irreversible oxidation of peroxide was detected on the cathodic scan. The oxidation potential was dependent on the scan rate which occurred at +1.2 V (vs. Ag/AgCl) at a scan rate of 400 V s−1. The sensor exhibited a detection limit of 2 μM in a linear range up to 2 mM and was successfully tested in brain slices.

However, in vivo biosensors for H2O2 were mostly developed for animal tissues. The in vivo measurements of H2O2 from oxidative bursts in plant (oilseed rape (Brassica napus L.)) which were induced by either biotic or abiotic stresses were successfully achieved by our group.222–224 The response of the oilseed rape to three different stresses including ultraviolet A and C, Sclerotinia sclerotiorum and Cd2+ were investigated, respectively. In our studies, a Pt microparticle modified Pt wire electrode with PPD coating as a selectively permeable layer was polarized at −0.1 V (vs. Ag/AgCl) which showed good sensitivity and selectivity in living plant tissues.

4. Conclusions and future prospects

This review has mainly concentrated on the recent advances of electrocatalytic H2O2 determination. The utilization of electrocatalytic activity of various materials is for effectively reducing the overpotentials for H2O2 reduction or oxidation, thus decreasing the possible interferences from other electroactive species. Though the earlier studies based on the conventional materials have shown considerable achievements for efficient H2O2 determination, the emergence of nanotechnology over the last decade still promoted tremendous progress in this area. Due to their unique physical and chemical properties, nanomaterials not only largely improve the sensitivity and selectivity, but also provide more strategies for H2O2 sensing. For example, CNTs and graphene can be used either as substrates with high specific area for catalytic materials or as electrocatalysts by themselves. In addition, nanomaterials facilitate the development of the third generation biosensor in which electrons can be more easily transferred from a protein prosthetic group to an electrode surface. For in vivo measurements, besides the above-mentioned merits, nanotechnology is expected to provide very high spatial and temporal resolutions for the analyte of interest with less injury to living organisms since the dimension of the probe can be made small enough in comparison with the size of cell.

On the other hand, research on artificial enzymes that mimic natural enzymes such as CAT is very meaningful. Natural enzymes are difficult to be manipulated for biosensor constructions because they are very sensitive to the environment and prone to denature by relatively harsh conditions such as high temperature or organic solvent. The poor stability of natural enzymes quite limits the fabrication process of biosensors and long-time usage in real applications. Furthermore, the cost of enzymes is still high even in mass manufacturing. In comparison, artificial enzymes can be more robust and more easily tailored to the desired properties for biosensor construction and application. For example, some progress has been made in CAT mimics.225–227

A wide range of newly introduced nanomaterials and electrocatalytic compounds is expected to continuously advance H2O2 sensing development. And these sensors will facilitate the understanding of the physiological process participated by H2O2 in living organisms. Furthermore, tremendous influence of H2O2 sensing on clinical diagnostic, pharmaceutical development, food safety and environmental monitoring will gradually appear in the near future.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (grant no. 81071229 and 91017013), Specialized Research Fund for the Doctoral Program of Higher Education of China (no. 20100142110002), the Fundamental Research Funds for the Central Universities (HUST: 2010ZD005, 2011QN235), and the Open Research Fund of State Key Laboratory of Bioelectronics, Southeast University.

Notes and references

  1. C. G. Tsiafoulis, P. N. Trikalitis and M. I. Prodromidis, Electrochem. Commun., 2005, 7, 1398–1404 CrossRef CAS .
  2. M. Geiszt and T. L. Leto, J. Biol. Chem., 2004, 279, 51715–51718 CrossRef CAS .
  3. M. Giorgio, M. Trinei, E. Migliaccio and P. G. Pelicci, Nat. Rev. Mol. Cell Biol., 2007, 8, 722–728 CrossRef CAS .
  4. C. Laloi, K. Apel and A. Danon, Curr. Opin. Plant Biol., 2004, 7, 323–328 CrossRef CAS .
  5. J. H. Lee, I. N. Tang and J. B. Weinstein-Lloyd, Anal. Chem., 1990, 62, 2381–2384 CrossRef CAS .
  6. S. Hanaoka, Anal. Chim. Acta, 2001, 426, 57–64 CrossRef CAS .
  7. E. Fernandes, A. Gomes and J. L. F. C. Lima, J. Biochem. Biophys. Methods, 2005, 65, 45–80 CrossRef .
  8. M. C. O. R. F. P. Nogueira and W. C. Paterlini, Talanta, 2005, 66, 86–91 CrossRef .
  9. J. Wang, Biosens. Bioelectron., 2006, 21, 1887–1892 CrossRef CAS .
  10. J. Wang, Analyst, 2005, 130, 421–426 RSC .
  11. Y. Zhang, Y. Sun, Z. Liu, F. Xu, K. Cui, Y. Shi, Z. Wen and Z. Li, J. Electroanal. Chem., 2011, 656, 23–28 CrossRef CAS .
  12. R. Yuan, L. J. Bai, Y. Q. Chai, Y. L. Yuan, L. Mao and Y. Zhuo, Analyst, 2011, 136, 1840–1845 RSC .
  13. I. Lee, J. Lu, I. Do, L. T. Drzal and R. M. Worden, ACS Nano, 2008, 2, 1825–1832 CrossRef .
  14. A. A. Karyakin, Electroanalysis, 2001, 13, 813–819 CrossRef CAS .
  15. A. A. Karyakin, E. E. Karyakina and L. Gorton, Anal. Chem., 2000, 72, 1720–1723 CrossRef CAS .
  16. A. A. Karyakin, E. E. Karyakina and L. Gorton, Talanta, 1996, 43, 1597–1606 CrossRef CAS .
  17. F. Ricci and G. Palleschi, Biosens. Bioelectron., 2005, 21, 389–407 CrossRef CAS .
  18. R. Koncki, Crit. Rev. Anal. Chem., 2002, 32, 79–96 CrossRef CAS .
  19. R. Garjonyte and A. Malinauskas, Sens. Actuators, B, 2000, 63, 122–128 CrossRef .
  20. M. Ferreira, P. A. Fiorito, O. N. Oliveira and S. I. C. de Torresi, Biosens. Bioelectron., 2004, 19, 1611–1615 CrossRef CAS .
  21. J. J. Zhu, M. H. Xue, Q. Xu and M. Zhou, Electrochem. Commun., 2006, 8, 1468–1474 CrossRef .
  22. D. Moscone, D. D'Ottavi, D. Compagnone, G. Palleschi and A. Amine, Anal. Chem., 2001, 73, 2529–2535 CrossRef CAS .
  23. F. Ricci, A. Amine, G. Palleschi and D. Moscone, Biosens. Bioelectron., 2003, 18, 165–174 CrossRef CAS .
  24. X. Y. Zou, X. C. Tan, M. J. Ll, P. X. Cai and L. J. Luo, Anal. Biochem., 2005, 337, 111–120 CrossRef .
  25. A. Karyakin, E. Karyakina and L. Gorton, Talanta, 1996, 43, 1597–1606 CrossRef CAS .
  26. Y. T. Wang, J. Z. Zhu, R. J. Zhu, Z. Q. Zhu, Z. S. Lai and Z. Y. Chen, Meas. Sci. Technol., 2003, 14, 831–836 CrossRef CAS .
  27. S. Varma, Y. Yigzaw and L. Gorton, Anal. Chim. Acta, 2006, 556, 319–325 CrossRef CAS .
  28. R. Garjonyte, Y. Yigzaw, R. Meskys, A. Malinauskas and L. Gorton, Sens. Actuators, B, 2001, 79, 33–38 CrossRef .
  29. F. Ricci, A. Amine, C. S. Tuta, A. A. Ciucu, F. Lucarelli, G. Palleschi and D. Moscone, Anal. Chim. Acta, 2003, 485, 111–120 CrossRef CAS .
  30. F. Ricci, C. Goncalves, A. Amine, L. Gorton, G. Palleschi and D. Moscone, Electroanalysis, 2003, 15, 1204–1211 CrossRef CAS .
  31. P. A. Fiorito and S. I. C. de Torresi, Talanta, 2004, 62, 649–654 CrossRef CAS .
  32. Y. J. Liu, L. H. Nie, W. Y. Tao and S. Z. Yao, Electroanalysis, 2004, 16, 1271–1278 CrossRef CAS .
  33. M. P. O'Halloran, M. Pravda and G. G. Guilbault, Talanta, 2001, 55, 605–611 CrossRef CAS .
  34. I. L. de Mattos, L. Gorton and T. Ruzgas, Biosens. Bioelectron., 2003, 18, 193–200 CrossRef .
  35. J. Li, J. D. Qiu, J. J. Xu, H. Y. Chen and X. H. Xia, Adv. Funct. Mater., 2007, 17, 1574–1580 CrossRef CAS .
  36. J. H. Chen, Z. F. Li, W. Li, K. Chen, L. H. Nie and S. Z. Yao, J. Electroanal. Chem., 2007, 603, 59–66 CrossRef .
  37. S. J. Dong, J. F. Zhai, Y. M. Zhai and D. Wen, Electroanalysis, 2009, 21, 2207–2212 CrossRef .
  38. B. Fang, W. Zhang, L. L. Wang, N. Zhang and G. F. Wang, Electroanalysis, 2009, 21, 2325–2330 CrossRef .
  39. D. Du, M. Wang, Y. Qin and Y. Lin, J. Mater. Chem., 2010, 20, 1532–1537 RSC .
  40. N. Q. Jia, Y. Z. Zhang, X. M. Sun, L. Z. Zhu and H. B. Shen, Electrochim. Acta, 2011, 56, 1239–1245 CrossRef .
  41. L. Cao, Y. Liu, B. Zhang and L. Lu, ACS Appl. Mater. Interfaces, 2010, 2, 2339–2346 CAS .
  42. S. Q. Liu, J. J. Xu and H. Y. Chen, Electrochem. Commun., 2002, 4, 421–425 CrossRef CAS .
  43. D. Zhang, K. Zhang, Y. L. Yao, X. H. Xia and H. Y. Chen, Langmuir, 2004, 20, 7303–7307 CrossRef CAS .
  44. W. Zhao, J. J. Xu, C. G. Shi and H. Y. Chen, Langmuir, 2005, 21, 9630–9634 CrossRef CAS .
  45. A. A. Karyakin, E. A. Puganova, I. A. Budashov, I. N. Kurochkin, E. E. Karyakina, V. A. Levchenko, V. N. Matveyenko and S. D. Varfolomeyev, Anal. Chem., 2004, 76, 474–478 CrossRef CAS .
  46. L. T. Jin, Y. Z. Xian, Y. Hu, F. Liu, Y. Xian and L. J. Feng, Biosens. Bioelectron., 2007, 22, 2827–2833 CrossRef .
  47. U. Scharf, Electrochim. Acta, 1996, 41, 233–239 CrossRef CAS .
  48. R. Vittal, M. Jayalakshmi, H. Gomathi and G. P. Rao, J. Electrochem. Soc., 1999, 146, 786–793 CrossRef CAS .
  49. R. Vittal, H. Gomathi and G. P. Rao, Electrochim. Acta, 2000, 45, 2083–2093 CrossRef CAS .
  50. S. M. S. Kumar and K. C. Pillai, Electrochem. Commun., 2006, 8, 621–626 CrossRef CAS .
  51. I. Dekany and V. Hornok, J. Colloid Interface Sci., 2007, 309, 176–182 CrossRef .
  52. S. M. S. Kumar and K. C. Pillai, J. Electroanal. Chem., 2006, 589, 167–175 CrossRef CAS .
  53. R. Vittal and H. Gomathi, J. Phys. Chem. B, 2002, 106, 10135–10143 CrossRef CAS .
  54. R. Vittal, K. J. Kim, H. Gomathi and V. Yegnaraman, J. Phys. Chem. B, 2008, 112, 1149–1156 CrossRef CAS .
  55. S. Q. Liu, H. Li, W. H. Sun, X. M. Wang, Z. G. Chen, J. J. Xu, H. X. Ju and H. Y. Chen, Electrochim. Acta, 2011, 56, 4007–4014 CrossRef CAS .
  56. B. Haghighi, S. Varma, F. M. Alizadeh, Y. Yigzaw and L. Gorton, Talanta, 2004, 64, 3–12 CrossRef CAS .
  57. R. Garjonyte and A. Malinauskas, Sens. Actuators, B, 1999, 56, 93–97 CrossRef .
  58. A. Schwake, B. Ross and K. Cammann, Sens. Actuators, B, 1998, 46, 242–248 CrossRef .
  59. J. Wang, X. J. Zhang and M. Prakash, Anal. Chim. Acta, 1999, 395, 11–16 CrossRef CAS .
  60. I. L. de Mattos, L. Gorton, T. Laurell, A. Malinauskas and A. A. Karyakin, Talanta, 2000, 52, 791–799 CrossRef CAS .
  61. P. A. Fiorito, C. M. A. Brett and S. I. C. de Torresi, Talanta, 2006, 69, 403–408 CrossRef CAS .
  62. M. H. Pournaghi-Azar and H. Razmi-Nerbin, Electroanalysis, 2000, 12, 209–215 CrossRef CAS .
  63. C. X. Cai, K. H. Xue, Y. M. Zhou and H. Yang, Talanta, 1997, 44, 339–347 CrossRef CAS .
  64. S. Milardovic, I. Kruhak, D. Ivekovic, V. Rumenjak, M. Tkalcec and B. S. Grabaric, Anal. Chim. Acta, 1997, 350, 91–96 CrossRef CAS .
  65. P. A. Fiorito and S. I. C. de Torresi, J. Electroanal. Chem., 2005, 581, 31–37 CrossRef CAS .
  66. M. S. Lin, Y. C. Wu and B. I. Jan, Biotechnol. Bioeng., 1999, 62, 56–61 CrossRef CAS .
  67. S. M. Chen, J. Electroanal. Chem., 2002, 521, 29–52 CrossRef CAS .
  68. A. Eftekhari, Microchim. Acta, 2003, 141, 15–21 CrossRef CAS .
  69. G. L. Shen, S. P. Wang, L. M. Lu, M. H. Yang, Y. Lei and R. Q. Yu, Anal. Chim. Acta, 2009, 651, 220–226 CrossRef .
  70. M. S. Lin, T. F. Tseng and W. C. Shih, Analyst, 1998, 123, 159–163 RSC .
  71. M. S. Lin and W. C. Shih, Anal. Chim. Acta, 1999, 381, 183–189 CrossRef CAS .
  72. N. Dale, F. Tian and E. Llaudet, Anal. Chem., 2007, 79, 6760–6766 CrossRef .
  73. J. M. Zen, A. S. Kumar and C. R. Chung, Anal. Chem., 2003, 75, 2703–2709 CrossRef CAS .
  74. A. Eftekhari, Talanta, 2001, 55, 395–402 CrossRef CAS .
  75. J. Bácskai, K. Martinusz, E. Czirók, G. Inzelt, P. J. Kulesza and M. A. Malik, J. Electroanal. Chem., 1995, 385, 241–248 CrossRef .
  76. S. Sinha, B. D. Humphrey and A. B. Bocarsly, Inorg. Chem., 1984, 23, 203–212 CrossRef CAS .
  77. S. Zamponi, M. Berrettoni, P. J. Kulesza, K. Miecznikowski, M. A. Malik, O. Makowski and R. Marassi, Electrochim. Acta, 2003, 48, 4261–4269 CrossRef CAS .
  78. M. A. Malik, K. Miecznikowski and P. J. Kulesza, Electrochim. Acta, 2000, 45, 3777–3784 CrossRef CAS .
  79. M. C. Rodriguez and G. A. Rivas, Anal. Lett., 2001, 34, 1829–1840 CrossRef CAS .
  80. S. A. Miscoria, G. D. Barrera and G. A. Rivas, Sens. Actuators, B, 2006, 115, 205–211 CrossRef .
  81. L. Gorton, Electroanalysis, 1995, 7, 23–45 CrossRef CAS .
  82. L. Gorton, A. Lindgren, T. Larsson, F. D. Munteanu, T. Ruzgas and I. Gazaryan, Anal. Chim. Acta, 1999, 400, 91–108 CrossRef CAS .
  83. R. S. Freire, C. A. Pessoa, L. D. Mello and L. T. Kubota, J. Braz. Chem. Soc., 2003, 14, 230–243 CrossRef CAS .
  84. S. P. Bi, J. W. Di, M. Zhang and K. A. Yao, Biosens. Bioelectron., 2006, 22, 247–252 CrossRef .
  85. Y. T. Kong, M. Boopathi and Y. B. Shim, Biosens. Bioelectron., 2003, 19, 227–232 CrossRef CAS .
  86. P. R. Solanki, A. Kaushik, A. A. Ansari, G. Sumana and B. D. Malhotra, Polym. Adv. Technol., 2011, 22, 903–908 CrossRef CAS .
  87. A. Safavi, N. Maleki, O. Moradlou and M. Sorouri, Electrochem. Commun., 2008, 10, 420–423 CrossRef CAS .
  88. B. Z. Zeng, R. Yan, F. Q. Zhao, J. W. Li, F. Mao and S. S. Fan, Electrochim. Acta, 2007, 52, 7425–7431 CrossRef .
  89. X. Shangguan, J. Zheng and Q. Sheng, Electroanalysis, 2009, 21, 1469–1474 CrossRef CAS .
  90. K. M. Razeeb, J. Xu, F. J. Shang, J. H. T. Luong and J. D. Glennon, Biosens. Bioelectron., 2010, 25, 1313–1318 CrossRef .
  91. H. Y. Chen, J. J. Feng and J. J. Xu, Biosens. Bioelectron., 2007, 22, 1618–1624 CrossRef .
  92. G. Y. Shi, Z. Y. Sun, M. C. Liu, L. Zhang, Y. Liu, Y. H. Qu and L. T. Jin, Anal. Chem., 2007, 79, 3581–3588 CrossRef CAS .
  93. Y. Wang, X. Ma, Y. Wen, Y. Xing, Z. Zhang and H. Yang, Biosens. Bioelectron., 2010, 25, 2442–2446 CrossRef CAS .
  94. B. Tang, F. Li, Y. Feng, Z. Wang, L. M. Yang and L. H. Zhuo, Biosens. Bioelectron., 2010, 25, 2244–2248 CrossRef .
  95. L. X. Sun, C. Xiang, Y. Zou and F. Xu, Sens. Actuators, B, 2009, 136, 158–162 CrossRef .
  96. J. Xu, F. Shang, J. H. T. Luong, K. M. Razeeb and J. D. Glennon, Biosens. Bioelectron., 2010, 25, 1313–1318 CrossRef CAS .
  97. Y. Z. Xian, F. H. Wu, J. J. Xu, Y. Tian, Z. C. Hu, L. W. Wang and L. T. Jin, Biosens. Bioelectron., 2008, 24, 198–203 CrossRef .
  98. H. Zhou, Q. A. Zhang, Y. Qiao, L. Zhang, S. Y. Wu, J. W. Xu and X. M. Song, Electroanalysis, 2011, 23, 900–906 CrossRef .
  99. M. Li, S. Xu, M. Tang, L. Liu, F. Gao and Y. Wang, Electrochim. Acta, 2011, 56, 1144–1149 CrossRef CAS .
  100. S. L. Luo, X. D. Zeng, X. F. Li, X. Y. Liu, Y. Liu, B. Kong, S. L. Yang and W. Z. Wei, Biosens. Bioelectron., 2009, 25, 896–900 CrossRef .
  101. S. F. Wang, F. Xie and G. D. Liu, Talanta, 2009, 77, 1343–1350 CrossRef CAS .
  102. Q. J. Xie, Z. J. Cao, X. Q. Jiang and S. Z. Yao, Biosens. Bioelectron., 2008, 24, 222–227 CrossRef .
  103. M.-Y. Hua, Y.-C. Lin, R.-Y. Tsai, H.-C. Chen and Y.-C. Liu, Electrochim. Acta, 2011, 56, 9488–9495 CrossRef CAS .
  104. A. Salimi, L. Miranzadeh, R. Hallaj and H. Mamkhezri, Electroanalysis, 2008, 20, 1760–1768 CrossRef CAS .
  105. A. Salimi, A. Noorbakhsh and M. Ghadermarz, Anal. Biochem., 2005, 344, 16–24 CrossRef CAS .
  106. K. H. Huang, K. J. Huang, D. J. Niu, X. Liu, Z. W. Wu, Y. Fan, Y. F. Chang and Y. Y. Wu, Electrochim. Acta, 2011, 56, 2947–2953 CrossRef CAS .
  107. Y. N. Tian, B. J. Zhou, J. W. Wang and X. L. Gao, Anal. Lett., 2008, 41, 1832–1849 CrossRef .
  108. A. Salimi, E. Sharifi, A. Noorbakhsh and S. Soltanian, Biophys. Chem., 2007, 125, 540–548 CrossRef CAS .
  109. H. Y. Chen, J. J. Feng, G. Zhao and J. J. Xu, Anal. Biochem., 2005, 342, 280–286 CrossRef .
  110. L. X. Sun, C. L. Xiang, Y. J. Zou and F. Xu, Electrochem. Commun., 2008, 10, 38–41 CrossRef .
  111. A. Zhu, Y. Tian, H. Liu and Y. Luo, Biomaterials, 2009, 30, 3183–3188 CrossRef CAS .
  112. J. B. Hu, S. Q. Li, J. Xia, C. Y. Liu, W. Cao and Q. L. Li, J. Electroanal. Chem., 2009, 633, 273–278 CrossRef .
  113. Q. Rui, K. Komori, Y. Tian, H. Liu, Y. Luo and Y. Sakai, Anal. Chim. Acta, 2010, 670, 57–62 CrossRef CAS .
  114. X. H. Xia, H. L. Guo, D. Y. Liu and X. D. Yu, Sens. Actuators, B, 2009, 139, 598–603 CrossRef .
  115. J. J. Fei, S. S. Jia, T. Tian and F. Q. Zhou, Electroanalysis, 2009, 21, 1424–1431 CrossRef .
  116. Y. D. Zhao, Q. Xu, J. H. Wang, Z. Wang, H. Q. Wang and Q. Yang, Anal. Lett., 2009, 42, 2496–2508 CrossRef .
  117. X. Yao, X. L. Xiao and W. Lu, Electroanalysis, 2008, 20, 2247–2252 CrossRef .
  118. H. M. Zhang, W. T. Xie, L. L. Kong, M. X. Kan, D. M. Han and X. J. Wang, J. Nanosci. Nanotechnol., 2010, 10, 6720–6724 CrossRef .
  119. G. Zhao, J. J. Xu and H. Y. Chen, Anal. Biochem., 2006, 350, 145–150 CrossRef CAS .
  120. T. Lotzbeyer, W. Schuhmann and H. L. Schmidt, Bioelectrochem. Bioenerg., 1997, 42, 1–6 CrossRef .
  121. Y. H. Wu, Q. C. Shen and S. S. Hu, Anal. Chim. Acta, 2006, 558, 179–186 CrossRef CAS .
  122. Y. M. Li, X. T. Chen, J. Li and H. H. Liu, Electrochim. Acta, 2004, 49, 3195–3200 CrossRef CAS .
  123. H.-P. Peng, R.-P. Liang and J.-D. Qiu, Biosens. Bioelectron., 2011, 26, 3005–3011 CrossRef CAS .
  124. D. W. Pang, S. F. Wang, T. Chen, Z. L. Zhang and K. Y. Wong, Electrochem. Commun., 2007, 9, 1709–1714 CrossRef .
  125. Y. L. Zhou, W. Wang, T. J. Zhang, D. W. Zhang, H. Y. Li, Y. R. Ma, L. M. Qi and X. X. Zhang, Talanta, 2011, 84, 71–77 CrossRef .
  126. J. Wang, G. D. Liu and M. R. Jan, J. Am. Chem. Soc., 2004, 126, 3010–3011 CrossRef CAS .
  127. F. Patolsky, Y. Weizmann and I. Willner, Angew. Chem., Int. Ed., 2004, 43, 2113–2117 CrossRef CAS .
  128. J. Wang, M. Musameh and Y. H. Lin, J. Am. Chem. Soc., 2003, 125, 2408–2409 CrossRef CAS .
  129. J. Wang and M. Musameh, Anal. Chem., 2003, 75, 2075–2079 CrossRef CAS .
  130. K. P. Lee, K. M. Manesh, H. T. Kim, P. Santhosh and A. I. Gopalan, Biosens. Bioelectron., 2008, 23, 771–779 CrossRef .
  131. M. M. Musameh, R. T. Kachoosangi, I. Abu-Yousef, J. M. Yousef, S. M. Kanan, L. Xiao, S. G. Davies, A. Russell and R. G. Compton, Anal. Chem., 2009, 81, 435–442 CrossRef .
  132. M. D. Rubianes and G. A. Rivas, Electrochem. Commun., 2003, 5, 689–694 CrossRef CAS .
  133. K. P. Lee, P. Santhosh, K. M. Manesh and A. Gopalan, Anal. Chim. Acta, 2006, 575, 32–38 CrossRef .
  134. F. L. Qu, M. H. Yang, J. H. Jiang, G. L. Shen and R. Q. Yu, Anal. Biochem., 2005, 344, 108–114 CrossRef CAS .
  135. T. Yu-Chen, L. Shih-Ci and L. Shang-Wei, Biosens. Bioelectron., 2006, 22, 495–500 CrossRef .
  136. T. Yu-Chen and H. Jing-Dae, J. Nanosci. Nanotechnol., 2007, 7, 3227–3232 CrossRef .
  137. B. Fang, N. Zhang, W. Zhang, A. X. Gu and G. F. Wang, J. Appl. Polym. Sci., 2009, 112, 3488–3493 CrossRef CAS .
  138. K. C. Lin, T. H. Tsai and S. M. Chen, Biosens. Bioelectron., 2010, 26, 608–614 CrossRef CAS .
  139. W. Feng, Z. G. Wu, Y. Y. Feng, Q. Liu, X. H. Xu, T. Sekino, A. Fujii and M. Ozaki, Carbon, 2007, 45, 1212–1218 CrossRef .
  140. D. Ragupathy, P. Gomathi, M. K. Kim, J. J. Park, A. Rajendran, S. C. Lee, J. C. Kim, S. H. Lee and H. D. Ghim, Sens. Actuators, B, 2011, 155, 897–902 CrossRef .
  141. J. Tkac and T. Ruzgas, Electrochem. Commun., 2006, 8, 899–903 CrossRef CAS .
  142. Z. G. Wu, W. Feng, Y. Y. Feng, Q. Liu, X. H. Xu, T. Sekino, A. Fujii and M. Ozaki, Carbon, 2007, 45, 1212–1218 CrossRef CAS .
  143. G. Chen, S. Qu, J. Wang, J. L. Kong and P. Y. Yang, Talanta, 2007, 71, 1096–1102 CrossRef .
  144. P. Gomathi, M. K. Kim, J. J. Park, D. Ragupathy, A. Rajendran, S. C. Lee, J. C. Kim, S. H. Lee and H. D. Ghim, Sens. Actuators, B, 2011, 155, 897–902 CrossRef .
  145. M. D. Rubianes and G. A. Rivas, Electrochem. Commun., 2007, 9, 480–484 CrossRef CAS .
  146. Y. H. Lin, F. Lu, Y. Tu and Z. F. Ren, Nano Lett., 2004, 4, 191–195 CrossRef CAS .
  147. S. Q. Liu, X. A. Xu, S. J. Jiang and Z. Hu, ACS Nano, 2010, 4, 4292–4298 CrossRef .
  148. L. M. Dai, K. P. Gong, F. Du, Z. H. Xia and M. Durstock, Science, 2009, 323, 760–764 CrossRef .
  149. R. G. Compton, B. Sljukic and C. E. Banks, Nano Lett., 2006, 6, 1556–1558 CrossRef .
  150. Y. H. Lin, Y. Y. Shao, J. Wang, H. Wu, J. Liu and I. A. Aksay, Electroanalysis, 2010, 22, 1027–1036 CrossRef .
  151. M. Segal, Nat. Nanotechnol., 2009, 4, 611–613 Search PubMed .
  152. S. J. Dong, M. Zhou and Y. M. Zhai, Anal. Chem., 2009, 81, 5603–5613 CrossRef .
  153. C. E. Banks and R. G. Compton, Analyst, 2005, 130, 1232–1239 RSC .
  154. C. E. Banks, T. J. Davies, G. G. Wildgoose and R. G. Compton, Chem. Commun., 2005, 829–841 RSC .
  155. C. E. Banks, R. R. Moore, T. J. Davies and R. G. Compton, Chem. Commun., 2004, 1804–1805 RSC .
  156. C. S. Shan, H. F. Yang, J. F. Song, D. X. Han, A. Ivaska and L. Niu, Anal. Chem., 2009, 81, 2378–2382 CrossRef CAS .
  157. L. Niu, C. S. Shan, H. F. Yang, D. X. Han, Q. X. Zhang and A. Ivaska, Biosens. Bioelectron., 2010, 25, 1070–1074 CrossRef .
  158. X. H. Xia, J. Li, J. D. Qiu, J. J. Xu and H. Y. Chen, Adv. Funct. Mater., 2007, 17, 1574–1580 CrossRef .
  159. J. H. T. Luong, S. Hrapovic, Y. L. Liu and K. B. Male, Anal. Chem., 2004, 76, 1083–1088 CrossRef .
  160. S. A. Miscoria, G. D. Barrera and G. A. Rivas, Electroanalysis, 2002, 14, 981–987 CrossRef CAS .
  161. M. C. Rodriguez and G. A. Rivas, Anal. Lett., 2000, 33, 2373–2389 CrossRef CAS .
  162. M. C. Rodriguez and G. A. Rivas, Electroanalysis, 2001, 13, 1179–1184 CrossRef CAS .
  163. G. L. Luque, N. F. Ferreyra and G. A. Rivas, Microchim. Acta, 2006, 152, 277–283 CrossRef CAS .
  164. G. A. Rivas and B. Maestroni, Anal. Lett., 1997, 30, 489–501 CrossRef CAS .
  165. M. C. Rodriguez and G. A. Rivas, Electroanalysis, 1999, 11, 558–564 CrossRef CAS .
  166. F. N. Comba, M. D. Rubianes, P. Herrasti and G. A. Rivas, Sens. Actuators, B, 2010, 149, 306–309 CrossRef .
  167. S. Guo, D. Wen, S. Dong and E. Wang, Talanta, 2009, 77, 1510–1517 CrossRef CAS .
  168. J. G. Liu, F. H. Meng, X. L. Yan, J. Gu and Z. G. Zou, Electrochim. Acta, 2011, 56, 4657–4662 CrossRef .
  169. R. Ramaraj and G. Maduralveeran, J. Electroanal. Chem., 2007, 608, 52–58 CrossRef .
  170. G. L. Shen, M. H. Yang, Y. H. Yang, Y. L. Liu and R. Q. Yu, Biosens. Bioelectron., 2006, 21, 1125–1131 CrossRef .
  171. G. L. Shen, M. H. Yang, Y. Yang, H. F. Yang and R. Q. Yu, Biomaterials, 2006, 27, 246–255 CrossRef .
  172. X. Y. Zou, X. H. Kang, Z. B. Mai, P. X. Cai and J. Y. Mo, Talanta, 2008, 74, 879–886 CrossRef .
  173. Y. H. Zhu, L. H. Xu, L. H. Tang, X. L. Yang and C. Z. Li, Electroanalysis, 2007, 19, 717–722 CrossRef .
  174. X. Y. Zou, X. H. Kang, Z. B. Mai, P. X. Cai and J. Y. Mo, Anal. Biochem., 2007, 369, 71–79 CrossRef .
  175. G. L. Shen, F. L. Qu, M. H. Yang and R. Q. Yu, Biosens. Bioelectron., 2007, 22, 1749–1755 CrossRef .
  176. S. Jeon, J. M. You, Y. N. Jeong, M. S. Ahmed, S. K. Kim and H. C. Choi, Biosens. Bioelectron., 2011, 26, 2287–2291 CrossRef .
  177. Z. Li, Y. Shi, Z. L. Liu, B. Zhao, Y. J. Sun, F. G. Xu, Y. Zhang, Z. W. Wen and H. B. Yang, J. Electroanal. Chem., 2011, 656, 29–33 CrossRef .
  178. Q. Chen, W. Zhao, H. C. Wang, X. Qin, X. S. Wang, Z. X. Zhao, Z. Y. Miao, L. L. Chen, M. M. Shan and Y. X. Fang, Talanta, 2009, 80, 1029–1033 CrossRef .
  179. T. J. Park, M. H. Yang, B. G. Choi, H. Park, W. H. Hong and S. Y. Lee, Electroanalysis, 2011, 23, 850–857 CrossRef .
  180. J. H. Jiang, Q. Zeng, J. S. Cheng, X. F. Liu and H. T. Bai, Biosens. Bioelectron., 2011, 26, 3456–3463 CrossRef .
  181. Y. Liu, D. Wang, L. Xu, H. Hou and T. You, Biosens. Bioelectron., 2011, 26, 4585–4590 CrossRef CAS .
  182. J. S. Huang, D. W. Wang, H. Q. Hou and T. Y. You, Adv. Funct. Mater., 2008, 18, 441–448 CrossRef CAS .
  183. L. W. Yang, J. J. Yin, X. Qi, G. L. Hao, J. Li and J. X. Zhong, Electrochim. Acta, 2011, 56, 3884–3889 CrossRef .
  184. M. Jamal, J. Xu and K. M. Razeeb, Biosens. Bioelectron., 2010, 26, 1420–1424 CrossRef CAS .
  185. F. Q. Zhao, F. Xiao, Y. F. Zhang, G. P. Guo and B. Z. Zeng, J. Phys. Chem. C, 2009, 113, 849–855 Search PubMed .
  186. P. Hernandez-Fernandez, S. Rojas, P. Ocon, J. L. GomezdelaFuente, J. SanFabian, J. Sanza, M. A. Pena, F. J. Garcia-Garcia, P. Terreros and J. L. G. Fierro, J. Phys. Chem. C, 2007, 111, 2913–2923 CAS .
  187. S. J. Dong, S. J. Guo and E. Wang, J. Phys. Chem. C, 2008, 112, 2389–2393 Search PubMed .
  188. A. Safavi and F. Farjami, Biosens. Bioelectron., 2011, 26, 2547–2552 CrossRef CAS .
  189. Y. Li, R. Yuan, Y. Chai and Z. Song, Electrochim. Acta, 2011, 56, 6715–6721 CrossRef CAS .
  190. Y. Tian, H. G. Liu, M. Wen, F. Zhang and D. Liu, Anal. Methods, 2010, 2, 143–148 RSC .
  191. Y. H. Lin, X. L. Cui and L. Y. Li, Electrochem. Commun., 2005, 7, 166–172 CrossRef CAS .
  192. S. J. Yao, J. H. Xu, Y. Wang, X. X. Chen, Y. X. Xu and S. S. Hu, Anal. Chim. Acta, 2006, 557, 78–84 CrossRef CAS .
  193. Y.-H. Bai, H. Zhang, J.-J. Xu and H.-Y. Chen, J. Phys. Chem. C, 2008, 112, 18984–18990 CAS .
  194. X. Chen, X. Zhang, W. Yang and D. G. Evans, Mater. Sci. Eng., C: Biomimetic Supramol. Syst., 2009, 29, 284–287 CrossRef CAS .
  195. B. Sljukic and R. G. Compton, Electroanalysis, 2007, 19, 1275–1280 CrossRef CAS .
  196. X. Cui, G. Liu and Y. Lin, Nanomed.: Nanotechnol., Biol. Med., 2005, 1, 130–135 CrossRef CAS .
  197. C. E. Langley, B. Sljukic, C. E. Banks and R. G. Compton, Anal. Sci., 2007, 23, 165–170 CrossRef .
  198. B. Xu, M.-L. Ye, Y.-X. Yu and W.-D. Zhang, Anal. Chim. Acta, 2010, 674, 20–26 CrossRef CAS .
  199. A. Salimi, R. Hallaj, S. Soltanian and H. Mamkhezri, Anal. Chim. Acta, 2007, 594, 24–31 CrossRef CAS .
  200. L.-C. Jiang and W.-D. Zhang, Electroanalysis, 2009, 21, 988–993 CrossRef CAS .
  201. R. Yuan, X. M. Miao, Y. Q. Chai, Y. T. Shi and Y. Y. Yuan, J. Electroanal. Chem., 2008, 612, 157–163 CrossRef .
  202. G. L. Luque, M. C. Rodriguez and G. A. Rivas, Talanta, 2005, 66, 467–471 CrossRef CAS .
  203. H. Elzanowska, E. Abu-Irhayem, B. Skrzynecka and V. I. Birss, Electroanalysis, 2004, 16, 478–490 CrossRef CAS .
  204. J. J. Xu, Y. H. Bai, Y. Du and H. Y. Chen, Electrochem. Commun., 2007, 9, 2611–2616 CrossRef .
  205. M. S. Lin and H. J. Len, Electroanalysis, 2005, 17, 2068–2073 CrossRef CAS .
  206. F. N. Comba, M. D. Rubianes, L. Cabrera, S. Gutierrez, P. Herrasti and G. A. Rivas, Electroanalysis, 2010, 22, 1566–1572 CAS .
  207. K. Wang, J.-J. Xu and H.-Y. Chen, Biosens. Bioelectron., 2005, 20, 1388–1396 CrossRef CAS .
  208. K. I. Ozoemena, P. N. Mashazi and T. Nyokong, Electrochim. Acta, 2006, 52, 177–186 CrossRef .
  209. M. S. M. Quintino, H. Winnischofer, K. Araki, H. E. Toma and L. Angnes, Analyst, 2005, 130, 221–226 RSC .
  210. K. Ozoemena and T. Nyokong, Electrochim. Acta, 2006, 51, 5131–5136 CrossRef CAS .
  211. J. S. Ye, Y. Wen, W. D. Zhang, H. F. Cui, G. Q. Xu and F. S. Sheu, Electroanalysis, 2005, 17, 89–96 CrossRef CAS .
  212. Y. Shimizu, H. Komatsu, S. Michishita, N. Miura and N. Yamazo, Sens. Actuators, B, 1996, 34, 493–498 CrossRef .
  213. G. L. Luque, N. F. Ferreyra, A. Gabriela Leyva and G. A. Rivas, Sens. Actuators, B, 2009, 142, 331–336 CrossRef .
  214. M. Y. Hua, H. C. Chen, C. K. Chuang, R. Y. Tsai, J. L. Jeng, H. W. Yang and Y. T. Chern, Biomaterials, 2011, 32, 4885–4895 CrossRef CAS .
  215. S. S. Narayanan and D. R. S. Jeykumari, Biosens. Bioelectron., 2008, 23, 1404–1411 CrossRef .
  216. J. W. Choi, A. K. Yagati, T. Lee and J. Min, Bioelectrochemistry, 2011, 80, 169–174 CrossRef .
  217. P. Salazar, M. Martin, R. Roche, R. D. O'Neill and J. L. Gonzalez-Mora, Electrochim. Acta, 2010, 55, 6476–6484 CrossRef CAS .
  218. P. Salazar, R. D. O'Neill, M. Martin, R. Roche and J. L. Gonzalez-Mora, Sens. Actuators, B, 2011, 152, 137–143 CrossRef .
  219. N. V. Kulagina and A. C. Michael, Anal. Chem., 2003, 75, 4875–4881 CrossRef CAS .
  220. L. A. Sombers, A. L. Sanford, S. W. Morton, K. L. Whitehouse, H. M. Oara, L. Z. Lugo-Morales and J. G. Roberts, Anal. Chem., 2010, 82, 5205–5210 CrossRef .
  221. R. G. Cooks, Z. Ouyang, Z. Takats and J. M. Wiseman, Science, 2006, 311, 1566–1570 CrossRef CAS .
  222. H. Chen, Q. Xu, S. Y. Liu, Q. J. Zou, X. L. Guo, X. Y. Dong, P. W. Li, D. Y. Song and Y. D. Zhao, Anal. Chim. Acta, 2009, 632, 21–25 CrossRef .
  223. Y. D. Zhao, Q. Xu, F. Wei, Z. Wang, Q. Yang and H. Chen, Sens. Actuators, B, 2009, 141, 599–603 CrossRef .
  224. Q. Xu, F. Wei, Z. Wang, Q. Yang, Y.-D. Zhao and H. Chen, Phytochem. Anal., 2010, 21, 192–196 CrossRef CAS .
  225. B. J. Day, Biochem. Pharmacol., 2009, 77, 285–296 CrossRef CAS .
  226. U. Rauen, T. Kettler-Thiel, H. De Groot, H.-G. Korth and R. Sustmann, Chem. Biol. Drug Des., 2009, 73, 494–501 CAS .
  227. M. U. Triller, W. Y. Hsieh, V. L. Pecoraro, A. Rompel and B. Krebs, Inorg. Chem., 2002, 41, 5544–5554 CrossRef CAS .

Footnotes

This article is part of a web theme in Analyst and Analytical Methods on Future Electroanalytical Developments, highlighting important developments and novel applications. Also in this theme is work presented at the Eirelec 2011 meeting, dedicated to Professor Malcolm Smyth on the occasion of his 60th birthday.
These authors made equal contribution to the paper.

This journal is © The Royal Society of Chemistry 2012