24
REVIEW 1 2 Electrochemical enzymatic biosensors based on metal micro-/nanoparticles- 3 modified electrodes: a review 4 5 Nina D. Dimcheva * , Elena G. Horozova 6 7 Department of Physical Chemistry, Plovdiv University, 24, Tsar Assen Street, BG-4000 8 Plovdiv, Bulgaria 9 10 Received 17 March 2014; Revised 6 June 2014; Accepted 6 June 2014 11 12 13 The functions of metal structures of micro- or nano-dimensions in the sensing 14 mechanisms of amperometric enzyme-based biosensors are considered in the light of the 15 principles of detection of the latter. The applications of metal mono- or bimetallic 16 nanoparticles-modified materials as catalytic electrodes in the fabrication of first-generation 17 and the role which metal nanoparticles play in promoting or enhancing the electron transfer 18 rates in third-generation electrochemical biosensors are reviewed. Some examples of gold 19 NPs functionalised with enzymes via goldthiol chemistry as a strategy for enzyme 20 immobilisation and spatial orientation when developing amperometric biosensors are also 21 discussed. 22 2014 Institute of Chemistry, Slovak Academy of Sciences 23 24 Keywords: nanoparticles, biosensors, electrocatalysis, direct electron transfer (DET) 25 26 27 Introduction 28 29 The first historical proof of the fabrication of metal nanoparticles dates back to the 30 times of the Roman Empire, as evidenced by an artefact currently in the British museum the 31 dichroic Lycurgus Cup. Due to the incorporation of gold (Au) and silver (Ag) nanoparticles in 32 *Corresponding author, e-mail: [email protected]

Electrochemical enzymatic biosensors based on metal micro-/nanoparticles-modified electrodes: a review

Embed Size (px)

Citation preview

REVIEW 1

2

Electrochemical enzymatic biosensors based on metal micro-/nanoparticles-3

modified electrodes: a review 4

5

Nina D. Dimcheva*, Elena G. Horozova 6

7

Department of Physical Chemistry, Plovdiv University, 24, Tsar Assen Street, BG-4000 8

Plovdiv, Bulgaria 9

10

Received 17 March 2014; Revised 6 June 2014; Accepted 6 June 2014 11

12

13

The functions of metal structures of micro- or nano-dimensions in the sensing 14

mechanisms of amperometric enzyme-based biosensors are considered in the light of the 15

principles of detection of the latter. The applications of metal mono- or bimetallic 16

nanoparticles-modified materials as catalytic electrodes in the fabrication of first-generation 17

and the role which metal nanoparticles play in promoting or enhancing the electron transfer 18

rates in third-generation electrochemical biosensors are reviewed. Some examples of gold 19

NPs functionalised with enzymes via gold–thiol chemistry as a strategy for enzyme 20

immobilisation and spatial orientation when developing amperometric biosensors are also 21

discussed. 22

2014 Institute of Chemistry, Slovak Academy of Sciences 23

24

Keywords: nanoparticles, biosensors, electrocatalysis, direct electron transfer (DET) 25

26

27

Introduction 28

29

The first historical proof of the fabrication of metal nanoparticles dates back to the 30

times of the Roman Empire, as evidenced by an artefact currently in the British museum – the 31

dichroic Lycurgus Cup. Due to the incorporation of gold (Au) and silver (Ag) nanoparticles in 32

*Corresponding author, e-mail: [email protected]

2

the glass, its colour appears greenish in a reflected light or ruby-red when light is transmitted 33

through the glass (Freestone et al., 2007). In the Middle Ages, Au nanoparticles of different 34

sizes were used to produce intensively-coloured glass for the stained-glass windows of the 35

cathedral churches of Europe, e.g. Notre Dame de Paris (Aili & Liedberg, 2010). Nowadays, 36

the ever-increasing interest in fabricating metal nanoparticles results from the diversity of 37

their possible applications,, ranging from tumour diagnosis (Wilson, 2008) through 38

electrocatalysts for energy production (Alexeyeva et al., 2006; Bron, 2008; El-Deab, 2010; 39

El-Deab & Ohsaka, 2002a, 2003, 2006, 2007; El-Deab et al., 2005a; Erikson et al., 2009; 40

Lima et al., 2010; Plyasova et al., 2006; Ramos et al., 2010; Sekol et al., 2013) to plasmonic 41

resonance-based sensing (Chandrasekharan & Kamat, 2001; Freeman et al., 1996; Gartia et 42

al., 2013; Ghosh & Pal, 2007; Ishikawa et al., 1996; Jäckel & Feldmann, 2012), e.g. surface 43

plasmon resonance (SPR) or surface-enhanced Raman scattering (SERS). This broad 44

spectrum of applications for nanosized metal particles predetermines the diversity of the 45

protocols for their fabrication, such as wet chemical techniques (Bharathi & Lev, 1997; Sau & 46

Rogach, 2012; Turkevich & Kim, 1970; Turkevich et al., 1951, 1953), photo-decomposition 47

(Krasnansky et al., 1991) and thermal decomposition of salts (Mao et al., 2013; Xue et al., 48

2001), vapour deposition (Whyman, 1996), sputtering (Ishikawa et al., 1996; Mizsei et al., 49

1998; Terauchi et al., 1995), acoustic (Mizukoshi et al., 1997) or radio-frequency-based 50

(Terauchi et al., 1995) methods and miscellaneous electrochemical-deposition techniques 51

(Chikae et al., 2006; Dai & Compton, 2006; Dobbs et al., 2006; El-Deab, 2010; El-Deab & 52

Ohsaka, 2002a; El-Deab et al., 2005b, 2006; Fu et al., 2010; Gao et al., 2005; Gloaguen et al., 53

1999; Ilias et al., 2013; Ma et al., 2009; Plyasova et al., 2006; Quinn et al., 2005; Sau & 54

Rogach, 2012; Tang et al., 2012; Wang et al., 2008, 2009), etc. Due to the numerous 55

advantages which they offer, e.g. the opportunity to tune the particle size and distribution 56

(Plyasova et al., 2006), improved mechanical stability (El-Deab, 2010) and better electrical 57

contact (El-Deab & Ohsaka, 2007; El-Nagar et al., 2013) with the underlying substrate, 58

electrochemical techniques for the deposition of metal nanoparticles on conductive supports 59

are favoured for producing electrocatalysts applicable in either electroanalysis (Zhang et al., 60

2013) or energy-producing devices (Alexeyeva et al., 2006; Bron, 2008; El-Deab, 2009; El-61

Deab & Ohsaka, 2002b, 2003, 2006, 2007; El-Deab et al., 2005a, 2005b, 2005c, 2006; 62

Erikson et al., 2009; Sekol et al., 2013; Shao et al., 2013). 63

Biosensor technologies represent one of the research areas in which nano-materials 64

play important roles in the sensing process. As defined by IUPAC (IUPAC, 2005–2014), a 65

biosensor is a device that makes use of specific biochemical reactions catalysed by a 66

3

biological element for molecular recognition, such as isolated enzymes, antigens/antibodies, 67

nucleic acids, tissues, organelles or whole cells in order to detect chemical compounds, 68

usually by electrical, thermal or optical signals. 69

Electrochemical biosensors possess two main advantages over the other types of 70

biosensors: (i) susceptibility to miniaturisation, and (ii) facility of operation with electrical 71

response – current, potential, or conductivity, which could be easily processed using relatively 72

inexpensive instrumentation. 73

Depending on the type of the measurable response, they could be sub-divided into 74

three types: (i) amperometric – the measurable response is the current, which varies upon 75

addition of the enzyme substrate; (ii) potentiometric – the potential difference between the 76

working and the reference electrode measured under equilibrium conditions (no current 77

flowing) is the electrode response; (iii) impedimetric – the impedance of the electrode is the 78

measurable response. Biosensors measuring the conductivity changes upon the addition of the 79

enzyme substrate can be also considered in this group. 80

Among the electrochemical biosensors, enzyme-based amperometric biosensors 81

represent the most commonly used group; this functions on the basis of monitoring the current 82

variation at a polarised electrode, induced by the reaction/interaction of the bio-recognition 83

element with the analyte of interest. Furthermore, amperometric enzyme-based biosensors 84

can, on their part, be classified into three categories in accordance with the mode of action 85

(Montornes et al., 2008): (i) first-generation biosensors: the signal corresponds to the 86

electrochemical reaction of an active reagent (monitoring the decrease in the current) or 87

product (monitoring the increase in the current) that is involved in biochemical transformation 88

of the target compound; (ii) second-generation biosensors: the architecture of these biosensors 89

includes a redox mediator either freely diffusing or spatially confined on the electrode surface 90

(small molecular mass compounds, able to effectively shuttle electrons between the electrode 91

surface and the enzyme active site); in this mode, the concentration of the target analyte 92

taking part in the biochemical reaction is proportional to the response resulting from the 93

oxidation/reduction of the mediator; (iii) third-generation biosensors: the bio-component is 94

capable of directly (mediators-free) exchanging electrons between the active site of the 95

enzyme and the transducer (Ghindilis et al., 1997; Gorton et al., 1999); consequently, the 96

concentration of analyte is directly proportional to the redox current generated at the polarised 97

electrode. The third-generation biosensors possess the advantages of simplicity of 98

construction, exclusion of additional supportive substances (e.g. mediator), increase in 99

4

specificity for towards target analyte, removal of interferences due to the usually low 100

polarisation potential at the working electrode, etc. 101

In the past decade, several review papers comprehensively discussed the uses of the 102

magnetic, optical, chemical and physical properties of nanoparticles of different natures in 103

bioanalytical techniques (de Dios & Díaz-García, 2010; Doria et al., 2012; Luo et al., 2006) 104

such as immune-assay, plasmonic biosensors, nucleic acids-based sensors, etc. The current 105

review focuses in particular on the roles played by metal nano- and microparticles-modified 106

electrodes in the first- and third-generation amperometric biosensors – they may have 107

electrocatalytic properties towards the redox transformation of a reaction product or reagent, 108

may bind to the biocatalyst, thereby retaining it closer to the electrode surface, they may also 109

affect enzyme-orientation or reduce the distance between its active site and the electrode 110

surface. Second-generation biosensors will be excluded from this review since they require a 111

redox-active species able to switch between the oxidised and the reduced form in order to 112

mediate the electron transport between the electrode and the enzyme active site. 113

114

115

First-generation biosensors 116

117

The simplest way to construct an enzyme electrode working for the first-generation 118

(1G) amperometric biosensors is to put a biocatalyst in contact with a catalytic electrode. As 119

stated above, the detection of the analyte is based on the monitoring of the current variation 120

due to the redox transformation of either an electroactive product or a reagent of the 121

biocatalysed process. Accordingly, a hydrogen peroxide-producing or -consuming enzyme or 122

protein is often used as a biocatalyst, although two non-oxidative enzymes, urease and 123

acetylcholinesterase (AChE), are also known to work for the first-generation amperometric 124

biosensors. In this case, the metal nano- or microparticles (NPs or MPs)-modified transducer 125

may function as either the electrocatalysts of the response-generating process or they act as 126

anchors to which the enzyme is attached. 127

128

129

Biosensors based on detection of enzymatically produced hydrogen peroxide 130

131

Hydrogen peroxide (H2O2) is usually generated as a by-product in most of the 132

enzyme-catalysed oxidative reactions, where its quantity is directly proportional to the 133

5

concentration of the analyte of interest – the enzyme substrate. H2O2 is the electrochemically 134

active compound and might be electrocatalytically oxidised on solid platinum (Pt) electrodes 135

at an applied potential of 0.6 V (vs Ag/AgCl) (Charmantray et al., 2013). Several examples, 136

however, have shown that the electro-oxidation of enzymatically generated H2O2 onto the 137

electrodes modified with Pt NPs (Olivia et al., 2004), Pt NPs–CNT composite (Zhao et al., 138

2007), Ir NPs (Hsu et al., 2010) or with bimetallic Pd/Pt microdeposits (Horozova et al., 139

2000) affords a more sensitive determination than the bulk Pt electrode. 140

The main drawback of this type of sensing is that, at the operating potentials at which 141

the electrooxidation of H2O2 takes place (i.e. above 0.5 V), organic compounds such as 142

ascorbic or uric acids, some neurotransmitters, antioxidants, pigments, drugs, etc., normally 143

present in the real samples, could be co-oxidised, thereby contributing to the electrode 144

response, which consequently results in overestimated analyte levels. One strategy to 145

overcome this issue is to cover the electrode surface with a perm-selective membrane to 146

restrict the access of the potentially interfering compounds (Kirwan et al., 2007). A much 147

more successful strategy, however, is to electrochemically reduce the peroxide at operating 148

potentials close to 0 V over electrodes modified with metal – mostly iron, hexacyanoferrates 149

(Karyakin, 2001), such as Prussian Blue (also known as artificial peroxidase). Although the 150

preparation of these modifiers is easy, their applications are rather limited because they form a 151

stable film only under acidic conditions, whilst at neutral and slightly alkaline pH they 152

dissolve and tend to lose their catalytic activity. 153

An alternative to the hexacyanoferrate-modified transducers for first-generation 154

biosensors which provides a stable electrode response over a pH range of 3 to 11, was 155

proposed previously (Dimcheva et al., 2002a, 2002b). These first-generation biosensors were 156

based on conventional graphite electrodes onto which a thin layer of bimetallic catalytic 157

phase, consisting of palladium (Pd) and Pt, was allowed to grow through electrodeposition at 158

a constant potential. Subsequent studies showed that the electrocatalytic activity towards 159

H2O2 reduction depended on the ratio at which the two metals were deposited (Dodevska et 160

al., 2006; Nagaiah et al., 2013). Hence, the biosensors thus produced exhibited a practically 161

interference-free response towards the target analyte (Dodevska et al., 2010; Horozova et al., 162

2009) due to the near-zero operating potential. Detailed SEM studies of the electrodeposited 163

bimetallic phases (Nagaiah et al., 2013) demonstrated that the deposits consisted of nanosized 164

crystallites which formed bigger agglomerates upon piling up, the shape and dimensions of 165

which were closely dependent on the ratio of the two metals. In the same study, XRD 166

confirmed that the two co-deposited metals formed alloys, thus explaining why their catalytic 167

6

activity towards H2O2 electroreduction greatly exceeded those of the solely Pd or Pt deposits. 168

Several examples of first-generation biosensors where the detection principle resides in the 169

electroreduction of the enzymatically generated H2O2 are summarised in Table 1. As seen 170

from the examples presented, in all the cases the detection of H2O2 is performed over the 171

potential range of −0.1–0.2 V, in neutral buffers. Under such operating conditions, it is 172

scarcely conceivable that some of the substances normally present in real samples, such as 173

glutathione or other antioxidants, neurotransmitters, phenols, etc. might be electrochemically 174

active and contribute to the electrode response. Therefore, as discussed above, the reduced 175

working potentials warrant a practically interference-free assay of the target analyte. 176

177

178

Two further examples of first-generation biosensors 179

180

Urease is a hydrolase-type enzyme which catalyses the hydrolysis of urea to 181

ammonium and bicarbonate ions according to the equation (Singh et al., 2008): 182

183

NH2CONH2 + 3H2O 2NH4 + OH– + HCO3

urease –+ (1) 184

185

Because the pH of the medium changes upon release of these ions, urease-based 186

potentiometric biosensors are quite commonly used for urea monitoring. This type of 187

biosensors was extensively discussed in two recent reviews (Dhawan et al., 2009; Singh et al., 188

2008), and will not be considered here as being beyond the scope of this review. 189

Electrodeposited rhodium nanoparticles were recently reported as exhibiting a 190

significant catalytic effect on the electro-oxidation of enzymatically liberated ammonia 191

(Velichkova et al., 2011) by the urease-based amperometric biosensor. A similar effect was 192

reported by Jia et al. (2011) due to the Pt-nanoflowers-decorated transducer of their biosensor 193

for urea. Some other nanoparticles-modified electrode materials have been found to contribute 194

to increased sensitivity of the determination, e.g. Ag NPs (Crespilho et al., 2009), Au NPs 195

(Do et al., 2011), or Au NPs-polymer composite material (Tiwari et al., 2009); however the 196

role of the metal nanoparticles is not justified. 197

Acetylcholinesterase (AChE) also belongs to the hydrolases group; it plays a key role 198

in the nervous system of mammalian organisms as one of the factors responsible for 199

transmitting nerve impulses to the cholinergic synapses (Arduini et al., 2010). It may readily 200

7

be inhibited by a variety of toxic compounds: heavy metals, nicotine and neuroparalytic gases, 201

medications for the treatment of Alzheimer’s or Parkinson’s diseases, or pesticides 202

(organophosphorous and carbamate types). The enzyme catalyses the hydrolysis of 203

acetylcholine to choline and acetic acid. Since none of these reaction products is 204

electrochemically active, acetylthiocholine is often used as an enzyme substrate (Pandey et 205

al., 2000). The principle of the detection of AChE enzyme inhibitors using acetylthiocholine 206

as the enzyme substrate lay in the following three stages (Arduini et al., 2010; Stoytcheva et 207

al., 1998). Stage 1: AChE catalyses the hydrolysis of acetylthiocholine to thiocholine and 208

acetic acid at ambient temperature: 209

210

Acetylthiocholine + H2O Thiocholine + CH3COOHAChE

(2) 211

212

Both the non-enzymatic and the enzymatic hydrolysis of acetylthiocholine to thiocholine 213

depended on the pH of the solution and were found to slow down in basic media; stage 2: 214

electrochemical oxidation of thiocholine onto the electrode surface: 215

216

2Thiocholine Disulphide + 2H+ + 2e–anodic oxidation

(3) 217

218

The process takes place with the transfer of one electron from the thiocholine and the 219

subsequent dimerisation of the intermediate to disulphide. When conventional electrodes are 220

used (Martorell et al., 1994; Marty et al., 1992, 1995), thiocholine electro-oxidation proceeds 221

at potentials greater than 0.7 V (vs SCE); stage 3: upon the inhibition of the AChE enzyme, 222

the thiocholine oxidation current decreases. The extent of enzyme inhibition is determined in 223

accordance with the equation: 224

225

%100InhibitionS

IS

I

II (4) 226

227

where IS is the steady-state electrode response to the control solution (at the large thiocholine 228

concentration present); II is the enzyme electrode response at the same thiocholine 229

concentration plus a given inhibitor concentration. 230

A simple and versatile technique for enzyme immobilisation based on gold-thiol 231

chemistry is the spontaneous chemisorption of the sulphur-containing organic compound 232

8

(cysteamine, mercaptobenzothiazole, etc.) on the colloid gold-modified electrode surfaces, 233

thereby forming a self-assembled monolayer, SAM, to which the AChE enzyme is attached 234

through intermolecular complexes (Du et al., 2008, 2009; Liu & Lin, 2006; Pedrosa et al., 235

2007; Somerset et al., 2009). This immobilisation approach provides conformational 236

flexibility of the enzyme and hence assures a more efficient biocatalytic process, albeit 237

enabling the formation of a thicker diffusion layer that reduces electrode sensitivity. An even 238

simpler method for AChE immobilisation is to chemisorb it directly on the surface of Au NPs 239

without the need for additional thiolated linkers (Du et al., 2007a, 2007b; Marinov et al., 240

2010; Shulga & Kirchhoff, 2007), since its protein shell contains terminal SH-groups. The 241

obvious challenge in this case, however, entails retaining the enzyme-functionalised NPs on 242

the electrode, which is implemented by either their incorporation in a porous membrane 243

(Marinov et al., 2010) or by inclusion in a composite film cast on the surface of the working 244

electrode (Du et al., 2007a, 2007b), thereby yielding a stable but thick catalytic layer that 245

triggers additional diffusion limitations. Direct growth of Au NPs on the surface of a compact 246

graphite electrode through electrodeposition allows for a much better electrical contact 247

between the two conductive materials, resulting in a much improved electrode sensitivity 248

towards thiocholine (Dimcheva et al., 2013a). Self-assembly of the AChE enzyme to the Au 249

NPs thus deposited yielded a membrane-free and, hence, fast-responding AChE-based 250

biosensor with a storage stability exceeding one month. 251

252

253

Direct electron transfer promoted by metal nanoparticles and its application in third-254

generation biosensors 255

256

The role of metal nanosized particles in third-generation biosensors is much more 257

complex, since it is believed that, in many cases, they promote the direct electron transfer 258

(DET) process between the electrode and the enzyme active site by shortening the distance 259

between them or by orienting the enzyme upon its immobilisation (Liu et al., 2003; Pingarrón 260

et al., 2008). The most systematically DET-promoting abilities of Au NPs have been explored 261

in relation to the development of cathodes for the enzymatic biofuel cells bearing multi-262

copper oxidases–laccases (Brondani et al., 2013; Krikstolaityte et al., 2014; Rahman et al., 263

2008) or bilirubin oxidases (Mano & Edembe, 2013; Pita et al., 2013), as catalysts for ORR 264

(oxygen reduction reaction). 265

9

Some of the earlier reports on the direct electrochemistry promoted by the 266

nanoparticles of noble metals concerned haem-containing redox proteins such as myoglobin 267

(Dimcheva et al., 2011; Liu & Ju, 2003), haemoglobin (Zhao et al., 2006) and the larger 268

molecule of the peroxidase enzyme (Di et al., 2005). Catalase also belongs to the haem-269

containing enzymes; it consists of four sub-units forming a tetramer with the huge molecular 270

mass of 275 kDa and its active site was found to be buried deep in the protein shell; all these 271

conditions suggest the electrochemical inactivity of the enzyme. Huang et al. (2011), 272

however, observed direct electrochemistry and electrocatalysis with catalase immobilised on 273

an Au NPs-amine-functionalised graphene composite, most probably due to the reduced 274

distance between the enzyme active centre and the electrode. In all of the examples discussed 275

above, where the direct electrochemistry of haem-containing proteins/enzymes was 276

established, it was further used to construct third-generation amperometric biosensors for 277

H2O2. 278

Although glucose oxidase, a flavoprotein with the active site placed deeply inside the 279

protein globule, is considered to be electrochemically inactive, several recent studies have 280

reported on its direct electrochemistry and electrocatalysis applicable in third-generation 281

glucose biosensors (Cui et al., 2014; Haghighi & Tabrizi, 2013; Holland et al., 2011; Liu & 282

Huang, 2012; Palanisamy et al., 2014; Ragupathy et al., 2009; Xu et al., 2012; Yu et al., 283

2014). In connection with this, a very successful strategy to greatly increase the biosensor 284

sensitivity towards its substrate (glucose) by electrically wiring the enzyme via highly ordered 285

nanostructures has been developed by Adeloju and co-workers (Cui et al., 2014; Xu et al., 286

2012). Their glucose nanobiosensor arrays exhibited extraordinary operational characteristics 287

(Cui et al., 2014), e.g. a limit of detection below 1 µM and a linear dynamic range spanning 288

three orders of magnitude (5–6000 µM) and the capacity to respond interference-free to 289

glucose in real samples. 290

Another interesting example of a flavoenzyme believed to be electrochemically inert, 291

but exhibiting direct electrochemistry when immobilised on metal NPs-modified electrodes, is 292

cholesterol oxidase (Karimi et al., 2012; Zhu et al., 2013). Similarly, third-generation 293

cholesterol biosensors demonstrate low limits of detection, high sensitivity and excellent 294

selectivity towards the analyte of interest. 295

Au NPs were found to be useful in promoting DET processes even for the more 296

complex oxidative enzyme – the two-domain sulphite oxidase (Frasca et al., 2012). The 297

strong electrocatalytic effect on the oxidation of sulphite was the basis for constructing a fast-298

responding and highly sensitive third-generation sulphite biosensor. Analogously, direct 299

10

electrochemistry and electrocatalysis by chemisorbed ascorbate oxidase (a multi-copper 300

oxidase, MCO) triggered by Au NPs-modified electrodes was recently reported by our group 301

(Dimcheva et al., 2013b; Dodevska et al., 2013). It is generally believed that two-substrate 302

MCOs only work for third-generation biosensors towards oxygen reduction, due to the 303

internal electron transfer between the three active centres of these enzymes, and that they can 304

eventually oxidise their second substrate (bilirubin, phenols or L-ascorbate) solely under de-305

oxygenated conditions. The AOx enzyme chemisorbed on Au NPs, however, demonstrates 306

the somewhat unusual ability to work for the third-generation L-ascorbate biosensor even 307

under oxygenated conditions. It is possible that a part of the enzyme binds to the Au NPs in 308

close proximity to the ascorbate-linking site (T1), thereby creating a shortcut between it and 309

the underlying electrode which renders the internal electron transport a less favourable 310

pathway. Comparing the electrode response towards L-ascorbate in the absence and in the 311

presence of oxygen, it may be estimated that approximately 1/3 of the enzyme chemisorbs 312

onto the Au NPs in this unusual conformation. Comparative experiments have shown that, on 313

a flat solid gold electrode, the AOx enzyme chemisorbs, shows the ability to 314

electrochemically reduce oxygen, but does not electrocatalytically oxidise L-ascorbate under 315

aerated conditions (Dimcheva et al., 2014). 316

317

318

Conclusions 319

320

In brief, the most beneficial feature of the micro-/nanoparticles-modified transducers, 321

in view of the development of first-generation amperometric biosensors, is their 322

electrocatalytic activity towards the target compound, which is an electroactive either reagent 323

or product of the enzyme-catalysed process. In addition, when the modifiers are Au NPs or Pt 324

NPs, the nanostructures may serve as anchors to which the biocatalyst is attached, largely 325

using the gold–thiol or platinum–thiol chemistry. 326

DET is an exclusive feature of the oxidative enzymes/proteins: their active sites 327

consist of either metal clusters (e.g. Fe/S or Cu clusters) or redox molecules (haemin, FAD, 328

NADH, PQQ), which are buried in the protein shell. Achieving the direct electrochemistry of 329

the enzyme and electrocatalysis in the presence of its typical substrate is the necessary 330

prerequisite to construct third-generation biosensors with exclusive selectivity and greatly 331

improved operational characteristics. Metal NPs or MPs may trigger these two events or may 332

even provoke some unexpected modifications of the electrocatalytic activity. 333

11

334

335

Acknowledgements. The authors gratefully acknowledge the financial support received 336

from the Bulgarian National Science Fund (grant no. DVU-02/38) and the technical help 337

received from the student, Boris Dzhuvinov (Computer Science, University of Plovdiv). 338

339

340

References 341

342

Aili, D., & Liedberg, B. (2010). Biomimetic approaches to self-assembly of nanomaterials. In 343

C. S. S. R. Kumar (Ed.), Biomimetic and bioinspired nanomaterials (Series: Nanomaterials 344

for the life sciences, Chapter 10, pp. 343–377). Weinheim, Germany: Wiley-VCH. DOI: 345

10.1002/9783527610419.ntls0212. 346

Alexeyeva, N., Laaksonen, T., Kontturi, K., Mirkhalaf, F., Schiffrin, D. J., & Tammeveski, K. 347

(2006). Oxygen reduction on gold nanoparticle/multi-walled carbon nanotubes modified 348

glassy carbon electrodes in acid solution. Electrochemistry Communications, 8, 1475–1480. 349

DOI: 10.1016/j.elecom.2006.07.002. 350

Arduini, F., Amine, A., Moscone, D., & Palleschi, G. (2010). Biosensors based on 351

cholinesterase inhibition for insecticides, nerve agents and aflatoxin B1 detection (review). 352

Microchimica Acta, 170, 193–214. DOI: 10.1007/s00604-010-0317-1. 353

Batra, B., & Pundir, C. S. (2013). An amperometric glutamate biosensor based on 354

immobilization of glutamate oxidase onto carboxylated multiwalled carbon nanotubes/gold 355

nanoparticles/chitosan composite film modified Au electrode. Biosensors and 356

Bioelectronics, 47, 496–501. DOI: 10.1016/j.bios.2013.03.063. 357

Bharathi, S., & Lev, O. (1997). Direct synthesis of gold nanodispersions in sol–gel derived 358

silicate sols, gels and films. Chemical Communications, 1997, 2303–2304. DOI: 359

10.1039/a705609e. 360

Bron, M. (2008). Carbon black supported gold nanoparticles for oxygen electroreduction in 361

acidic electrolyte solution. Journal of Electroanalytical Chemistry, 624, 64–68. DOI: 362

10.1016/j.jelechem.2008.07.026. 363

Brondani, D., de Souza, B., Souza, B. S., Neves, A., & Vieira, I. C. (2013). PEI-coated gold 364

nanoparticles decorated with laccase: A new platform for direct electrochemistry of enzymes 365

and biosensing applications. Biosensors and Bioelectronics, 42, 242–247. DOI: 366

10.1016/j.bios.2012.10.087. 367

12

Chandrasekharan, N., & Kamat, P. V. (2001). Assembling gold nanoparticles as 368

nanostructured films using an electrophoretic approach. Nano Letters, 1, 67–70. DOI: 369

10.1021/nl000184f. 370

Charmantray, F., Touisni, N., Hecquet, L., & Mousty, C. (2013). Amperometric biosensor 371

based on galactose oxidase immobilized in clay matrix. Electroanalysis, 25, 630–635. DOI: 372

10.1002/elan.201200274. 373

Chauhan, N., Narang, J., Sunny, & Pundir, C. S. (2013). Immobilization of lysine oxidase on 374

a gold–platinum nanoparticles modified Au electrode for detection of lysine. Enzyme and 375

Microbial Technology, 52, 265–271. DOI: 10.1016/j.enzmictec.2013.01.006. 376

Chikae, M., Idegami, K., Kerman, K., Nagatani, N., Ishikawa, M., Takamura, Y., & Tamiya, 377

E. (2006). Direct fabrication of catalytic metal nanoparticles onto the surface of a screen-378

printed carbon electrode. Electrochemistry Communications, 8, 1375–1380. DOI: 379

10.1016/j.elecom.2006.06.019. 380

Choi, H. N., Han, J. H., Park, J. A., Lee, J. M., & Lee, W. Y. (2007). Amperometric glucose 381

biosensor based on glucose oxidase encapsulated in carbon nanotube–titania–Nafion 382

composite film on platinized glassy carbon electrode. Electroanalysis, 19, 1757–1763. DOI: 383

10.1002/elan.200703958. 384

Comba, F. N., Rubianes, M. D., Herrasti, P., & Rivas, G. A. (2010). Glucose biosensing at 385

carbon paste electrodes containing iron nanoparticles. Sensors and Actuators B: Chemical, 386

149, 306–309. DOI: 10.1016/j.snb.2010.06.020. 387

Crespilho, F. N., Iost, R. M., Travain, S. A., Oliveira, O. N., Jr., & Zucolotto, V. (2009). 388

Enzyme immobilization on Ag nanoparticles/polyaniline nanocomposites. Biosensors and 389

Bioelectronics, 24, 3073–3077. DOI: 10.1016/j.bios.2009.03.026. 390

Cui, J., Adeloju, S. B., & Wu, Y. (2014). Integration of a highly ordered gold nanowires array 391

with glucose oxidase for ultra-sensitive glucose detection. Analytica Chimica Acta, 809, 392

134–140. DOI: 10.1016/j.aca.2013.11.024. 393

Dai, X., & Compton, R. G. (2006). Direct electrodeposition of gold nanoparticles onto indium 394

tin oxide film coated glass: Application to the detection of arsenic(III). Analytical Sciences, 395

22, 567–570. DOI: 10.2116/analsci.22.567. 396

de Dios, A. S., & Díaz-García, M. E. (2010). Multifunctional nanoparticles: Analytical 397

prospects. Analytica Chimica Acta, 666, 1–22. DOI: 10.1016/j.aca.2010.03.038. 398

Dhawan, G., Sumana, G., & Malhotra, B. D. (2009). Recent developments in urea biosensors. 399

Biochemical Engineering Journal, 44, 42–52. DOI: 10.1016/j.bej.2008.07.004. 400

13

Di, J., Shen, C., Peng, S., Tu, Y., & Li, S. (2005). A one-step method to construct a third-401

generation biosensor based on horseradish peroxidase and gold nanoparticles embedded in 402

silica sol–gel network on gold modified electrode. Analytica Chimica Acta, 553, 196–200. 403

DOI: 10.1016/j.aca.2005.08.013. 404

Dimcheva, N., Horozova, E., & Jordanova, Z. (2002a). An amperometric xanthine oxidase 405

enzyme electrode based on hydrogen peroxide electroreduction. Zeitschrift für 406

Naturforschung C – A Journal of Biosciences, 57c, 883–889. 407

Dimcheva, N., Horozova, E., & Jordanova, Z. (2002b). A glucose oxidase immobilized 408

electrode based on modified graphite. Zeitschrift für Naturforschung C – A Journal of 409

Biosciences, 57c, 705–711. 410

Dimcheva, N. D., Horozova, E. G., & Dodevska, T. M. (2011). Direct electrochemistry of 411

myoglobin immobilized on non-modified and modified graphite. Bulgarian Chemical 412

Communications, 43, 17–22. 413

Dimcheva, N., Horozova, E., Ivanov, Y., & Godjevargova, T. (2013a). Self-assembly of 414

acetylcholinesterase on gold nanoparticles electrodeposited on graphite. Central European 415

Journal of Chemistry, 11, 1740–1748. DOI: 10.2478/s11532-013-0307-3. 416

Dimcheva, N., Dodevska, T., & Horozova, E. (2013b). Direct electrochemistry of ascorbate 417

oxidase self-assembled on Au-modified glassy carbon. Journal of the Electrochemical 418

Society, 160, H414–H419. DOI: 10.1149/2.025308jes. 419

Dimcheva, N., Horozova, E., & Dodevska, T. (2014). Electrochemical activity of two redox 420

biocatalysts promoted by gold nanostructures: A proof of principle and applications in 421

biosensing. In Book of Abstracts of the 8th National Conference on Chemistry: Chemistry 422

for Sustainable Development, June 26–28, 2014 (pp. 11-O7). Sofia, Bulgaria: University of 423

Chemical Technology and Metallurgy. 424

Do, J. S., Lin, K. H., & Ohara, R. (2011). Preparation of urease/nano-structured polyaniline-425

Nafion®/Au/Al2O3 electrode for inhibitive detection of mercury ion. Journal of the Taiwan 426

Institute of Chemical Engineers, 42, 662–668. DOI: 10.1016/j.jtice.2010.11.007. 427

Dobbs, W., Suisse, J. M., Douce, L., & Welter, R. (2006). Electrodeposition of silver particles 428

and gold nanoparticles from ionic liquid-crystal precursors. Angewandte Chemie, 118, 4285–429

4288. DOI: 10.1002/ange.200600929. 430

Dodevska, T., Horozova, E., & Dimcheva, N. (2006). Electrocatalytic reduction of hydrogen 431

peroxide on modified graphite electrodes: Application to the development of glucose 432

biosensors. Analytical and Bioanalytical Chemistry, 386, 1413–1418. DOI: 10.1007/s00216-433

006-0682-0. 434

14

Dodevska, T., Horozova, E., & Dimcheva, N. (2010). Design of an amperometric xanthine 435

biosensor based on a graphite transducer patterned with noble metal microparticles. Central 436

European Journal of Chemistry, 8, 19–27. DOI: 10.2478/s11532-009-0102-3. 437

Dodevska, T., Horozova, E., & Dimcheva, N. (2013). Electrochemical behavior of ascorbate 438

oxidase immobilized on graphite electrode modified with Au-nanoparticles. Materials 439

Science and Engineering B: Solid-State Materials for Advanced Technology, 178, 1497–440

1502. DOI: 10.1016/j.mseb.2013.08.012. 441

Doria, G., Conde, J., Veigas, B., Giestas, L., Almeida, C., Assunção, M., Rosa, J., & Baptista, 442

P. V. (2012). Noble metal nanoparticles for biosensing applications. Sensors, 12, 1657–443

1687. DOI: 10.3390/s120201657. 444

Du, D., Ding, J., Cai, J., & Zhang, A. (2007a). Electrochemical thiocholine inhibition sensor 445

based on biocatalytic growth of Au nanoparticles using chitosan as template. Sensors and 446

Actuators B: Chemical, 127, 317–322. DOI: 10.1016/j.snb.2007.04.023. 447

Du, D., Huang, X., Cai, J., & Zhang, A. (2007b). Amperometric detection of triazophos 448

pesticide using acetylcholinesterase biosensor based on multiwall carbon nanotube–chitosan 449

matrix. Sensors and Actuators B: Chemical, 127, 531–535. DOI: 10.1016/j.snb.2007.05.006. 450

Du, D., Chen, W., Cai, J., Zhang, J., Qu, F., & Li, H. (2008). Development of 451

acetylcholinesterase biosensor based on CdTe quantum dots modified cysteamine self-452

assembled monolayers. Journal of Electroanalytical Chemistry, 623, 81–85. DOI: 453

10.1016/j.jelechem.2008.06.020. 454

Du, D., Chen, W., Cai, J., Zhang, J., Tu, H., & Zhang, A. (2009). Acetylcholinesterase 455

biosensor based on gold nanoparticles and cysteamine self assembled monolayer for 456

determination of monocrotophos. Journal of Nanoscience and Nanotechnology, 9, 2368–457

2373. DOI: 10.1166/jnn.2009.se24. 458

El-Deab, M. S., & Ohsaka, T. (2002a). An extraordinary electrocatalytic reduction of oxygen 459

on gold nanoparticles-electrodeposited gold electrodes. Electrochemistry Communications, 460

4, 288–292. DOI: 10.1016/s1388-2481(02)00263-1. 461

El-Deab, M. S., & Ohsaka, T. (2002b). Hydrodynamic voltammetric studies of the oxygen 462

reduction at gold nanoparticles-electrodeposited gold electrodes. Electrochimica Acta, 47, 463

4255–4261. DOI: 10.1016/s0013-4686(02)00487-5. 464

El-Deab, M. S., & Ohsaka, T. (2003). Electrocatalysis by nanoparticles: oxygen reduction on 465

gold nanoparticles-electrodeposited platinum electrodes. Journal of Electroanalytical 466

Chemistry, 553, 107–115. DOI: 10.1016/s0022-0728(03)00291-2. 467

15

El-Deab, M. S., Sotomura, T., & Ohsaka, T. (2005a). Oxygen reduction at electrochemically 468

deposited crystallographically oriented Au(1 0 0)-like gold nanoparticles. Electrochemistry 469

Communications, 7, 29–34. DOI: 10.1016/j.elecom.2004.10.010. 470

El-Deab, M. S., Sotomura, T., & Ohsaka, T. (2005b). Morphological selection of gold 471

nanoparticles electrodeposited on various substrates. Journal of the Electrochemical Society, 472

152, C730–C737. DOI: 10.1149/1.2041948. 473

El-Deab, M. S., Sotomura, T., & Ohsaka, T. (2005c). Size and crystallographic orientation 474

controls of gold nanoparticles electrodeposited on GC electrodes. Journal of the 475

Electrochemical Society, 152, C1–C6. DOI: 10.1149/1.1824041. 476

El-Deab, M. S., & Ohsaka, T. (2006). Electrocatalytic reduction of oxygen at Au 477

nanoparticles–manganese oxide nanoparticle binary catalysts. Journal of the 478

Electrochemical Society, 153, A1365–A1371. DOI: 10.1149/1.2196707. 479

El-Deab, M. S., Sotomura, T., & Ohsaka, T. (2006). Oxygen reduction at Au nanoparticles 480

electrodeposited on different carbon substrates. Electrochimica Acta, 52, 1792–1798. DOI: 481

10.1016/j.electacta.2005.12.057. 482

El-Deab, M. S., & Ohsaka, T. (2007). Electrocatalysis by design: Effect of the loading level 483

of Au nanoparticles–MnOx nanoparticles binary catalysts on the electrochemical reduction 484

of molecular oxygen. Electrochimica Acta, 52, 2166–2174. DOI: 485

10.1016/j.electacta.2006.08.041. 486

El-Deab, M. S. (2009). On the preferential crystallographic orientation of Au nanoparticles: 487

Effect of electrodeposition time. Electrochimica Acta, 54, 3720–3725. DOI: 488

10.1016/j.electacta.2009.01.054. 489

El-Deab, M. S. (2010). Electrocatalysis by nanoparticles: Oxidation of formic acid at 490

manganese oxide nanorods-modified Pt planar and nanohole-arrays. Journal of Advanced 491

Research, 1, 87–93. DOI: 10.1016/j.jare.2010.01.001. 492

El-Nagar, G. A., Mohammad, A. M., El-Deab, M. S., & El-Anadouli, B. E. (2013). 493

Electrocatalysis by design: Enhanced electrooxidation of formic acid at platinum 494

nanoparticles–nickel oxide nanoparticles binary catalysts. Electrochimica Acta, 94, 62–71. 495

DOI: 10.1016/j.electacta.2013.01.133. 496

Erikson, H., Jürmann, G., Sarapuu, A., Potter, R. J., & Tammeveski, K. (2009). 497

Electroreduction of oxygen on carbon-supported gold catalysts. Electrochimica Acta, 54, 498

7483–7489. DOI: 10.1016/j.electacta.2009.08.001. 499

16

Frasca, S., Rojas, O., Salewski, J., Neumann, B., Stiba, K., Weidinger, I. M., Tiersch, B., 500

Leimkühler, S., Koetz, J., & Wollenberger, U. (2012). Human sulfite oxidase 501

electrochemistry on gold nanoparticles modified electrode. Bioelectrochemistry, 87, 33–41. 502

DOI: 10.1016/j.bioelechem.2011.11.012. 503

Freeman, R. G., Hommer, M. B., Grabar, K. C., Jackson, M. A., & Natan, M. J. (1996). Ag-504

clad Au nanoparticles: Novel aggregation, optical, and surface-enhanced Raman scattering 505

properties. The Journal of Physical Chemistry, 100, 718–724. DOI: 10.1021/jp951379s. 506

Freestone, I., Meeks, N., Sax, M., & Higgitt, C. (2007). The Lycurgus Cup – a Roman 507

nanotechnology. Gold Bulletin, 40, 270–277. DOI: 10.1007/bf03215599. 508

Fu, C., Zhou, H., Xie, D., Sun, L., Yin, Y., Chen, J., & Kuang, Y. (2010). Electrodeposition 509

of gold nanoparticles from ionic liquid microemulsion. Colloid and Polymer Science, 288, 510

1097–1103. DOI: 10.1007/s00396-010-2238-2. 511

Gao, F., El-Deab, M. S., Okajima, T., & Ohsaka, T. (2005). Electrochemical preparation of a 512

Au crystal with peculiar morphology and unique growth orientation and its catalysis for 513

oxygen reduction. Journal of the Electrochemical Society, 152, A1226–A1232. DOI: 514

10.1149/1.1906023. 515

Gartia, M. R., Hsiao, A., Pokhriyal, A., Seo, S., Kulsharova, G., Cunningham, B. T., Bond, T. 516

C., & Liu, G. L. (2013). Colorimetric plasmon resonance imaging using nano Lycurgus Cup 517

arrays. Advanced Optical Materials, 1, 68–76. DOI: 10.1002/adom.201200040. 518

Ghindilis, A. L., Atanasov, P., & Wilkins, E. (1997). Enzyme-catalyzed direct electron 519

transfer: Fundamentals and analytical applications. Electroanalysis, 9, 661–674. DOI: 520

10.1002/elan.1140090902. 521

Ghosh, S. K., & Pal, T. (2007). Interparticle coupling effect on the surface plasmon resonance 522

of gold nanoparticles: From theory to applications. Chemical Reviews, 107, 4797–4862. 523

DOI: 10.1021/cr0680282. 524

Gloaguen, F., Léger, J. M., Lamy, C., Marmann, A., Stimming, U., & Vogel, R. (1999). 525

Platinum electrodeposition on graphite: electrochemical study and STM imaging. 526

Electrochimica Acta, 44, 1805–1816. DOI: 10.1016/s0013-4686(98)00332-6. 527

Gorton, L., Lindgren, A., Larsson, T., Munteanu, F. D., Ruzgas, T., & Gazaryan, I. (1999). 528

Direct electron transfer between heme-containing enzymes and electrodes as basis for third 529

generation biosensors. Analytica Chimica Acta, 400, 91–108. DOI: 10.1016/s0003-530

2670(99)00610-8. 531

17

Haghighi, B., & Tabrizi, M. A. (2013). Direct electron transfer from glucose oxidase 532

immobilized on an overoxidized polypyrrole film decorated with Au nanoparticles. Colloids 533

and Surfaces B: Biointerfaces, 103, 566–571. DOI: 10.1016/j.colsurfb.2012.11.010. 534

Holland, J. T., Lau, C., Brozik, S., Atanassov, P., & Banta, S. (2011). Engineering of glucose 535

oxidase for direct electron transfer via site-specific gold nanoparticle conjugation. Journal of 536

the American Chemical Society, 133, 19262–19265. DOI: 10.1021/ja2071237. 537

Horozova, E. G., Dimcheva, N. D., & Jordanova, Z. J. (2000). Study of xanthine oxidase 538

immobilized electrode based on modified graphite. Zeitschrift für Naturforschung C – A 539

Journal of Biosciences, 55c, 60–65. 540

Horozova, E., Dodevska, T., & Dimcheva, N. (2009). Modified graphites: Application to the 541

development of enzyme-based amperometric biosensors. Bioelectrochemistry, 74, 260–264. 542

DOI: 10.1016/j.bioelechem.2008.09.003. 543

Hsu, S. Y., Bartling, B., Wang, C., Shieu, F. S., & Liu, C. C. (2010). Enzymatic determination 544

of diglyceride using an iridium nano-particle based single use, disposable biosensor. 545

Sensors, 10, 5758–5773. DOI: 10.3390/s100605758. 546

Huang, K. J., Niu, D. J., Liu, X., Wu, Z. W., Fan, Y., Chang, Y. F., & Wu, Y. Y. (2011). 547

Direct electrochemistry of catalase at amine-functionalized graphene/gold nanoparticles 548

composite film for hydrogen peroxide sensor. Electrochimica Acta, 56, 2947–2953. DOI: 549

10.1016/j.electacta.2010.12.094. 550

Ilias, S. H., Kok, K. Y., Ng, I. K., & Saidin, N. U. (2013). Electrochemical synthesis and 551

characterization of palladium nanostructures. Journal of Physics: Conference Series, 431. 552

012003. DOI: 10.1088/1742-6596/431/1/012003. 553

Ishikawa, H., Ida, T., & Kimura, K. (1996). Plasmon absorption of gold nanoparticles and 554

their morphologies observed by AFM. Surface Review and Letters, 3, 1153–1156. DOI: 555

10.1142/s0218625x96002060. 556

IUPAC (2005–2014). IUPAC compendium of chemical terminology – gold book. Version 557

2.3.3 (2014-02-24). Research Triangle Park, NC, USA International Union of Pure and 558

Applied Chemistry. DOI: 10.1351/goldbook. 559

Jäckel, F., & Feldmann, J. (2012). Surface-enhanced Raman scattering using complex-shaped 560

metal nanostructures In T. K. Sau & A. L. Rogach (Eds.), Complex-shaped metal 561

nanoparticles: Bottom-up syntheses and applications (Chapter 13, pp. 429–454). Weinheim, 562

Germany: Wiley-VCH. DOI: 10.1002/9783527652570.ch13. 563

Jia, W., Su, L., & Lei, Y. (2011). Pt nanoflower/polyaniline composite nanofibers based urea 564

biosensor. Biosensors and Bioelectronics, 30, 158–164. DOI: 10.1016/j.bios.2011.09.006. 565

18

Karimi, S., Ghourchian, H., Rahimi, P., & Rafiee-Pour, H. A. (2012). A nanocomposite based 566

biosensor for cholesterol determination. Analytical Methods, 4, 3225–3231. DOI: 567

10.1039/c2ay25826a. 568

Karyakin, A. A. (2001). Prussian blue and its analogues: Electrochemistry and analytical 569

applications. Electroanalysis, 13, 813–819. DOI: 10.1002/1521-570

4109(200106)13:10<813::AID-ELAN813>3.0.CO;2-Z. 571

Kirwan, S. M., Rocchitta, G., McMahon, C. P., Craig, J. D., Killoran, S. J., O’Brien, K. B., 572

Serra, P. A., Lowry, J. P., & O’Neill, R. D. (2007). Modifications of poly(o-573

phenylenediamine) permselective layer on Pt-Ir for biosensor application in neurochemical 574

monitoring. Sensors, 7, 420–437. DOI: 10.3390/s7040420. 575

Krasnansky, R., Yamamura, S., Thomas, J. K., & Dellaguardia, R. (1991). Photoproduction of 576

gold colloids and films. Langmuir, 7, 2881–2886. DOI: 10.1021/la00060a003. 577

Krikstolaityte, V., Barrantes, A., Ramanavicius, A., Arnebrant, T., Shleev, S., & Ruzgas, T. 578

(2014). Bioelectrocatalytic reduction of oxygen at gold nanoparticles modified with laccase. 579

Bioelectrochemistry, 95, 1–6. DOI: 10.1016/j.bioelechem.2013.09.004. 580

Lim, S. H., Wei, J., Lin, J., Li, Q., & You, J. K. (2005). A glucose biosensor based on 581

electrodeposition of palladium nanoparticles and glucose oxidase onto Nafion-solubilized 582

carbon nanotube electrode. Biosensors and Bioelectronics, 20, 2341–2346. DOI: 583

10.1016/j.bios.2004.08.005. 584

Lima, F. H. B., Profeti, D., Chatenet, M., Riello, D., Ticianelli, E. A., & Gonzalez, E. R. 585

(2010). Electro-oxidation of ethanol on Rh/Pt and Ru/Rh/Pt sub-monolayers deposited on 586

Au/C nanoparticles. Electrocatalysis, 1, 72–82. DOI: 10.1007/s12678-010-0014-1. 587

Liu, S., & Ju, H. (2003). Electrocatalysis via direct electrochemistry of myoglobin 588

immobilized on colloidal gold nanoparticles. Electroanalysis, 15, 1488–1493. DOI: 589

10.1002/elan.200302722. 590

Liu, S., Leech, D., & Ju, H. (2003). Application of colloidal gold in protein immobilization, 591

electron transfer, and biosensing. Analytical Letters, 36, 1–19. DOI: 10.1081/al-120017740. 592

Liu, G., & Lin, Y. (2006). Biosensor based on self-assembling acetylcholinesterase on carbon 593

nanotubes for flow injection/amperometric detection of organophosphate pesticides and 594

nerve agents. Analytical Chemistry, 78, 835–843. DOI: 10.1021/ac051559q. 595

Liu, A. R., & Huang, S. M. (2012). A glucose biosensor based on direct electrochemistry of 596

glucose oxidase immobilized onto platinum nanoparticles modified graphene electrode. 597

Science China - Physics, Mechanics & Astronomy, 55, 1163–1167. DOI: 10.1007/s11433-598

012-4782-x. 599

19

Luo, X., Morrin, A., Killard, A. J., & Smyth, M. R. (2006). Application of nanoparticles in 600

electrochemical sensors and biosensors. Electroanalysis, 18, 319–326. DOI: 601

10.1002/elan.200503415. 602

Ma, Y., Di, J., Yan, X., Zhao, M., Lu, Z., & Tu, Y. (2009). Direct electrodeposition of gold 603

nanoparticles on indium tin oxide surface and its application. Biosensors and Bioelectronics, 604

24, 1480–1483. DOI: 10.1016/j.bios.2008.10.007. 605

Mano, N., & Edembe, L. (2013). Bilirubin oxidases in bioelectrochemistry: Features and 606

recent findings. Biosensors and Bioelectronics, 50, 478–485. DOI: 607

10.1016/j.bios.2013.07.014. 608

Mao, S., Long, Y., Li, W., Tu, Y., & Deng, A. (2013). Core–shell structured Ag@C for direct 609

electrochemistry and hydrogen peroxide biosensor applications. Biosensors and 610

Bioelectronics, 48, 258–262. DOI: 10.1016/j.bios.2013.04.026. 611

Marinov, I., Ivanov, Y., Gabrovska, K., & Godjevargova, T. (2010). Amperometric 612

acetylthiocholine sensor based on acetylcholinesterase immobilized on nanostructured 613

polymer membrane containing gold nanoparticles. Journal of Molecular Catalysis B: 614

Enzymatic, 62, 67–75. DOI: 10.1016/j.molcatb.2009.09.005. 615

Martorell, D., Céspedes, F., Martínez-Fàbregas, E., & Alegret, S. (1994). Amperometric 616

determination of pesticides using a biosensor based on a polishable graphite-epoxy 617

biocomposite. Analytica Chimica Acta, 290, 343–348. DOI: 10.1016/0003-2670(94)80121-618

5. 619

Marty, J. L., Mionetto, N., & Rouillon, R. (1992). Entrapped enzymes in photocrosslinkable 620

gel for enzyme electrodes. Analytical Letters, 25, 1389–1398. DOI: 621

10.1080/00032719208017123. 622

Marty, J. L., Mionetto, N., Lacorte, S., & Barceló, D. (1995). Validation of an enzymatic 623

biosensor with various liquid chromatographic techniques for determining 624

organophosphorus pesticides and carbaryl in freeze-dried waters. Analytica Chimica Acta, 625

311, 265–271. DOI: 10.1016/0003-2670(94)00617-u. 626

Mizsei, J., Sipilä, P., & Lantto, V. (1998). Structural studies of sputtered noble metal catalysts 627

on oxide surfaces. Sensors and Actuators B: Chemical, 47, 139–144. DOI: 10.1016/s0925-628

4005(98)00015-x. 629

Mizukoshi, Y., Okitsu, K., Maeda, Y., Yamamoto, T. A., Oshima, R., & Nagata, Y. (1997). 630

Sonochemical preparation of bimetallic nanoparticles of gold/palladium in aqueous solution. 631

The Journal of Physical Chemistry B, 101, 7033–7037. DOI: 10.1021/jp9638090. 632

20

Montornes, J. M., Vreeke, M. S., & Katakis, I. (2008). Glucose biosensors. In P. N. Bartlett 633

(Ed.), Bioelectrochemistry: Fundamentals, experimental techniques and applications 634

(Chapter 5, pp. 199–217). Chichester, UK: Wiley. DOI: 10.1002/9780470753842.ch5. 635

Nagaiah, T. C., Schäfer, D., Schuhmann, W., & Dimcheva, N. (2013). Electrochemically 636

deposited Pd-Pt and Pd-Au codeposits on graphite electrodes for electrocatalytic H2O2 637

reduction. Analytical Chemistry, 85, 7897–7903. DOI: 10.1021/ac401317y. 638

Olivia, H., Sarada, B. V., Honda, K., & Fujishima, A. (2004). Continuous glucose monitoring 639

using enzyme-immobilized platinized diamond microfiber electrodes. Electrochimica Acta, 640

49, 2069–2076. DOI: 10.1016/j.electacta.2003.10.026. 641

Palanisamy, S., Karuppiah, C., & Chen, S. M. (2014). Direct electrochemistry and 642

electrocatalysis of glucose oxidase immobilized on reduced graphene oxide and silver 643

nanoparticles nanocomposite modified electrode. Colloids and Surfaces B: Biointerfaces, 644

114, 164–169. DOI: 10.1016/j.colsurfb.2013.10.006. 645

Pandey, P. C., Upadhyay, S., Pathak, H. C., Pandey, C. M. D., & Tiwari, I. (2000). 646

Acetylthiocholine/acetylcholine and thiocholine/choline electrochemical biosensors/sensors 647

based on an organically modified sol–gel glass enzyme reactor and graphite paste electrode. 648

Sensors and Actuators, B: Chemical, 62, 109–116. DOI: 10.1016/s0925-4005(99)00367-6. 649

Pedrosa, V. A., Caetano, J., Machado, S. A. S., Freire, R. S., & Bertotti, M. (2007). 650

Acetylcholinesterase immobilization on 3-mercaptopropionic acid self assembled monolayer 651

for determination of pesticides. Electroanalysis, 19, 1415–1420. DOI: 652

10.1002/elan.200703872. 653

Pingarrón, J. M., Yáñez-Sedeño, P., & González-Cortés, A. (2008). Gold nanoparticle-based 654

electrochemical biosensors. Electrochimica Acta, 53, 5848–5866. DOI: 655

10.1016/j.electacta.2008.03.005. 656

Pita, M., Gutierrez-Sanchez, C., Toscano, M. D., Shleev, S., & De Lacey, A. L. (2013). 657

Oxygen biosensor based on bilirubin oxidase immobilized on a nanostructured gold 658

electrode. Bioelectrochemistry, 94, 69–74. DOI: 10.1016/j.bioelechem.2013.07.001. 659

Plyasova, L. M., Molina, I. Y., Gavrilov, A. N., Cherepanova, S. V., Cherstiouk, O. V., 660

Rudina, N. A., Savinova, E. R., & Tsirlina, G. A. (2006). Electrodeposited platinum 661

revisited: Tuning nanostructure via the deposition potential. Electrochimica Acta, 51, 4477–662

4488. DOI: 10.1016/j.electacta.2005.12.027. 663

Qiaocui, S., Tuzhi, P., Yunu, Z., & Yang, C. F. (2005). An electrochemical biosensor with 664

cholesterol oxidase/sol-gel film on a nanoplatinum/carbon nanotube electrode. 665

Electroanalysis, 17, 857–861. DOI: 10.1002/elan.200403162. 666

21

Quinn, B. M., Dekker, C., & Lemay, S. G. (2005). Electrodeposition of noble metal 667

nanoparticles on carbon nanotubes. Journal of the American Chemical Society, 127, 6146–668

6147. DOI: 10.1021/ja0508828. 669

Ragupathy, D., Gopalan, A. I., & Lee, K. P. (2009). Synergistic contributions of multiwall 670

carbon nanotubes and gold nanoparticles in a chitosan–ionic liquid matrix towards improved 671

performance for a glucose sensor. Electrochemistry Communications, 11, 397–401. DOI: 672

10.1016/j.elecom.2008.11.048. 673

Rahman, M. A., Noh, H. B., & Shim, Y. B. (2008). Direct electrochemistry of laccase 674

immobilized on Au nanoparticles encapsulated-dendrimer bonded conducting polymer: 675

Application for a catechin sensor. Analytical Chemistry, 80, 8020–8027. DOI: 676

10.1021/ac801033s. 677

Ramos, S. G., Moreno, M. S., Andreasen, G. A., & Triaca, W. E. (2010). Facetted platinum 678

electrocatalysts for electrochemical energy converters. International Journal of Hydrogen 679

Energy, 35, 5925–5929. DOI: 10.1016/j.ijhydene.2009.12.106. 680

Sau, T. K., & Rogach, A. L. (2012). Colloidal synthesis of noble metal nanoparticles of 681

complex morphologies. In T. K. Sau, & A. L. Rogach (Eds.), Complex-shaped metal 682

nanoparticles: Bottom-up syntheses and applications (Chapter 1, pp. 7–90). Weinheim, 683

Germany: Wiley-VCH. 10.1002/9783527652570.ch1. 684

Sekol, R. C., Li, X., Cohen, P., Doubek, G., Carmo, M., & Taylor, A. D. (2013). Silver 685

palladium core–shell electrocatalyst supported on MWNTs for ORR in alkaline media. 686

Applied Catalysis B: Environmental, 138–139, 285–293. DOI: 687

10.1016/j.apcatb.2013.02.054. 688

Shao, M., Odell, J. H., Choi, S. I., & Xia, Y. (2013). Electrochemical surface area 689

measurements of platinum- and palladium-based nanoparticles. Electrochemistry 690

Communications, 31, 46–48. DOI: 10.1016/j.elecom.2013.03.011. 691

Shulga, O., & Kirchhoff, J. R. (2007). An acetylcholinesterase enzyme electrode stabilized by 692

an electrodeposited gold nanoparticle layer. Electrochemistry Communications, 9, 935–940. 693

DOI: 10.1016/j.elecom.2006.11.021. 694

Singh, M., Verma, N., Garg, A. K., & Redhu, N. (2008). Urea biosensors. Sensors and 695

Actuators B: Chemical, 134, 345–351. DOI: 10.1016/j.snb.2008.04.025. 696

Somerset, V., Baker, P., & Iwuoha, E. (2009). Mercaptobenzothiazole-on-gold organic phase 697

biosensor systems: 1. Enhanced organosphosphate pesticide determination. Journal of 698

Environmental Science and Health, Part B: Pesticides, Food Contaminants, and 699

Agricultural Wastes, 44, 164–178. DOI: 10.1080/03601230802599092. 700

22

Stoytcheva, M., Sharkova, V., & Magnin, J. P. (1998). Electrochemical approach in studying 701

the inactivation of immobilized acetylcholinesterase by arsenate(III). Electroanalysis, 10, 702

994–998. DOI: 10.1002/(SICI)1521-4109(199810)10:14<994::AID-ELAN994>3.0.CO;2-0. 703

Tang, J., Tian, X. C., Pang, W. H., Liu, Y. Q., & Lin, J. H. (2012). Codeposition of AuPd 704

bimetallic nanoparticles on to ITO and their electrocatalytic properties for ethanol oxidation. 705

Electrochimica Acta, 81, 8–13. DOI: 10.1016/j.electacta.2012.07.048. 706

Terauchi, S., Koshizaki, N., & Umehara, H. (1995). Fabrication of Au nanoparticles by radio-707

frequency magnetron sputtering. Nanostructured Materials, 5, 71–78. DOI: 10.1016/0965-708

9773(95)00011-3. 709

Tiwari, A., Aryal, S., Pilla, S., & Gong, S. (2009). An amperometric urea biosensor based on 710

covalently immobilized urease on an electrode made of hyperbranched polyester 711

functionalized gold nanoparticles. Talanta, 78, 1401–1407. DOI: 712

10.1016/j.talanta.2009.02.038. 713

Turkevich, J., Stevenson, P. C., & Hillier, J. (1951). A study of the nucleation and growth 714

processes in the synthesis of colloidal gold. Discussions of the Faraday Society, 11, 55–75. 715

DOI: 10.1039/df9511100055. 716

Turkevich, J., Stevenson, P. C., & Hillier, J. (1953). The formation of colloidal gold. The 717

Journal of Physical Chemistry, 57, 670–673. DOI: 10.1021/j150508a015. 718

Turkevich, J., & Kim, G. (1970). Palladium: Preparation and catalytic properties of particles 719

of uniform size. Science, 169, 873–879. 720

Velichkova, Y., Ivanov, Y., Marinov, I., Ramesh, R., Kamini, N. R., Dimcheva, N., 721

Horozova, E., & Godjevargova, T. (2011). Amperometric electrode for determination of urea 722

using electrodeposited rhodium and immobilized urease. Journal of Molecular Catalysis B: 723

Enzymatic, 69, 168–175. DOI: 10.1016/j.molcatb.2011.01.015. 724

Wang, L., Mao, W., Ni, D., Di, J., Wu, Y., & Tu, Y. (2008). Direct electrodeposition of gold 725

nanoparticles onto indium/tin oxide film coated glass and its application for electrochemical 726

biosensor. Electrochemistry Communications, 10, 673–676. DOI: 727

10.1016/j.elecom.2008.02.009. 728

Wang, J., Wang, L., Di, J., & Tu, Y. (2009). Electrodeposition of gold nanoparticles on 729

indium/tin oxide electrode for fabrication of a disposable hydrogen peroxide biosensor. 730

Talanta, 77, 1454–1459. DOI: 10.1016/j.talanta.2008.09.034. 731

Whyman, R. (1996). Gold nanoparticles: A renaissance in gold chemistry. Gold Bulletin, 29, 732

11–15. DOI: 10.1007/bf03214736. 733

23

Wilson, R. (2008). The use of gold nanoparticles in diagnostics and detection. Chemical 734

Society Reviews, 37, 2028–2045. DOI: 10.1039/b712179m. 735

Xu, G., Adeloju, S. B., Wu, Y., & Zhang, X. (2012). Modification of polypyrrole nanowires 736

array with platinum nanoparticles and glucose oxidase for fabrication of a novel glucose 737

biosensor. Analytica Chimica Acta, 755, 100–107. DOI: 10.1016/j.aca.2012.09.037. 738

Xue, B., Chen, P., Hong, Q., Lin, J., & Tan, K. L. (2001). Growth of Pd, Pt, Ag and Au 739

nanoparticles on carbon nanotubes. Journal of Materials Chemistry, 11, 2378–2381. DOI: 740

10.1039/b100618p. 741

Yang, M., Qu, F., Lu, Y., He, Y., Shen, G., & Yu, R. (2006). Platinum nanowire 742

nanoelectrode array for the fabrication of biosensors. Biomaterials, 27, 5944–5950. DOI: 743

10.1016/j.biomaterials.2006.08.014. 744

Yu, Y., Chen, Z., He, S., Zhang, B., Li, X., & Yao, M. (2014). Direct electron transfer of 745

glucose oxidase and biosensing for glucose based on PDDA-capped gold nanoparticle 746

modified graphene/multi-walled carbon nanotubes electrode. Biosensors and Bioelectronics, 747

52, 147–152. DOI: 10.1016/j.bios.2013.08.043. 748

Zhang, X., Cao, Y., Yu, S., Yang, F., & Xi, P. (2013). An electrochemical biosensor for 749

ascorbic acid based on carbon-supported PdNi nanoparticles. Biosensors and Bioelectronics, 750

44, 183–190. DOI: 10.1016/j.bios.2013.01.020. 751

Zhao, S., Zhang, K., Sun, Y., & Sun, C. (2006). Hemoglobin/colloidal silver nanoparticles 752

immobilized in titania sol–gel film on glassy carbon electrode: Direct electrochemistry and 753

electrocatalysis. Bioelectrochemistry, 69, 10–15. DOI: 10.1016/j.bioelechem.2005.09.004. 754

Zhao, K., Zhuang, S., Chang, Z., Songm, H., Dai, L., He, P., & Fang, Y. (2007). 755

Amperometric glucose biosensor based on platinum nanoparticles combined aligned carbon 756

nanotubes electrode. Electroanalysis, 19, 1069–1074. DOI: 10.1002/elan.200603823. 757

Zhu, L., Xu, L., Tan, L., Tan, H., Yang, S., & Yao, S. (2013). Direct electrochemistry of 758

cholesterol oxidase immobilized on gold nanoparticles-decorated multiwalled carbon 759

nanotubes and cholesterol sensing. Talanta, 106, 192–199. DOI: 760

10.1016/j.talanta.2012.12.036. 761

762

763

764

Table 1. First-generation amperometric biosensors with H2O2 electrocatalytic reduction as response-generating process 765

766

Analyte Immobilised enzyme Electrode modifier Ea/V Reference

Glucose Glucose oxidase Pd/Pt alloys (MPs) 0 Dimcheva et al. (2002b); Dodevska et al. (2006); Nagaiah et al. (2013)

Pt nanowires 0 Yang et al. (2006)

Fe NPs −0.1 Comba et al. (2010)

Pt NPs −0.1 Choi et al. (2007)

Pd NPs n.a. Lim et al. (2005)

Cholesterol Cholesterol oxidase CNT + Pt NPs n.a. Qiaocui et al. (2005)

Glutamate Glutamate oxidase Au NPs 0.135 Batra and Pundir (2013)

L-Lysine L-Lysine oxidase Au NPs + Pt NPs 0.2 Chauhan et al. (2013)

Xanthine Xanthine oxidase Pd/Pt alloys (MPs) −0.05 Dodevska et al. (2010); Horozova et al. (2009)

a) All operating potentials (E) were measured against Ag/AgCl; n.a. means not available. 767

768