Next Article in Journal
An Automatic and Novel SAR Image Registration Algorithm: A Case Study of the Chinese GF-3 Satellite
Next Article in Special Issue
Electrochemical Sensor for Bilirubin Detection Using Screen Printed Electrodes Functionalized with Carbon Nanotubes and Graphene
Previous Article in Journal
Towards an Iterated Game Model with Multiple Adversaries in Smart-World Systems
Previous Article in Special Issue
Fabrication of a Textile-Based Wearable Blood Leakage Sensor Using Screen-Offset Printing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Integrated Affinity Biosensing Platforms on Screen-Printed Electrodes Electrografted with Diazonium Salts

by
Paloma Yáñez-Sedeño
*,
Susana Campuzano
and
José M. Pingarrón
Departamento de Química Analítica, Facultad de CC. Químicas, Universidad Complutense de Madrid, E-28040 Madrid, Spain
*
Author to whom correspondence should be addressed.
Sensors 2018, 18(2), 675; https://doi.org/10.3390/s18020675
Submission received: 29 January 2018 / Revised: 15 February 2018 / Accepted: 20 February 2018 / Published: 24 February 2018
(This article belongs to the Special Issue Screen-Printed Electrodes)

Abstract

:
Adequate selection of the electrode surface and the strategies for its modification to enable subsequent immobilization of biomolecules and/or nanomaterials integration play a major role in the performance of electrochemical affinity biosensors. Because of the simplicity, rapidity and versatility, electrografting using diazonium salt reduction is among the most currently used functionalization methods to provide the attachment of an organic layer to a conductive substrate. This particular chemistry has demonstrated to be a powerful tool to covalently immobilize in a stable and reproducible way a wide range of biomolecules or nanomaterials onto different electrode surfaces. Considering the great progress and interesting features arisen in the last years, this paper outlines the potential of diazonium chemistry to prepare single or multianalyte electrochemical affinity biosensors on screen-printed electrodes (SPEs) and points out the existing challenges and future directions in this field.

1. Introduction

Preparation of electrochemical affinity biosensors by immobilization of appropriate bioreceptors onto an electrode platform implies the selection of the electrode material and the methods for immobilization as key steps to achieve the desired final performance [1,2,3]. Accordingly, electrode surface modification using different physical or chemical strategies to be utilized for further biomolecules immobilization is an extensively studied subject in the literature. Physical adsorption, mainly based on the electrostatic interaction between the biomolecule and the support surface, is a simple and economic method that does not damage the activity of the biological material [4,5]. However, poor reproducibility and low sensitivity due to leaching of the adsorbed component may be observed. Another physical immobilization involves entrapment of biomolecules into a three-dimensional network of a natural or synthetic gel [6,7,8] or conducting polymer [9,10]. However, small size molecules are difficult to immobilize using this approach because they can be filtered from the matrix and, moreover, this simple immobilization method is not adequate for the preparation of affinity biosensors because the affinity binding can be strongly hindered.
As it is well known, an interesting alternative to physical methods is the covalent coupling of biomolecules to activated surfaces containing functional moieties with binding capacity. Using the most common electrode materials, which include carbon in all forms (glassy carbon, graphite, carbon nanotubes or graphene), and gold, fabrication of self-assembled monolayers (SAMs) of thiols on gold [11,12], preparation of electrodes by carboxyl-confined [13], silanized [14] or aldehyde-derivatives [15,16], and functionalization by means of click-chemistry [17] have been widely explored. Furthermore, aryl radicals generated from the electrochemical reduction of diazonium salts [18] have also been used to modify electrode surfaces. The reductive process of such salts results in the formation of aryl centered radicals covalently attached onto electrode surfaces after the spontaneous elimination of dinitrogen [19]. The so-called grafting method has demonstrated to be an excellent strategy to be used for further immobilization of biomolecules because of the simple preparation and versatility. In fact, conductive and semiconductive surfaces can be modified with a wide range of functional groups in aqueous solution at room temperature and without sophisticated equipment. Other advantages are the reproducibility, uniformity and stability of the covalently attached organic layer on the surface, which have made this method a suitable choice to develop a wide range of attractive electrochemical affinity biosensors by immobilizing proteins, antibodies, nucleic acids and anchoring nanomaterials [5,20,21,22,23]. In addition, the ability to create a diazonium-modified surface by application of an appropriate potential scan or previous modification of the biomolecules with aryl diazonium allows the selective functionalization of closely spaced microelectrode surfaces with different molecules leading to the construction of multianalyte biosensors [24,25,26]. Electrografting as a method for surface modification was extensively reviewed by Bélanger and Pinson [27]. More recently, applications in electroanalysis were also reviewed [28].
While the use of carbon surfaces (glassy carbon, screen-printed carbon electrodes or SPCEs, graphite, graphene, carbon nanotubes and diamond) has dominated this field, a wide range of other conducting materials such as metals, silicon and indium tin oxide (ITO) have been also employed for diazonium grafting modification with a second to minutes scale reaction time [29,30]. Moreover, a surface can be functionalized with either one or multiple types of aryldiazonium salts to create single or mixed layers, respectively. Indeed, modification with mixed layers is especially relevant in applications where the interface is required to perform multiple functions, and hence different chemical species should be incorporated into the layer, or when the bioplatform should operate in complex biological fluids and a mixture of antifouling and bio-recognition components are needed [30]. However, it is worth to mention that formation of mixed layers using this chemistry is still in the infancy and some challenges remain to be addressed to precisely control their structure to the level reached using alkanethiol mixed SAMs. For example, the unselective reactivity of two aryldiazonium cations makes it difficult to control the surface composition and the highly reactive radicals can grow both on the bare electrode surface and on an as-deposited layer resulting in multilayer formation.
Regarding grafting onto SPEs, it should be noted that the use of these electrodes is becoming more and more widespread for the preparation of electrochemical affinity biosensors because their disposable nature which simplifies their use, avoids problems with electrode fouling and makes it easier to perform decentralized assays [31,32]. Moreover, the use of screen-printing technology also offers other general interesting advantages such as the utilization of small sample volumes, and the mass production of inexpensive and robust strip solid electrodes suitable to be fabricated in miniaturized sizes, with different materials, in diverse geometries and multiplexed formats [5,33]. Among the different strategies, immobilization of biorecognition components and nanomaterials onto these planar and disposable surfaces pre-functionalized with electrografted diazonium salts has been widely reported in the latest years to develop competitive integrated affinity biosensing platforms (Figure 1). The use of non-disposable electrodes is less attractive for the application of this methodology since once the aryl radical is grafted onto the electrode surface, a regeneration protocol should be implemented according with the employed electrode material for subsequent reutilization. Moreover, one of the main current trends in the development of biosensors is the ability to allow multiplexed detection of several biomarkers with a single sensing platform. However, the selective functionalization of different electrodes remains an obstacle to produce massively multiplexed sensors. In this context, the combination of the screen-printed electrodes ability for multiplexing with the versatility, simplicity, speed and capability offered by electrografting methodology to individually functionalize each working electrode opens up substantial possibilities for the massive fabrication of integrated electroanalytical bioplatforms for the simultaneous determination of biomarkers of the same or different molecular level. These very interesting advantages have led to the fact that in recent years most of electrochemical biosensors configurations using electrodes modified by grafting of diazonium salts have been implemented with screen-printed electrodes.
Considering the aforementioned aspects, this review article discusses the basic concepts of the chemistry involved in aryldiazonium grafting and provides an updated overview of the main strategies reported so far to develop affinity biosensing platforms at SPEs modified with bioreceptors or nanomaterials through representative selected examples. Highlighted examples are classified and discussed in the next sections according to the SPE material and the type of biomolecule immobilized on the electrografted aryldiazonium salt.

2. Aryl Diazonium Salt Chemistry in Electrochemical Affinity Biosensing

Since in 1992 Pinson and co-workers described the reaction mechanism for the modification of carbon electrodes by electrochemical reduction of one or multiple types of aryldiazonium salts (see Scheme 1), this chemistry have demonstrated to be an efficient way to introduce many types of functional groups onto a variety of surfaces (carbon surfaces, metals, silicon and indium tin oxide) with a reaction time scale of seconds to minutes [30].
However, the use of aryl diazonium salt chemistry for the preparation of biosensors may show some drawbacks: (i) the complicated synthesis of new aryl diazonium salts because of the reactivity of the head group; (ii) the fabrication of only single component layers; (iii) the presence of side reactions which compromises surface chemistry control by creating multilayers instead of monolayers; (iv) the need for connection of each electrode to a source of potential for modification; (v) the absence of established methods to apply different chemistries at well-defined surface locations; (vi) when used carbon electrodes, the natural heterogeneity of such surfaces may limit the reproducibility of devices. Despite these problems, the great number of advances made in recent years has provided potential solutions to all these disadvantages and greatly broaden the utility of this powerful surface chemistry for biosensing applications. It is worth to note, for example, that using diazonium salt chemistry, the density of the deposited organic layer on the electrode surface can be controlled by choosing the electrografting protocol and time [20]. Moreover, as indicated in the Introduction section, this chemistry has been used so far for the modification of a variety of conducting surfaces: carbon electrodes (glassy carbon, graphite, screen-printed carbon electrodes, carbon nanotubes [34,35], diamond [36,37,38]), but also metals, silicon and ITO electrodes. In this context, it must be considered that differences existing in the surface chemistry on all these materials have also been investigated.
The in situ formation of aryldiazonium salts has significantly reduced the synthetic burden with no significant differences between the prepared films and the corresponding pre-made salts [39,40]. This is a particularly relevant achievement for the preparation of biosensors since the diazotization step, followed by purification, is a step fraught with problems for some biomolecules. In addition, the great progress made in the preparation of sensing interfaces with mixed layers [30,41,42,43,44] has provided useful insights regarding factors determining the ratio of the two components on the surface compared with the assembly solution [45]. These studies demonstrated, for example, that the concentration of the cation most easily reduced in the electrografted surface is larger than its relative concentration in the mixed solution used for the deposition. In the specific case of electrochemical affinity biosensors, the modification with mixed layers where one component is further modified with bio-recognition species (e.g., antibodies, nucleic acid) and the other component is used to space the attached biorecognition species attached, to immobilize redox species or to minimize nonspecific adsorptions, is particularly relevant [30].
New strategies have been developed for preparing monolayers derived from aryl diazonium salts where the 3- and 5-positions of surface bound diazonium salts are blocked [46] as well as from molecules with a large sacrificial component attached to the 4-position so that, once the layer is formed, is cleaved from the surface leaving a near monolayer. This later was achieved using aryl alkyl hydrazones where hydrolysis of the surface bound molecule leaves a benzaldehyde which is compatible with subsequent coupling of recognition species or a disulfide leaving a surface thiol. Moreover, spontaneous adsorption of aryldiazonium salts and several different strategies which allow simple patterning of the sensing substrate have been reported also on a range of different surfaces [47,48,49,50,51,52]. This is especially important for sensors as it is much more compatible with the bulk manufacture of devices.
Although the application of aryldiazonium salts for the preparation of biosensors was recognized soon after the first report on electrode modification with such derivatives [53], the great explosion of this chemistry for biosensing started ten years ago. Since then, the number of approaches has grown rapidly and has been used for the immobilization of enzymes [54,55,56], antibodies [5,19,20,24,29,33,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74], DNA [24,65,66,67], and PNA probes [68], aptamers [21,69,70], peptides [71,72,73], and even whole cells [74]. Furthermore, the aryldiazonium salt derived layers have demonstrated also very interesting advantages for attaching nanomaterials and other components, to prepare more sophisticated (bio)sensing layers [75,76,77,78,79,80,81,82,83,84].
Table 1 summarizes the methods and analytical characteristics of some selected configurations of electrochemical affinity biosensors using SPEs modified by grafting with aryldiazonium salts, which have been applied to clinical and food samples. In the following sections, we discuss the relevant aspects of such designs.

2.1. Electrochemical Immunosensors Involving Aryl Diazonium Salt Chemistry onto Screen-Printed Electrodes

Strategies involving conjugation of capture antibodies with 4-carboxy methylaniline followed by diazotization to the respective diazonium salt modified antibody [23,24] (Figure 2), or with 4-carboxyphenyldiazonium salt to avoid the diazotization step in the presence of the protein, have been reported. Due to the high number of primary amines on the exterior, the proteins are indeed modified with several diazonium species.
More conventional strategies involve antibody immobilization onto electrodes previously modified with the diazonium salt. These strategies include: (i) modification of SPCEs [5,58,68,69,70,85], carbon nanomaterial-modified SPEs [33], graphene-modified SPEs (GSPEs) [59,60], carbon nanofiber-modified SPCEs (CNF SPCEs) [61] and reduced graphene oxide (rGO)-SPCEs [62,63] with 4-carboxyphenyl diazonium salt; (ii) modification of GSPEs with p-nitroaniline diazonium salt [20]; (iii) modification of graphite-based SPEs with p-aminophenyl acetic acid [58]; (iv) modification of glassy carbon electrodes (GCEs) with a mixed layer of oligo(ethylene glycol) species (OEG) and oligo(phenylethynylene) molecular wires (MW) [44].
In these immunosensing approaches, antibodies were covalently immobilized onto the electrografted surface using carbodiimide chemistry [5,33,58,59,60,61,62,63], glutaraldehyde [20], via an amidic bond [20] and through aminophenylboronic acid [85] or biotin-streptavidin interaction [44].
Electrochemical immunosensors have gained prominence for protein detection in recent years due to their sensitivity, selectivity, portability, simplicity, low cost, fast response time and compatibility with multiplexed determination and miniaturization [2,33,61]. The key points for the design of an electrochemical immunosensor are the choice of the electrode support and the immobilization of the antibody or the analyte onto its surface. As the following examples show, the chemistry of diazonium salts offers a very interesting alternative for the manufacture of immunosensors onto SPEs with attractive performance.
A label-free voltammetric immunosensor for the detection of β-lactoglobulin was developed by electrochemical modification of GSPEs with 4-nitrophenyldiazonium cations. The diazonium cations were prepared by the diazotization of 4-nitroaniline. Subsequently, the in situ generated diazonium cations were covalently grafted through an electrochemical reduction step on the graphene electrode surface, followed by the electrochemical reduction of the terminal nitro to amine groups used, upon their activation with glutaraldehyde for covalent immobilization of β-lactoglobulin antibodies [20]. By monitoring the decrease in the DPV reduction peak current of [Fe(CN)6]3−/4− in the presence of β-lactoglobulin, the immunosensor enabled a dynamic range from 1 pg mL−1 to 100 ng mL−1 and a LOD of 0.85 pg mL−1. Moreover, the immunosensor was applied to the analysis of different samples (cake, cheese snacks and sweet biscuits). Same authors proposed the use of the same GSPEs modified by cyclic voltammetry reduction of in situ generated 4-carboxyphenyldiazonium salt to develop voltammetric immunosensors for the sensitive detection of okadaic acid (OA) [59] and ovalbumin (OVA) [60]. The antibody was covalently immobilized onto the electrografted GSPE via carbodiimide chemistry. The immunosensor for OA determination was based on a competitive assay between OA and a fixed concentration of OA–ovalbumin conjugate (OA–OVA) for the immobilized antibodies and on measuring the [Fe(CN)6]3−/4− reduction peak current obtained by SWV (Figure 3). The higher the concentration of OA, the larger the SWV signal due to the lower amount of immobilized OA-OVA. The method allowed reaching a LOD of 19 ng L−1, a linear determination range up to ~5000 ng L−1 and successful applicability for the analysis of spiked shellfish tissue extracts and certified reference mussel samples. The method developed for OVA, based on the decrease in the reduction peak of [Fe(CN)6]3−/4− measured by DPV in the presence of the antigen, demonstrated to be appropriate for OVA determination in the concentration range from 1 pg mL−1 to 0.5 μg mL−1 with a LOD of 0.83 pg mL−1 and for the analysis of spiked cake extracts.
Hayat et al. [5] developed another electrochemical immunosensor for OA through its covalent immobilization onto SPCEs modified by grafting with 4-carboxyphenyl film followed by terminal carboxylic group activation by N-hydroxysuccinimide (NHS) and N-(3-dimethylamino propyl)-N-ethylcarbodiimide hydrochloride (EDC). Hexamethyldiamine was then covalently bound by one of its terminal amine group to the activated carboxylic group. The carboxyl group of OA, activated by EDC/NHS, was then conjugated to the second terminal amine group on other side of the hexamethyldiamine through amide bond formation. After immobilization of OA, an indirect competitive immunoassay involving the labeling of a specific OA antibody with an alkaline phosphatase (AP)-labeled secondary antibody and DPV detection in the presence of 1-naphtyl phosphate (1-NP) was employed to detect OA. This immunosensor allowed a LOD of 1.44 ng L−1 and showed applicability to the analysis of certified reference mussel samples. One year later, same authors developed an impedimetric immunosensor for OA by covalent immobilization of the specific antibody onto the same carboxyphenyl-modified SPCE [58]. The increase in electron transfer resistance measured by electrochemical impedance spectroscopy (EIS) in the presence of [Fe(CN)6]4−/3− was linearly proportional to the OA concentration in the 0.195–12.5 μg L−1 range, with a LOD of 0.3 μg L−1. The analysis performed in spiked mussel samples demonstrated acceptable recovery percentages.
A label-free electrochemical immunosensor for the detection of porcine serum albumin (pSA) using CNF SPEs with a 4-carboxyphenyl layer electrografted using the 4-carboxyphenyl diazonium salt was proposed by Lim et al. (Figure 4) [61]. Antibodies were covalently immobilized onto the modified electrodes previously activated with EDC/NHS and the increase in the SWV cathodic peak current of anionic redox probe recorded after immunocomplex formation with antibodies, attributed to the strong affinities of serum albumins like pSA towards anions, was used for the detection. This immunosensor exhibited a wide linear range (0.5–500 pg mL−1) and a low LOD (0.5 pg mL−1) in buffer solution. Moreover, the feasibility of this approach for practical application was demonstrated by satisfactory recoveries in spiked undiluted fresh raw pork samples.
Pingarrón´s group proposed the development of immunosensing platforms using SPCEs for the covalent immobilization of capture antibodies through a 4-ABA diazonium salt grafting strategy. Using this strategy, an electrochemical immunosensor for the determination of adrenocorticotropin hormone (ACTH) was reported [19]. The immunoelectrode design involved grafting with 4-ABA followed by using of amino phenylboronic acid for the oriented immobilization of anti-ACTH antibodies onto SPCE-modified electrode surfaces. A competitive immunoassay between the antigen and the biotinylated hormone for the binding sites of the immobilized antibody was performed. The electroanalytical response was generated by using alkaline phosphatase-labelled streptavidin and 1-naphtyl phosphate as the enzyme substrate. The electrochemical oxidation of the enzyme reaction product, 1-naphtol, measured by DPV was employed to monitor the affinity reaction. Under the optimized working conditions, an extremely low detection limit of 18 pg L−1 was obtained and an excellent selectivity against other hormones (cortisol, estradiol, testosterone, progesterone, hGH and prolactin) were achieved. The immunosensor was used to analyze a human serum sample containing a certified amount of ACTH with good results. This strategy was further extended to the construction of a novel dual electrochemical immunosensor for the multiplexed determination of adrenocorticotropin (ACTH) and cortisol onto dual screen-printed carbon electrodes (SPdCEs) [85].
Another configuration of a disposable immunosensing platform for the simultaneous determination of two obesity-related hormones, ghrelin (GHRL) and peptide YY (PYY) was also reported by this group [62,63]. SPdCEs were modified with reduced graphene oxide (rGO) and, after grafting of the diazonium salt of 4-ABA on these modified electrode surfaces, the corresponding capture antibody for each target hormone was covalently immobilized onto the 4-carboxyphenyl moieties of each working electrode via EDC/NHS chemistry (Figure 5). The determination of each hormone was performed by direct competitive immunoassays with the corresponding biotinylated hormones for the immobilized capture antibody. After labeling the attached biotinylated hormones with a polymer of streptavidin-phosphatase (AP-Strep), the DPV signal obtained in the presence of 1-NP was used to monitor the affinity reactions. This dual immunosensing scaffold provided linear current vs. log [hormone] plots extending between 10−3 and 100 ng mL−1 and 10−4 and 10 ng mL−1 for GHRL and PYY, respectively. The usefulness of this approach was also demonstrated by its application for the accurate determination of the analytes in spiked human serum and saliva samples.
SPCEs modified with double-walled carbon nanotubes (DWCNTs) previously functionalized with an aryl diazonium salt by the Bahr and Tour’ method [96] were used for developing an electrochemical immunosensor for adiponectin (APN). DWCNTs were treated with p-ABA in the presence of isoamyl nitrite in N-methyl-2-pyrrolidone (NMP) resulting in the formation of 4-carboxyphenyl-DWCNTs without any electrochemical treatment. The oriented binding of specific antibodies toward APN was accomplished by using the metallic-complex chelating polymer Mix&GoTM (Figure 6). A calibration plot for APN was constructed with a range of linearity extending between 0.05 and 10.0 μg mL−1, and a detection limit of 14.5 ng mL−1. The usefulness of the immunosensor for the analysis of real samples was demonstrated by analyzing human serum from female or male healthy individuals [86].
A simple label-free impedimetric immunosensor for determination of mucin 4 (MUC 4) protein was proposed using graphite SPEs modified with p-aminophenylacetic acid for antibody immobilization via amidic bond [58]. The immunosensor provided a LOD of 330 pg μL−1. Eissa et al. [33] carried out a comparative study of six different carbon nanomaterial-modified electrodes (carbon, graphene (G), GO, single wall carbon nanotube (SWCNT), multi-wall carbon nanotube (MWCNT), and carbon nanofiber (CNF)) which were modified by grafting with 4-carboxyphenyl layers to develop voltammetric immunosensors for the detection of survival motor neuron (SMN) protein. The SMN antibody was covalently immobilized onto the terminal carboxylic moieties on the electrode surfaces through EDC/NHS chemistry. Results showed that the CNF-modified electrode exhibited the best performance for the SMN immunosensor. By measuring the increase in the SWV reduction peak current of [Fe(CN)6]3−/4− in the presence of SMN due to the positive charge of the protein, the CNF-based immunosensor provided a LOD of 0.75 pg mL−1 and feasibility to perform the determination in a spiked blood sample from healthy volunteer.
As it is well known, strategies allowing a high control over the fabrication of sensing interfaces on gold electrodes include mainly the use of SAMs [97,98] and, more recently, also modification with aryldiazonium salts [72]. This latter method provides not only an effective method for immobilization of bioreagents but also the stable incorporation of gold nanoparticles onto gold electrodes [82,83]. Although alkanethiols assembling offers a simply and versatile way to prepare sensing interfaces, their stability is a big issue when the as prepared sensors should operate over long measurement times or their fabrication involves multiple coupling steps. In these cases, the oxidation of alkanethiols and the resultant loss of SAM becomes problematical. In contrast to thiol–Au chemistry, diazonium modification on Au surfaces improves stability in terms of long-term storage in air, potential cycling under acidic conditions, and wider potential window for electrochemical detection methods [24,26]. Various examples of application of this methodology to gold electrode surfaces have been reported in the literature. For example, the direct modification of a gold electrode with aminophenyl groups by electrochemical reduction of in situ generated aminophenyl monodiazonium cations synthesized from p-phenylendiamine and NaNO2 was reported by Lyskawa and Bélanger [99]. Very recently, Phal et al. [100] prepared a gold electrode electrografted with 4-carboxybenzenediazonium for developing a methotrexate (MTX) immunosensor. However, only one work has been found regarding the use of screen printed gold electrodes [89]. In this method, SPAuE was modified with 4-nitrophenyl groups assembled from 4-nitrophenyl diazonium salt and, then, the nitro groups were electrochemically reduced to amines followed by activation with glutaraldehyde for covalently bind ochratoxin A (OTA) antibodies. A direct competitive-type immunosensor using OTA-HRP was prepared using 3,3′,5,5′-tetramethyl-benzidine (TMB) for the amperometric detection of OTA in a dynamic range up to 60 ng mL−1 with a LOD of 12 ng mL−1.

2.2. Electrochemical Nucleic Acid Biosensors Involving Aryl Diazonium Salt Chemistry onto Screen-Printed Electrodes

Nucleic acid biosensors typically involve the immobilization of a single strand (ss) of a nucleic acid, typically a short oligonucleotide, to detect a complementary strand [101] or nucleic acid duplexes to detect small molecules, such as drugs [102] and of aptamers for a wide variety of (bio)molecules determination [21,69,70]. The immobilization of capture probes on electrode surfaces is thus a crucial step to obtain reliable nucleic acid biosensors [103]. A very attractive route involves covalent attachment of thiol-functionalized oligonucleotides on gold electrodes [104]. Regarding carbon electrode surfaces, although adsorption is one of the simplest techniques to immobilize nucleic acids [105,106,107,108], the resulting multiple-point linkage usually leads to poor hybridization efficiency and the release of the nucleic acids from the surface during the hybridization is a potential problem. To overcome these drawbacks, the covalent attachment of an oligonucleotide monolayer on a chemically-functionalized carbon surface appears as more advantageous because it allows increasing the sensitivity of the assay and the use of more stringent washing conditions to reduce the background signal. A very attractive method to covalently attach oligonucleotides on carbon surfaces relies on surface functionalization with diazonium groups [109,110]. During the coupling reaction, the amine electron-releasing group contained in the aromatic rings of adenine, guanine and cytosine undergo an electrophilic attack by in situ generated diazonium ions [65]. Moreover, as DNA sensors sometimes require reasonably high temperatures to ensure efficient duplexes denaturation or hybridization, the thermal stability of the linkage of the DNA to a substrate using diazonium salts chemistry is especially attractive [76]. The selected examples described below demonstrate the potential of this chemistry for developing electrochemical nucleic acid sensors with very attractive characteristics.
The novelty of diazonium chemistry towards the electroaddressable selective functionalization of a single electrode in an electrode array [24] was successfully explored to develop arrays for multi-determination of different proteins [26,111]. Moreover, dual functionality sensors able to detect simultaneously target DNAs and proteins were proposed also using this chemistry [24,25]. These approaches relied on modification of the electrodes in the array with 4-carboxyphenyl-diazonium salts further activated with EDC/NHS to allow the covalent attachment of biomolecules [24] and on the biomolecules modification with aryl diazonium prior to the addressing [25].
The first example of using aryldiazonium salt-derived layers for the preparation of nucleic acid-based biosensors involved the attachment of the DNA probe onto SPCEs modified with p-aminophenyl using p-nitroanilinodiazonium salt and further converted to diazophenyl functions [65]. Moreover, to reduce nonspecific adsorptions, the interface was modified with a mixed layer by spacing the nitrophenyl moieties using carboxyphenyl molecules. Other binary films of 4-((tri-methylsilyl)ethynyl) benzene (TMSi-Eth-Ar) and p-nitrobenzene (p-NO2-Ar) prepared sequentially by electrografting onto SPCEs were used after deprotection and in the presence of copper(I) catalyst to attach aptamers with azide function by forming a covalent 1,2,3-triazole linkage [21].
Covalent immobilization of amino terminated aptameric [69,70] and PNA [68] probes have been successfully achieved using carbodiimide chemistry onto SPCEs modified via diazonium coupling reaction with 4-aminobenzoic acid. A different DNA immobilization strategy relied on the selective covalent binding between thiols and maleimides [112] and the coupling of the 4-phenylmaleimide diazonium salt to a thiol-terminated oligonucleotide prior surface assembly. Another novel strategy for immobilizing DNA and preparing aryldiazonium salts [113] involved the use of a triazene (a protected aryldiazonium salt) and its activation using dimethyl sulfate to reveal the diazonium moiety. In the DNA electrochemical sensor reported, a unique ferrocene derivative (used as electrochemical reporter) was prepared with a phenyl triazene attached to one of the ferrocene cyclopentadienyl (Cp) ring and a succinimide ester to the other Cp ring. An aminated DNA was further attached to this activated ester.
DNA-sensing platforms for the determination of an amplified herpes virus DNA sequence were prepared by attaching covalently oligonucleotide capture probes onto p-aminophenyl functionalized SPCEs using p-nitroanilinodiazonium salt [65]. The subsequent conversion of the p-aminophenyl groups to diazophenyl moieties provided a convenient and versatile way to covalently link nucleic acids on carbon surfaces. In this approach, the PCR-amplified 406 base-pairs (bp) human cytomegalovirus (HCMV) DNA sequences were detected through a sandwich-type hybridization assay with the immobilized DNA probe and a colloidal gold-labeled detector probe. The hybridization event was followed by measuring the Au(III) ions generated by acid dissolution of the gold metal nanoparticles attached to the hybrids using anodic stripping voltammetry at a screen-printed microband electrode (SPMBE).
Hayat et al. [21] developed a highly sensitive and reusable aptasensor for the impedimetric detection of ochratoxin A (OTA) by covalent immobilization of a specific aptamer onto SPCEs modified with two binary films of diazonium salts via click chemistry. SPCEs were modified sequentially by electrografting of protected layers of 4-((trimethylsilyl)ethynyl) benzene (TMSi-Eth-Ar) and p-nitrobenzene (p-NO2-Ar) by means of electrochemical reduction of their corresponding diazonium salts. After deprotection, by treatment with tetrabutylammonium fluoride (TBAF), and in the presence of copper (I) catalyst, the active ethynyl groups of the film and the azide moiety of the aptamer reacted efficiently forming a covalent 1,2,3-triazole linkage (Figure 7). The increase in the electron-transfer resistance (RCT) values measured by EIS in the presence of [Fe(CN)6]4−/3− was proportional to the OTA concentration between 1.25 and 500 ng L−1 with a LOD of 0.25 ng L−1. The aptasensor could be regenerated 10 times with a mild solution, showed a storage stability of at least 10 days at 4 °C, and was successfully applied to the analysis of spiked beer samples.
The use of an aptamer immobilized on SPCEs through diazonium coupling reaction using 4-ABA was also exploited by Del Valle’s group to develop competitive aptasensors for lysozyme (Lys) determination. One of these approaches involved a direct assay and impedimetric detection providing linearity between 0.025 and 0.8 mM, and a LOD of 25 nM (Figure 8) [69]. Another reported method relied on an aptamer-antibody sandwich assay involving the use of a biotinylated antibody further labeled with avidin-AP. By DPV monitoring of the oxidation signal of 1-naphthol generated by enzymatic hydrolysis of 1-naphthyl phosphate substrate, the method exhibited a wide detection range (5 fM–5 nM) and a LOD 4.3 fM [70]. Both aptasensors were used for the quantification of the target protein in spiked wine samples. Additional claimed advantages of these aptasensors are the use of simple instrumentation, low production cost and rapid response.
SPCEs fabricated using Low Temperature Co-fired Ceramics (LTCC) technology and functionalized with amino-modified PNA probes by electrografting via in situ generated diazonium cations (4-ABA) were proposed for label-free molecular detection of antibiotic resistance by targeting blaNDM, one of the main antimicrobial resistance factors in carbapenem-resistant Enterobacteriaceae. This impedimetric PNA-based biosensor provided a LOD of 200 nM [68]. The excellent molecular detection performance combined with a low cost and accelerated sensor manufacturing and functionalization process (over six times faster than protocols used for gold electrodes) make this platform very attractive for rapid detection of antimicrobial resistance at point-of-care POC with competitive costs.

2.3. Other Electrochemical Biosensors Involving Aryl Diazonium Salt Chemistry onto Screen-Printed Electrodes

Apart from antibodies and nucleic acids, aryl diazonium salt derived layers have also been used for cells immobilization [74,114]. For this purpose, a novel phenyl boronic acid pinacol ester diazonium salt was synthesized so that, after assembly onto the electrode surface, addition of sodium iodate removed the pinacol protecting group to yield a phenylboronic acid functionalized surface. By exploiting the selective reaction of boronic acid with sugars, these scaffolds were used to immobilize yeast cells [74] and murine macrophages (mammalian cells belonging to the immune system) [114]. Claimed advantages of these bioscaffolds include the possibility to release the captured cells from the surface by exposure to fructose due to the competitive reaction of this sugar with the cells for the boronic acid of the functionalized surface. Furthermore, a label free impedimetric aptamer sensor for Salmonella typhimurium (S. typhimurium) detection was reported by Bagheryan et al. [95]. A diazonium-supporting layer was fabricated onto a SPCE followed by immobilization of aminated aptamer (Figure 9). The developed design responded linearly over the 10–108 CFU mL−1 range with a limit of detection of 6 CFU mL−1. Good recoveries were obtained in the application to spiked apple juice.

3. General Considerations, Challenges and Prospects

Electrochemical grafting consisting of covalent modification of carbon surfaces by aryl radicals generated from electrochemical reduction of diazonium salts has demonstrated to facilitate the electron transfer and provide a highly stable binding surface with enhanced properties for selective and controlled immobilization of chemical and biological compounds [62,64]. Advantages of this method include the ease of diazonium preparation, a covalent attachment to the electrode, the speed of the chemistry involved [29] and the ability to modify closely-spaced electrodes with different biological entities (proteins, nucleic acid strands, and peptides) [23,29]. Moreover, the application of aryldiazonium salts for (bio)sensing has recently seen the integration of nanomaterials together with the recognition species into interfaces thus taking advantage of the nanomaterials properties.
As it has been pointed out above, diazonium chemistry-based biosensing platforms include SPCEs [5,58,68,69,70,85], carbon nanomaterial-modified SPEs [33], GrSPEs [59,60], CNF SPEs [55], rGO-SPCEs [62,63] and graphite-based SPEs [64]. Regarding the aryl diazonium substrates, 4-ABA [5,33,58,59,60,61,62,63,68,69,70,85], p-nitroaniline [20] and p-aminophenylacetic acid [64] were mostly used.
Despite the modification of surfaces using aryl diazonium salts has been explored for almost 20 years now, the application of aryl diazonium salts for fabricating biosensing interfaces is still in its infancy. Although variations of aryldiazonium salts mixed layers have been employed for affinity biosensing at conventional electrodes [30], they have been scarcely explored onto disposable electrodes. Therefore, additional efforts should be focused on exploring multicomponent layers and to prepare more stable analogues of sensing interfaces that those developed with another surface chemistry.
However, it is worth to mention that despite these issues, the exciting advances made in the understanding of aryl diazonium salt chemistry, and how to assemble complex layers on surfaces, the range of new synthesized aryldiazonium salts, and the incorporation of nanomaterials are providing aryl diazonium salts derived sensors with the stability and flexibility advantages this chemistry provides. Furthermore, possible drawbacks, relative to other surface chemistries such as that of the alkanethiols, that are in a much more advanced state, are rapidly being resolved which makes aryldiazonium salt chemistry almost the ideal surface chemistry for preparing sensing interfaces.

Acknowledgments

The financial support of the CTQ2015-70023-R and CTQ2015-64402-C2-1-R (Spanish Ministerio de Economía y Competitividad Research Projects) and S2013/MT3029 (NANOAVANSENS Program from the Comunidad de Madrid) are gratefully acknowledged.

Author Contributions

P.Y.-S., S.C. and J.M.P. contributed equally to the writing of this paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kokkinos, C.; Economou, A.; Prodromidis, M.I. Electrochemical immunosensors: Critical survey of different architectures and transduction strategies. Trends Anal. Chem. 2016, 79, 88–105. [Google Scholar] [CrossRef]
  2. Felix, F.S.; Angnes, L. Electrochemical immunosensors—A powerful tool for analytical applications. Biosens. Bioelectron. 2018, 102, 470–478. [Google Scholar] [CrossRef] [PubMed]
  3. Du, Y.; Dong, S. Nucleic acid biosensors: Recent advances and perspectives. Anal. Chem. 2017, 89, 189–215. [Google Scholar] [CrossRef] [PubMed]
  4. Schreiber, A.; Feldbrugge, R.; Key, G.; Glatz, J.F.C.; Spencer, F. An immunosensor based on disposable electrodes for rapid estimation of fatty acid-binding protein, an early marker of myocardial infarction. Biosens. Bioelectron. 1997, 12, 1131–1137. [Google Scholar] [CrossRef]
  5. Hayat, A.; Barthelmebs, L.; Sassolas, A.; Marty, J.-L. An electrochemical immunosensor based on covalent immobilization of okadaic acid onto screen printed carbon electrode via diazotization-coupling reaction. Talanta 2011, 85, 513–518. [Google Scholar] [CrossRef] [PubMed]
  6. Cosnier, S. Biomolecule immobilization on electrode surfaces by entrapment or attachment to electrochemically polymerized films: A review. Biosens. Bioelectron. 1999, 14, 443–456. [Google Scholar] [CrossRef]
  7. Butterfield, D.A.; Colvin, J.; Liu, J.L.; Wang, J.Q.; Bachas, L.; Bhattacharya, D. Electron paramagnetic resonance spin label titration: A novel method to investigate random and site-specific immobilization of enzymes onto polymeric membranes with different properties. Anal. Chim. Acta 2002, 470, 29–36. [Google Scholar] [CrossRef]
  8. Shen, L.; He, P. An electrochemical immunosensor based on agarose hydrogel films for rapid determination of ractopamine. Electrochem. Commun. 2007, 9, 657–662. [Google Scholar] [CrossRef]
  9. Cosnier, S. Recent advances in biological sensors based on electrogenerated polymers: A review. Anal. Lett. 2007, 40, 1260–1279. [Google Scholar] [CrossRef]
  10. Teles, F.R.R.; Fonseca, L.P. Applications of polymers for biomolecule immobilization in electrochemical biosensors. Mater. Sci. Eng. C 2008, 28, 1530–1543. [Google Scholar] [CrossRef]
  11. Nuzzo, R.G.; Allara, D.L. Adsorption of bifunctional organic disulfides on gold surfaces. J. Am. Chem. Soc. 1983, 105, 4481–4483. [Google Scholar] [CrossRef]
  12. Wilson, D.L.; Martin, R.; Hong, S.; Cronin-Golomb, M.; Mirkin, C.; Kaplan, D. Surface organization and nanopatterning of collagen by dip-pen nanolithography. Proc. Natl. Acad. Sci. USA 2001, 98, 13660–13664. [Google Scholar] [CrossRef] [PubMed]
  13. Sassolas, A.; Leca-Bouvier, B.D.; Blum, L.J. DNA biosensors and microarrays. Chem. Rev. 2008, 108, 109–139. [Google Scholar] [CrossRef] [PubMed]
  14. Demers, L.M.; Ginger, D.S.; Park, S.J.; Li, Z.; Chung, S.W.; Mirkin, C.A. Direct patterning of modified oligonucleotides on metals and insulators by dip-pen nanolithography. Science 2002, 296, 1836–1838. [Google Scholar] [CrossRef] [PubMed]
  15. MacBeath, G.; Schreiber, S.L. Printing proteins as microarrays for high-throughput function determination. Science 2000, 289, 1760–1763. [Google Scholar] [PubMed]
  16. Zhu, H.; Bilgin, M.; Bangham, R.; Hall, D.; Casamayor, A.; Bertone, P.; Lan, N.; Jansen, R.; Bidlingmaier, S.; Houfek, T.; et al. Global analysis of protein activities using proteome chips. Science 2001, 293, 2101–2105. [Google Scholar] [CrossRef] [PubMed]
  17. Cernat, A.; Tertiş, M.; Cristea, C.; Săndulescu, R. Applications of click chemistry in the development of electrochemical sensors. Int. J. Electrochem. Sci. 2015, 10, 6324–6337. [Google Scholar]
  18. Verma, P.; Maire, P.; Novak, P. Concatenation of electrochemical grafting with chemical or electrochemical modification for preparing electrodes with specific surface functionality. Electrochim. Acta 2011, 56, 3555–3561. [Google Scholar] [CrossRef]
  19. Moreno-Guzmán, M.; Ojeda, I.; Villalonga, R.; González-Cortés, A.; Yáñez-Sedeño, P.; Pingarrón, J.M. Ultrasensitive detection of adrenocorticotropin hormone (ACTH) using disposable phenylboronic-modified electrochemical immunosensors. Biosens. Bioelectron. 2012, 35, 82–86. [Google Scholar] [CrossRef] [PubMed]
  20. Eissa, S.; Tlili, C.; L’Hocine, L.; Zourob, M. Electrochemical immunosensor for the milk allergen β-lactoglobulin based on electrografting of organic film on graphene modified screen-printed carbon electrodes. Biosens. Bioelectron. 2012, 38, 308–313. [Google Scholar] [CrossRef] [PubMed]
  21. Hayat, A.; Sassolas, A.; Marty, J.-L.; Radi, A.-E. Highly sensitive ochratoxin A impedimetric aptasensor based on the immobilization of azido-aptamer onto electrografted binary film via click chemistry. Talanta 2013, 103, 14–19. [Google Scholar] [CrossRef] [PubMed]
  22. Zhang, Y.; Geng, X.; Ai, J.; Gao, Q.; Qi, H.; Zhang, C. Signal amplification detection of DNA using a sensor fabricated by one-step covalent immobilization of amino-terminated probe DNA onto the polydopamine-modified screen-printed carbon electrode. Sens. Actuators B Chem. 2015, 221, 1535–1541. [Google Scholar] [CrossRef]
  23. Zheng, W.; van den Hurk, R.; Cao, Y.; Du, R.; Sun, X.; Wang, Y.; McDermott, M.T.; Evoy, S. Aryl diazonium chemistry for the surface functionalization of glassy Biosensors. Biosensors 2016, 6, 8. [Google Scholar] [CrossRef] [PubMed]
  24. Harper, J.C.; Polsky, R.; Wheeler, D.R.; Dirk, S.M.; Brozik, S.M. Selective immobilization of DNA and antibody probes on electrode arrays: Simultaneous electrochemical detection of DNA and protein on a single platform. Langmuir 2007, 23, 8285–8287. [Google Scholar] [CrossRef] [PubMed]
  25. Corgier, B.P.; Laurent, A.; Perriat, P.; Blum, L.J.; Marquette, C.A. A versatile method for direct and covalent immobilization of DNA and proteins on biochips. Angew. Chem. Int. Ed. 2007, 46, 4108–4110. [Google Scholar] [CrossRef] [PubMed]
  26. Polsky, R.; Harper, J.C.; Wheeler, D.R.; Dirk, S.M.; Arango, D.C.; Brozik, S.M. Electrically addressable diazonium-functionalized antibodies for multianalyte electrochemical sensor applications. Biosens. Bioelectron. 2008, 23, 757–764. [Google Scholar] [CrossRef] [PubMed]
  27. Bélanger, D.; Pinson, J. Electrografting: A powerful method for surface modification. Chem. Soc. Rev. 2011, 40, 3995–4048. [Google Scholar] [CrossRef] [PubMed]
  28. Randriamahazaka, H.; Ghilane, J. Electrografting and controlled surface functionalization of carbon based surfaces for electroanalysis. Electroanalysis 2016, 2, 13–26. [Google Scholar] [CrossRef]
  29. Corgier, B.P.; Marquette, C.A.; Blum, L.J. Diazonium-protein adducts for graphite electrode microarrays modification: Direct and addressed electrochemical immobilization. J. Am. Chem. Soc. 2005, 127, 18328–18332. [Google Scholar] [CrossRef] [PubMed]
  30. Jiang, C.; Silva, S.M.; Fan, S.; Wu, Y.; Alam, M.T.; Liu, G.; Gooding, J.J. Aryldiazonium salt derived mixed organic layers: From surface chemistry to their applications. J. Electroanal. Chem. 2017, 785, 265–278. [Google Scholar] [CrossRef]
  31. Tudorache, M.; Bala, C. Biosensors based on screen-printing technology and their applications in environmental and food analysis. Anal. Bioanal. Chem. 2007, 388, 565–578. [Google Scholar] [CrossRef] [PubMed]
  32. Taleat, Z.; Khoshroo, A.; Mazloum-Ardakani, M. Screen-printed electrodes for biosensing: A review (2008–2013). Microchim. Acta 2014, 181, 865–891. [Google Scholar] [CrossRef]
  33. Eissa, S.; Alshehri, N.; Abdel Rahman, A.M.; Dasouki, M.; Abu Salah, K.M.; Zourob, M. Electrochemical immunosensors for the detection of survival motor neuron (SMN) protein using different carbon nanomaterials-modified electrodes. Biosens. Bioelectron. 2018, 101, 282–289. [Google Scholar] [CrossRef] [PubMed]
  34. Mitchell, C.A.; Bahr, J.L.; Arepalli, S.; Tour, J.M.; Krishnamoorti, R. Dispersion of functionalized carbon nanotubes in polystyrene. Macromolecules 2002, 35, 8825–8830. [Google Scholar] [CrossRef]
  35. Dyke, C.A.; Stewart, M.P.; Maya, F.; Tour, J.M. Diazonium-based functionalization of carbon nanotubes: XPS and GC-MS analysis and mechanistic implications. Synlett 2004, 155–160. [Google Scholar] [CrossRef]
  36. Lud, S.Q.; Steenackers, M.; Jordan, R.; Bruno, P.; Gruen, D.M.; Feulner, P.; Garrido, J.A.; Stutzmann, M. Chemical grafting of biphenyl self-assembled monolayers on ultrananocrystalline diamond. J. Am. Chem. Soc. 2006, 128, 16884–16891. [Google Scholar] [CrossRef] [PubMed]
  37. Zhou, Y.L.; Zhi, J.F. Development of an amperometric biosensor based on covalent immobilization of tyrosinase on a boron-doped diamond electrode. Electrochem. Commun. 2006, 8, 1811–1816. [Google Scholar] [CrossRef]
  38. Szunerits, S.; Ghodbane, S.; Niedziolka-Jonsson, J.; Galopin, E.; Klauser, F.; Akjouj, A.; Pennec, Y.; Djafari-Rouhani, B.; Boukherroub, R.; Steinmuller-Nethl, D. Development and characterization of a diamond-based localized surface plasmon resonance interface. J. Phys. Chem. C 2010, 114, 3346–3353. [Google Scholar] [CrossRef]
  39. Baranton, S.; Belanger, D. Electrochemical derivatization of carbon surface by reduction of in situ generated diazonium cations. J. Phys. Chem. B 2005, 109, 24401–24410. [Google Scholar] [CrossRef] [PubMed]
  40. Liu, G.Z.; Chockalingham, M.; Khor, S.M.; Gui, A.L.; Gooding, J.J. A comparative study of the modification of gold and glassy carbon surfaces with mixed layers of in situ generated aryl diazonium compounds. Electroanalysis 2010, 22, 918–926. [Google Scholar] [CrossRef]
  41. Liu, G.Z.; Gooding, J.J. An interface comprising molecular wires and poly(ethylene glycol) spacer units self-assembled on carbon electrodes for studies of protein electrochemistry. Langmuir 2006, 22, 7421–7430. [Google Scholar] [CrossRef] [PubMed]
  42. Liu, G.Z.; Paddon-Row, M.N.; Gooding, J.J. A molecular wire modified glassy carbon electrode for achieving direct electron transfer to native glucose oxidase. Electrochem. Commun. 2007, 9, 2218–2223. [Google Scholar] [CrossRef]
  43. Liu, G.Z.; Paddon-Row, M.N.; Gooding, J.J. Protein modulation of electrochemical signals: Application to immunobiosensing. Chem. Commun. 2008, 3870–3872. [Google Scholar] [CrossRef] [PubMed]
  44. Khor, S.M.; Liu, G.; Fairman, C.; Iyengar, S.G.; Gooding, J.J. The importance of interfacial design for the sensitivity of a label-free electrochemical immuno-biosensor for small organic molecules. Biosens. Bioelectron. 2011, 26, 2038–2044. [Google Scholar] [CrossRef] [PubMed]
  45. Louault, C.; D’Amours, M.; Belanger, D. The electrochemical grafting of a mixture of substituted phenyl groups at a glassy carbon electrode surface. Chem. Phys. Chem. 2008, 9, 1164–1170. [Google Scholar] [CrossRef] [PubMed]
  46. Combellas, C.; Kanoufi, F.; Pinson, J.; Podvorica, F.I. Sterically hindered diazonium salts for the grafting of a monolayer on metals. J. Am. Chem. Soc. 2008, 130, 8576–8577. [Google Scholar] [CrossRef] [PubMed]
  47. Hurley, B.L.; McCreery, R.L. Covalent bonding of organic molecules to Cu and Al alloy 2024 T3 surfaces via diazonium ion reduction. J. Electrochem. Soc. 2004, 151, B252–B259. [Google Scholar] [CrossRef]
  48. Stewart, M.P.; Maya, F.; Kosynkin, D.V.; Dirk, S.M.; Stapleton, J.J.; McGuiness, C.L.; Allara, D.L.; Tour, J.M. Direct covalent grafting of conjugated molecules onto Si, Ga As, and Pd surfaces from aryldiazonium salts. J. Am. Chem. Soc. 2004, 126, 370–378. [Google Scholar] [CrossRef] [PubMed]
  49. Adenier, A.; Cabet-Deliry, E.; Chausse, A.; Griveau, S.; Mercier, F.; Pinson, J.; Vautrin-Ul, C. Grafting of nitrophenyl groups on carbon and metallic surfaces without electrochemical induction. Chem. Mater. 2005, 17, 491–501. [Google Scholar] [CrossRef]
  50. Combellas, C.; Delamar, M.; Kanoufi, F.; Pinson, J.; Podvorica, F.J. Spontaneous grafting of iron surfaces by reduction of aryldiazonium salts in acidic or neutral aqueous solution. Application to the protection of iron against corrosion. Chem. Mater. 2005, 17, 3968–3975. [Google Scholar] [CrossRef]
  51. Garrett, D.J.; Lehr, J.; Miskelly, G.M.; Downard, A.J. Microcontact printing using the spontaneous reduction of aryldiazonium salts. J. Am. Chem. Soc. 2007, 129, 15456–15457. [Google Scholar] [CrossRef] [PubMed]
  52. Lehr, J.; Williamson, B.E.; Flavel, B.S.; Downard, A.J. Reaction of gold substrates with diazonium salts in acidic solution at open-circuit potential. Langmuir 2009, 25, 13503–13509. [Google Scholar] [CrossRef] [PubMed]
  53. Delamar, M.; Hitmi, R.; Pinson, J.; Saveant, J.M. Covalent modification of carbon surfaces by grafting of functionalized aryl radicals produced from electrochemical reduction of diazonium salts. J. Am. Chem. Soc. 1992, 114, 5883–5884. [Google Scholar] [CrossRef]
  54. Bourdillon, C.; Delamar, M.; Demaille, C.; Hitmi, R.; Moiroux, J.; Pinson, J. Immobilization of glucose oxidase on a carbon surface derivatized by electrochemical reduction of diazonium salts. J. Electroanal. Chem. 1992, 336, 113–123. [Google Scholar] [CrossRef]
  55. Polsky, R.; Harper, J.C.; Dirk, S.M.; Arango, D.C.; Wheeler, D.R.; Brozik, S.M. Diazonium-functionalized horseradish peroxidase immobilized via addressable electrodeposition: Direct electron transfer and electrochemical detection. Langmuir 2007, 23, 364–366. [Google Scholar] [CrossRef] [PubMed]
  56. Alonso-Lomillo, M.A.; Yardimci, C.; Dominguez- Renedo, O.; Arcos-Martinez, M.J. CYP450 2B4 covalently attached to carbon and gold screen printed electrodes by diazonium salt and thiols monolayers. Anal. Chim. Acta 2009, 633, 51–56. [Google Scholar] [CrossRef] [PubMed]
  57. Corgier, B.P.; Marquette, C.A.; Blum, L.J. Direct electrochemical addressing of immunoglobulins: Immunochip on screen-printed microarray. Biosens. Bioelectron. 2007, 22, 1522–1526. [Google Scholar] [CrossRef] [PubMed]
  58. Hayat, A.; Barthelmebs, L.; Marty, J.-L. Electrochemical impedimetric immunosensor for the detection of okadaic acid in mussel simple. Sens. Actuators B Chem. 2012, 171–172, 810–815. [Google Scholar] [CrossRef]
  59. Eissa, S.; Zourob, M. A graphene-based electrochemical competitive immunosensor for the sensitive detection of okadaic acid in shellfish. Nanoscale 2012, 4, 7593–7599. [Google Scholar] [CrossRef] [PubMed]
  60. Eissa, S.; L’Hocineb, L.; Siaj, M.; Zourob, M. Graphene-based label-free voltammetric immunosensor for sensitive detection of the egg allergen ovalbumin. Analyst 2013, 7, 4378–4384. [Google Scholar] [CrossRef] [PubMed]
  61. Lim, S.A.; Minhaz, U.A. A label free electrochemical immunosensor for sensitive detection of porcine serum albumin as a marker for pork adulteration in raw meat. Food Chem. 2016, 206, 197–203. [Google Scholar] [CrossRef] [PubMed]
  62. Martínez-García, G.; Agüí, L.; Yáñez-Sedeño, P.; Pingarrón, J.M. Multiplexed electrochemical immunosensing of obesity-related hormones at grafted graphene-modified electrodes. Electrochim. Acta 2016, 202, 209–215. [Google Scholar] [CrossRef]
  63. Martínez-García, G.; Agüí, L.; Yáñez-Sedeño, P.; Pingarrón, J.M. Multiplexed electrochemical immunosensor for obesity-related hormones using grafted graphene-modified electrodes as platforms for antibodies immobilization. Proc. Technol. 2017, 27, 187–189. [Google Scholar] [CrossRef]
  64. Hosu, O.; Tertiş, M.; Melinte, G.; Săndulescu, R.; Cristea, C. Mucin 4 immunosensor based on p-aminophenylacetic acid grafting on carbon electrodes as immobilization platform. Proc. Technol. 2017, 27, 110–111. [Google Scholar] [CrossRef]
  65. Ruffien, A.; Dequaire, M.; Brossier, P. Covalent immobilization of oligonucleotides on p-aminophenyl-modified carbon screen-printed electrodes for viral DNA sensing. Chem. Commun. 2003, 7, 912–913. [Google Scholar] [CrossRef]
  66. Shabani, A.; Mak, A.W.H.; Gerges, I.; Cuccia, L.A.; Lawrence, M.F. DNA immobilization onto electrochemically functionalized Si(100) surfaces. Talanta 2006, 70, 615–623. [Google Scholar] [CrossRef] [PubMed]
  67. Hajdukiewicz, J.; Boland, S.; Kavanagh, P.; Leech, D. An enzyme-amplified amperometric DNA hybridisation assay using DNA immobilised in a carboxymethylated dextran film anchored to a graphite surface. Biosens. Bioelectron. 2010, 25, 1037–1042. [Google Scholar] [CrossRef] [PubMed]
  68. Obaje, E.A.; Cummins, G.; Schulze, H.; Mahmood, S.; Desmulliez, M.P.Y.; Bachmann, T.T. Carbon screen-printed electrodes on ceramic substrates for label-free molecular detection of antibiotic resistance. J. Interdiscip. Nanomed. 2016, 1. [Google Scholar] [CrossRef]
  69. Ocaña, C.; Hayat, A.; Mishra, R.K.; Vasilescu, A.; del Valle, M.; Marty, J.-L. Label free aptasensor for Lysozyme detection: A comparison of the analytical performance of two aptamers. Bioelectrochemistry 2015, 105, 72–77. [Google Scholar] [CrossRef] [PubMed]
  70. Ocaña, C.; Hayat, A.; Mishra, R.; Vasilescu, A.; del Valle, M.; Marty, J.-L. A novel electrochemical aptamer–antibody sandwich assay for lysozyme detection. Analyst 2015, 140, 4148–4153. [Google Scholar] [CrossRef] [PubMed]
  71. Liu, G.Z.; Nguyen, Q.T.; Chow, E.; Bocking, T.; Hibbert, D.B.; Gooding, J.J. Study of factors affecting the performance of voltammetric copper sensors based on Gly-Gly-His modified glassy carbon and gold electrodes. Electroanalysis 2006, 18, 1141–1151. [Google Scholar] [CrossRef]
  72. Liu, G.Z.; Bocking, T.; Gooding, J.J. Diazonium salts: Stable monolayers on gold electrodes for sensing applications. J. Electroanal. Chem. 2007, 600, 335–344. [Google Scholar] [CrossRef]
  73. Zhang, X.; Sun, G.G.; Hovestadt, M.; Syritski, V.; Esser, N.; Volkmer, R.; Janietz, S.; Rappich, J.; Hinrichs, K. A new strategy for the preparation of maleimide-functionalised gold surfaces. Electrochem. Commun. 2010, 12, 1403–1406. [Google Scholar] [CrossRef]
  74. Harper, J.C.; Polsky, R.; Wheeler, D.R.; Lopez, D.M.; Arango, D.C.; Brozik, S.M. A multifunctional thin film Au electrode surface formed by consecutive electrochemical reduction of aryl diazonium salts. Langmuir 2009, 25, 3282–3288. [Google Scholar] [CrossRef] [PubMed]
  75. Harnisch, J.A.; Pris, A.D.; Porter, M.D. Attachment of gold nanoparticles to glassy carbon electrodes via a mercaptobenzene film. J. Am. Chem. Soc. 2001, 123, 5829–5830. [Google Scholar] [CrossRef] [PubMed]
  76. Chehimi, M.M. Aryl Diazonium Salts: New Coupling Agents and Surface Science; Editorial Wiley-VCH: New York, NY, USA, 2012; ISBN 978-3-527-32998-4. [Google Scholar]
  77. Chen, H.J.; Wang, Y.L.; Qu, J.Y.; Dong, S.J. Self-assembled silver nanoparticle monolayer on glassy carbon: An approach to SERS substrate. J. Raman Spectrosc. 2007, 38, 1444–1448. [Google Scholar] [CrossRef]
  78. Baron, R.; Wildgoose, G.G.; Compton, V. Metallic nanoparticles deposited on carbon microspheres: Novel materials for combinatorial electrochemistry and electroanalysis. J. Nanosci. Nanotechnol. 2009, 9, 2274–2282. [Google Scholar] [CrossRef] [PubMed]
  79. Gehan, H.; Fillaud, L.; Felidj, N.; Aubard, J.; Lang, P.; Chehimi, M.M.; Mangeney, C. A general approach combining diazonium salts and click chemistries for gold surface functionalization by nanoparticle assemblies. Langmuir 2010, 26, 3975–3980. [Google Scholar] [CrossRef] [PubMed]
  80. Noel, J.M.; Zigah, D.; Simonet, J.; Hapiot, P. Synthesis and immobilization of Ag0 nanoparticles on diazonium modified electrodes: SECM and cyclic voltammetry studies of the modified interfaces. Langmuir 2010, 26, 7638–7643. [Google Scholar] [CrossRef] [PubMed]
  81. Garrett, D.J.; Flavel, B.S.; Shapter, J.G.; Baronian, K.H.R.; Downard, A.J. Robust forests of vertically aligned carbon nanotubes chemically assembled on carbon substrates. Langmuir 2010, 26, 1848–1854. [Google Scholar] [CrossRef] [PubMed]
  82. Liu, G.Z.; Luais, E.; Gooding, J.J. The fabrication of stable gold nanoparticle-modified interfaces for electrochemistry. Langmuir 2011, 27, 4176–4183. [Google Scholar] [CrossRef] [PubMed]
  83. Liu, G.Z.; Liu, J.Q.; Davis, T.P.; Gooding, J.J. Electrochemical impedance immunosensor based on gold nanoparticles and aryldiazonium salt functionalized gold electrodes for the detection of antibody. Biosens. Bioelectron. 2011, 26, 3660–3665. [Google Scholar] [CrossRef] [PubMed]
  84. Serafín, V.; Torrente-Rodríguez, R.M.; González-Cortés, A.; García de Frutos, P.; Sabaté, M.; Campuzano, S.; Yáñez-Sedeño, P.; Pingarrón, J.M. An electrochemical immunosensor for brain natriuretic peptide prepared with screen-printed carbon electrodes nanostructured with gold nanoparticles grafted through aryl diazonium salt chemistry. Talanta 2018, 179, 131–138. [Google Scholar] [CrossRef] [PubMed]
  85. Moreno-Guzmán, M.; González-Cortés, A.; Yáñez-Sedeño, P.; Pingarrón, J.M. Multiplexed ultrasensitive determination of adrenocorticotropin and cortisol hormones at a dual electrochemical immunosensor. Electroanalysis 2012, 24, 1100–1108. [Google Scholar] [CrossRef]
  86. Ojeda, I.; Barrejón, M.; Arellano, L.M.; González-Cortés, A.; Yáñez-Sedeño, P.; Langa, F.; Pingarrón, J.M. Grafted-double walled carbon nanotubes as electrochemical platforms for immobilization of antibodies using a metallic-complex chelating polymer: Application to the determination of adiponectin cytokine in serum. Biosens. Bioelectron. 2015, 74, 24–29. [Google Scholar] [PubMed]
  87. Sánchez-Tirado, E.; Salvo, C.; González-Cortés, A.; Yáñez-Sedeño, P.; Langa, F.; Pingarrón, J.M. Electrochemical immunosensor for simultaneous determination of interleukin-1 beta and tumor necrosis factor alpha in serum and saliva using dual screen printed electrodes modified with functionalized double–walled carbon nanotubes. Anal. Chim. Acta 2017, 959, 66–73. [Google Scholar] [CrossRef] [PubMed]
  88. Sánchez-Tirado, E.; Martínez-García, G.; González-Cortés, A.; Yáñez-Sedeño, P.; Pingarrón, J.M. Electrochemical immunosensor for sensitive determination of transforming growth factor (TGF)-β1 in urine. Biosens. Bioelectron. 2017, 88, 9–14. [Google Scholar] [CrossRef] [PubMed]
  89. Radi, A.-E.; Muñoz-Berbel, X.; Cortina-Puig, M.; Marty, J.-L. An electrochemical immunosensor for ochratoxin A based on immobilization of antibodies on diazonium-functionalized gold electrode. Electrochim. Acta 2009, 54, 2180–2184. [Google Scholar] [CrossRef]
  90. Qi, H.; Ling, C.; Ma, Q.; Gao, Q.; Zhang, C. Sensitive electrochemical immunosensor array for the simultaneous detection of multiple tumor markers. Analyst 2012, 137, 393–399. [Google Scholar] [CrossRef] [PubMed]
  91. Goud, K.Y.; Catanantea, G.; Hayat, A.; Satyanarayana, M.; Gobi, K.V.; Marty, J.-L. Disposable and portable electrochemical aptasensor for label free detection of aflatoxin B1 in alcoholic beverages. Sens. Actuators B Chem. 2016, 235, 466–473. [Google Scholar] [CrossRef]
  92. Mishra, R.K.; Hayat, A.; Catanante, G.; Ocaña, C.; Marty, J.-L. A label free aptasensor for Ochratoxin A detection in cocoa beans: An application to chocolate industries. Anal. Chim. Acta 2015, 889, 106–112. [Google Scholar] [CrossRef] [PubMed]
  93. Istamboulié, G.; Paniel, N.; Zara, L.; Reguillo Granados, L.; Barthelmebs, L.; Noguer, T. Development of an impedimetric aptasensor for the determination of aflatoxin M1 in milk. Talanta 2016, 146, 464–469. [Google Scholar] [CrossRef] [PubMed]
  94. Revenga-Parra, M.; Garcia-Mendiola, T.; Gonzalez-Costas, J.; Gonzalez-Romero, E.; Garcia Marin, A.; Pau, J.L.; Pariente, F.; Lorenzo, E. Simple diazonium chemistry to develop specific gene sensing platforms. Anal. Chim. Acta 2014, 813, 41–47. [Google Scholar] [CrossRef] [PubMed]
  95. Bagheryan, Z.; Raoof, J.-B.; Golabi, M.; Turner, A.P.F.; Beni, V. Diazonium-based impedimetric aptasensor for the rapid label-free detection of Salmonella typhimurium in food sample. Biosens. Bioelectron. 2016, 80, 566–573. [Google Scholar] [CrossRef] [PubMed]
  96. Bahr, J.L.; Tour, J.M. Highly functionalized carbon nanotubes using in situ generated diazonium compounds. Chem. Mater. 2001, 13, 3823–3824. [Google Scholar] [CrossRef]
  97. Gooding, J.J.; Mearns, F.; Yang, W.; Liu, J. Self-assembled monolayers into the 21st century: Recent advances and applications. Electroanalysis 2003, 15, 81–96. [Google Scholar] [CrossRef]
  98. Gooding, J.J. Advances in interfacial design for electrochemical biosensors and sensors: Aryl diazonium salts for modifying carbon and metal electrodes. Electroanalysis 2008, 20, 573–582. [Google Scholar] [CrossRef]
  99. Lyskawa, J.; Belanger, D. Direct modification of a gold electrode with aminophenyl groups by electrochemical reduction of in situ generated aminophenyl monodiazonium cations. Chem. Mater. 2006, 18, 4755–4763. [Google Scholar] [CrossRef]
  100. Phal, S.; Shatri, B.; Berisha, A.; Geladi, P.; Lindholm-Sethson, B.; Tesfalidet, S. Covalently electrografted carboxyphenyl layers onto gold surface serving as a platform for the construction of an immunosensor for detection of methotrexate. J. Electroanal. Chem. 2018, 812, 235–243. [Google Scholar] [CrossRef]
  101. Gooding, J.J. Electrochemical DNA hybridization biosensors. Electroanalysis 2002, 14, 1149–1156. [Google Scholar] [CrossRef]
  102. Rauf, S.; Gooding, J.J.; Akhtar, K.; Ghauri, M.-A.; Rahman, M.; Anwar, M.A.; Khalid, A.M. Electrochemical approach of anticancer drugs–DNA interaction. J. Pharm. Biomed. Anal. 2005, 37, 205–217. [Google Scholar] [CrossRef] [PubMed]
  103. Pividori, M.I.; Merckoci, A.; Alegret, S. Electrochemical genosensor design: Immobilisation of oligonucleotides onto transducer surfaces and detection methods. Biosen. Bioelectron. 2000, 15, 291–303. [Google Scholar] [CrossRef]
  104. Umek, R.M.; Lin, S.W.; Vielmetter, J.; Terbrueggen, R.H.; Irvine, B.; Yu, C.J.; Kayyem, J.F.; Yowanto, H.; Blackburn, G.F.; Farkas, D.H.; et al. Electronic detection of nucleic acids: A versatile platform for molecular diagnostics. J. Mol. Diagn. 2001, 3, 74–84. [Google Scholar] [CrossRef]
  105. Wang, J.; Rivas, G.; Cai, X. Screen-printed electrochemical hybridization biosensor for the detection of DNA sequences from the Escherichia coli pathogen. Electroanalysis 1997, 9, 395–398. [Google Scholar] [CrossRef]
  106. Marrazza, G.; Chianella, I.; Mascini, M. Disposable DNA electrochemical sensor for hybridization detection. Biosens. Bioelectron. 1999, 14, 43–51. [Google Scholar] [CrossRef]
  107. Azek, F.; Grossiord, C.; Joannes, M.; Limoges, B.; Brossier, P. Hybridization assay at a disposable electrochemical biosensor for the attomole detection of amplified human cytomegalovirus DNA. Anal. Biochem. 2000, 284, 107–113. [Google Scholar] [CrossRef] [PubMed]
  108. Authier, L.; Grossiord, C.; Limoges, B.; Brossier, P. Gold nanoparticle-based quantitative electrochemical detection of amplified human cytomegalo-virus DNA using disposable microband electrodes. Anal. Chem. 2001, 73, 4450–4456. [Google Scholar] [CrossRef] [PubMed]
  109. Alwine, J.C.; Kemp, D.J.; Stark, G.R. Method for detection of specific RNAs in agarose gels by transfer to diazobenzyloxymethyl-paper and hybridization with DNA probes. Proc. Natl. Acad. Sci. USA 1977, 74, 5350–5353. [Google Scholar] [CrossRef]
  110. Dolan, P.L.; Wu, Y.; Ista, L.K.; Metzenberg, R.L.; Nelson, M.A.; Lopez, G.P. Robust and efficient synthetic method for forming DNA microarrays. Nucleic Acids Res. 2001, 29, e107. [Google Scholar] [CrossRef] [PubMed]
  111. Roth, K.M.; Peyvan, K.; Schwarzkopf, K.R.; Ghindilis, A. Electrochemical detection of short DNA oligomer hybridization using the combi matrix electra sense microarray reader. Electroanalysis 2006, 18, 1982–1988. [Google Scholar] [CrossRef]
  112. Harper, J.C.; Polsky, R.; Wheeler, D.R.; Brozik, S.M. Maleimide-Activated Aryl Diazonium Salts for Electrode Surface Functionalization with Biological and Redox-Active Molecules. Langmuir 2008, 24, 2206–2211. [Google Scholar] [CrossRef] [PubMed]
  113. Hansen, M.N.; Farjami, E.; Kristiansen, M.; Clima, L.; Pedersen, S.U.; Daasbjerg, K.; Ferapontova, E.E.; Gothelf, K.V. Synthesis and application of a triazene−ferrocene modifier for immobilization and characterization of oligonucleotides at electrodes. J. Org. Chem. 2010, 75, 2474–2481. [Google Scholar] [CrossRef] [PubMed]
  114. Polsky, R.; Harper, J.C.; Wheeler, D.R.; Arango, D.C.; Brozik, S.M. Electrically addressable cell immobilization using phenylboronic acid diazonium salts. Angew. Chem. Int. Ed. 2008, 47, 2631–2634. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Functionalization of different carbon nanomaterial-modified SPEs (a) via diazonium salt reduction and affinity biosensor fabrication (b). Reprinted and adapted from [33] with permission.
Figure 1. Functionalization of different carbon nanomaterial-modified SPEs (a) via diazonium salt reduction and affinity biosensor fabrication (b). Reprinted and adapted from [33] with permission.
Sensors 18 00675 g001
Scheme 1. Reaction mechanisms proposed for the preparation of single (a) or multiple (b) layers onto electrode surfaces through electroreductive electrografting of aryldiazonium salts. In (b) R1 and R2 represent two different substituents. Reprinted from [30] with permission.
Scheme 1. Reaction mechanisms proposed for the preparation of single (a) or multiple (b) layers onto electrode surfaces through electroreductive electrografting of aryldiazonium salts. In (b) R1 and R2 represent two different substituents. Reprinted from [30] with permission.
Sensors 18 00675 sch001
Figure 2. Preparation of a diazonium-modified antibody electrode: carboxyl diazonium is covalently attached to antibody by EDC/NHS (1) and diazonium–antibody is deposited onto an electrode by cyclic voltammetry (2). Reprinted and adapted from [26] with permission.
Figure 2. Preparation of a diazonium-modified antibody electrode: carboxyl diazonium is covalently attached to antibody by EDC/NHS (1) and diazonium–antibody is deposited onto an electrode by cyclic voltammetry (2). Reprinted and adapted from [26] with permission.
Sensors 18 00675 g002
Figure 3. Schematic display of the working principle of the immunosensor based on a direct competitive format and SWVs of the immunosensor before the competition step (1) and after incubation with different concentrations of OA: 0.00 (2), 1.00 (3), 10.0 (4), 100 (5) and 1000 (6) ng L−1. Reprinted and adapted from [59] with permission.
Figure 3. Schematic display of the working principle of the immunosensor based on a direct competitive format and SWVs of the immunosensor before the competition step (1) and after incubation with different concentrations of OA: 0.00 (2), 1.00 (3), 10.0 (4), 100 (5) and 1000 (6) ng L−1. Reprinted and adapted from [59] with permission.
Sensors 18 00675 g003
Figure 4. Electrochemical immunosensor developed for pSA determination onto a CNF SPE modified with a 4-carboxyphenyl layer (a) and comparison of the DPV responses provided by the immunosensor to 100 pg mL−1 of serum albumin from porcine, bovine, rabbit and albumin in chicken egg (b). Reprinted and adapted from [61] with permission.
Figure 4. Electrochemical immunosensor developed for pSA determination onto a CNF SPE modified with a 4-carboxyphenyl layer (a) and comparison of the DPV responses provided by the immunosensor to 100 pg mL−1 of serum albumin from porcine, bovine, rabbit and albumin in chicken egg (b). Reprinted and adapted from [61] with permission.
Sensors 18 00675 g004
Figure 5. Schematic illustration of the different steps and protocols involved in the preparation and functioning of the dual GHRL and PYY immunosensor. Reprinted from [63] with permission.
Figure 5. Schematic illustration of the different steps and protocols involved in the preparation and functioning of the dual GHRL and PYY immunosensor. Reprinted from [63] with permission.
Sensors 18 00675 g005
Figure 6. Schematic display of the different steps involved in the construction of an amperometric immunosensor for APN involving grafted DWCNTs and oriented immobilization of anti-APN by using the metallic-complex chelating polymer Mix & Go. Reprinted from [86] with permission.
Figure 6. Schematic display of the different steps involved in the construction of an amperometric immunosensor for APN involving grafted DWCNTs and oriented immobilization of anti-APN by using the metallic-complex chelating polymer Mix & Go. Reprinted from [86] with permission.
Sensors 18 00675 g006
Figure 7. Schematic display of the working principle of the impedimetric aptasensor for the determination of OTA. Reprinted from [21] with permission.
Figure 7. Schematic display of the working principle of the impedimetric aptasensor for the determination of OTA. Reprinted from [21] with permission.
Sensors 18 00675 g007
Figure 8. Impedimetric aptasensor developed for Lys determination on a 4-ABA-modified SPCE. Reprinted from [69] with permission.
Figure 8. Impedimetric aptasensor developed for Lys determination on a 4-ABA-modified SPCE. Reprinted from [69] with permission.
Sensors 18 00675 g008
Figure 9. Overview of the preparation of the S. typhimurium aptasensor. Reprinted from [95] with permission.
Figure 9. Overview of the preparation of the S. typhimurium aptasensor. Reprinted from [95] with permission.
Sensors 18 00675 g009
Table 1. Electrochemical affinity biosensors involving aryldiazonium salt chemistry onto screen printed electrodes.
Table 1. Electrochemical affinity biosensors involving aryldiazonium salt chemistry onto screen printed electrodes.
ElectrodeConfigurationAnalyteDetection SchemeDetection TechniqueLinear RangeLODSampleRef.
SPCESandwich immunosensor, grafted p-ABA-4-APBA-anti-ACTH-ACTH-Biotin-Strept-APACTH1-NPP + AP → 1-NP
1-NP → quinone
DPV0.025–1.0 pg mL−118 pg L−1serum[19]
SPdCESandwich immunosensor, grafted p-ABA-4-APBA-anti-ACTH-ACTH-Biotin-Strept-AP
Direct competitive, grafted p-ABA-4-APBA-anti-cortisol-cortisol-AP
ACTH
cortisol
1-NPP + AP → 1-NP
1-NP → quinone
DPV5.0 × 10−5–0.1ng mL−1
0.1–500 ng mL−1
40 pg L−1 37 pg mL−1serum[85]
SPCESandwich immunosensor, grafted p-aminothio-phenol-AuNPs-anti-BNP-BNP-anti-BNP-HRPBNPH2O2 + HRP + HQ
→ quinone
amperometry0.014–15 ng mL−14 pg L−1serum[84]
SPCESandwich immunosensor, grafted p-ABA-DWCNTs-M&G-anti-APN-APN-Biotin-Strept-HRPAPNH2O2 + HRP + HQ
→ quinone
amperometry0.05–10.0 μg mL−114.5 ng mL−1serum[86]
SPdCEDirect competitive immunosensor, grafted p-ABA-anti-GHRL GHRL-Biotin-Strept-HRPDirect competitive, grafted p-ABA-anti-PYY- PYY-Biotin-Strept-HRPGHRLPYYH2O2 + HRP + HQ
→ quinone
DPV10−3–100 ng mL−1
10−4–10 ng mL−1
1.0 pg mL−1 0.02 pg mL−1serum,
saliva
[62]
SPdCESandwich immunosensor, grafted p-ABA-DWCNTs -M&G-anti-TNF-TNF-anti-TNF-Biotin-poly-Strept-HRP. Sandwich, grafted p-ABA-DWCNTs-M&G-anti-IL1β-IL1β-anti-IL1β-Biotin-poly-Strept-HRPTNF-α
IL-1β
H2O2 + HRP + HQ
→ quinone
amperometry1–200 pg mL−1
0.5–100 pg mL−1
0.85 pg mL−1
0.38 pg mL−1
serum
saliva
[87]
SPCESandwich immunosensor, grafted p-ABA-Strept-Biotin-anti-TGF-TGF-anti-TGF-SWCNT(-HRP)-Phe-VTGF-β1H2O2 + HRP + HQ
→ quinone
amperometry2.5–1000 pg mL−10.95 pg mL−1saliva[88]
SPAuEDirect competitive immunosensor, grafted 4-nitrophenyl red.-GA-anti-OTA-HRPOTATMB reductionamperometryup to 60 ng mL−112 ng mL−1-[89]
SPCEIndirect competitive immunosensor, grafted 4-carboxy-phenyl-HMDA-OA-anti-OA-IgG-APOA1-NPP + AP → 1-NP 1-NP → quinoneDPV-1.44 ng L−1mussel[5]
SPCE arraySandwich immunosensor, grafted 4-aminophenyl-GA-anti-CEA(or anti-AFP)-anti-CEA(or anti-AFP)-HRPCEAAFPH2O2 + HRP + HQ
→ quinone
amperometry0.10–50 ng mL−10.03 ng mL−1
0.05 ng mL−1
serum[90]
SPCELabel free immunosensor, grafted 4-carboxyphenyl-anti-OA-OAOARCT [Fe(CN)6]4−/3−EIS0.195–12.5 μg mL−10.3 μg L−1mussel[58]
GrSPELabel free direct competitive immunosensor, grafted 4-carboxyphenyl-anti-OA-OA-OVAOAreduction current [Fe(CN)6]4−/3−SWVup to 5000 ng mL−119 ng L−1shellfish tissue[59]
GrSPELabel free immunosensor, grafted 4-nitrophenyl red.-GA-anti-β-LGB-β-LGBβ-LGBreduction current [Fe(CN)6]4−/3−DPV1 pg mL−1–100 ng mL−10.85 pg mL−1cake, cheese, snacks[20]
CnFSPELabel free immunosensor, grafted 4-carboxyphenyl-anti-pSA-pSApSAreduction current [Fe(CN)6]4−/3−DPV0.5–500 pg mL−10.5 pg mL−1fresh meat[61]
CnFSPELabel free immunosensor, grafted 4-carboxyphenyl-anti-pSA-pSASMNreduction current [Fe(CN)6]4−/3−SWV1.0 pg mL−1–100 ng mL−10.75 pg mL−1whole blood[33]
GrSPELabel-free immunosensor, grafted 4-carboxyphenyl-anti-OVA-OVAOVAreduction current [Fe(CN)6]4−/3−DPV1.0 pg mL−1–0.5 μg mL−10.83 pg mL−1egg-free cake[60]
SPCELabel free aptasensor, grafted 4-carboxyphenyl-aptamer AFB1AFB1RCT [Fe(CN)6]4−/3−EIS0.125 ng mL−1–16 ng mL−10.12 ng mL−1beer
wine
[91]
SPCELabel free immunosensor, grafted 4-carboxyphenyl-acetic-anti-MUC4MUC4RCT [Fe(CN)6]4−/3−EIS1–15 μg mL−10.33 μg mL−1serum[64]
SPCELabel free aptasensor, grafted p-ABA -aptamer-OTAOTARCT [Fe(CN)6]4−/3−EIS0.15–2.5 ng mL−10.15 ng mL−1cocoa beans[92]
SPCELabel free aptasensor, grafted 4-carboxyphenyl-aptamer-AFM1AFM1RCT [Fe(CN)6]4−/3−EIS2–150 ng L−11.15 ng L−1milk[93]
SPCELabel free DNA sensor, grafted 4-nitrophenyl red.-DNAMRP3 genereduction current Ru(NH3)63+DPV1.0–7.3 ng μL−1210 pg μL−1-[94]
SPCELabel free aptasensor, grafted 4-carboxyphenyl-NH2-aptamer-salmonellaS. typhi.RCT [Fe(CN)6]4−/3−EIS10–108 CFU mL−16 CFU mL−1apple juice[95]
SPCESandwich aptasensor, grafted 4-carboxyphenyl-aptamer-anti-Lys-Biotin-avidin-APLys1-NPP + AP → 1-NP 1-NP → quinoneDPV5 fM–5 nM4.3 fMwine[70]
SPCELabel free aptasensor, grafted 4-((trimethylsilyl)ethynyl) benzene- p-nitrobenzene + click azide aptamer-OTAOTARCT [Fe(CN)6]4−/3−EIS1.25–500 ng L−10.25 ng L−1beer[21]
ACTH, adrenocorticotropin; 2-ABA, 2-aminobenzyl amine; p-ABA, p-aminobenzoic acid; AFB1, aflatoxin B1; AFM1, aflatoxin M1; AP, alkaline phosphatase; 4-APBA, 4-aminophenylboronic acid; APN, adiponectin; AuNP, gold nanoparticles; BNP, brain natriuretic peptide; CEA, carcinoembrionic antigen; CnFSPE, carbon nanofibers screen printed electrode; DPV, differential pulse voltammetry; DWCNTs, double walled carbon nanotubes; EIS, electrochemical impedance spectroscopy; GQDs, graphene quantum dots; GHRL, ghrelin; GrSPE, graphene-modified screen-printed carbon electrode; HQ, hydroquinone; HRP, horseradish peroxidase; IL-1β; interleukin-1β; β-LGB, β-lactoglobulina; Lys, lysozyme; M&G, Mix&Go;MRP3 gene; MUC4, mucin 4; 1-NP, 1-naphthylphenol; 1-NPP, 1-naphthylphosphate; OA, okadaic acid; OTA, ochratoxin A; OVA, ovoalbumin; PYY, peptide YY; pSA, porcine serum albumin; RCT, charge transfer resistance; S. typhi., Salmonella typhimurium; SMN, survival motor neuron: SPAuE, screen printed gold electrode; SPCE, screen printed carbon electrode; SPdCE, screen printed dual carbon electrode; SWCNT, single walled carbon nanotubes; SWV, square wave voltammetry; TGF-β1, transforming growth factor β1; TNF-α, tumor necrosis factor alpha; V, viologen.

Share and Cite

MDPI and ACS Style

Yáñez-Sedeño, P.; Campuzano, S.; Pingarrón, J.M. Integrated Affinity Biosensing Platforms on Screen-Printed Electrodes Electrografted with Diazonium Salts. Sensors 2018, 18, 675. https://doi.org/10.3390/s18020675

AMA Style

Yáñez-Sedeño P, Campuzano S, Pingarrón JM. Integrated Affinity Biosensing Platforms on Screen-Printed Electrodes Electrografted with Diazonium Salts. Sensors. 2018; 18(2):675. https://doi.org/10.3390/s18020675

Chicago/Turabian Style

Yáñez-Sedeño, Paloma, Susana Campuzano, and José M. Pingarrón. 2018. "Integrated Affinity Biosensing Platforms on Screen-Printed Electrodes Electrografted with Diazonium Salts" Sensors 18, no. 2: 675. https://doi.org/10.3390/s18020675

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop