Skip to main content

MINI REVIEW article

Front. Public Health, 05 May 2017
Sec. Environmental health and Exposome
This article is part of the Research Topic The environment-animal-human web: a “One Health” view of toxicological risk analysis View all 14 articles

Portable Bio/Chemosensoristic Devices: Innovative Systems for Environmental Health and Food Safety Diagnostics

\r\nRoberto DragoneRoberto Dragone1Gerardo Grasso*Gerardo Grasso1*Michele MucciniMichele Muccini2Stefano Toffanin\r\nStefano Toffanin2
  • 1Institute of Nanostructured Materials (ISMN), Consiglio Nazionale delle Ricerche (CNR), Rome, Italy
  • 2Institute of Nanostructured Materials (ISMN), Consiglio Nazionale delle Ricerche (CNR), Bologna, Italy

This mini-review covers the newly developed biosensoristic and chemosensoristic devices described in recent literature for detection of contaminants in both environmental and food real matrices. Current needs in environmental and food surveillance of contaminants require new simplified, sensitive systems, which are portable and allow for rapid and on-site monitoring and diagnostics. Here, we focus on optical and electrochemical bio/chemosensoristic devices as promising tools with interesting analytical features that can be potentially exploited for innovative on-site and real-time applications for diagnostics and monitoring of environmental and food matrices (e.g., agricultural waters and milk). In near future, suitably developed and implemented bio/chemosensoristic devices will be a new and modern technological solution for the identification of new quality and safety marker indexes as well as for a more proper and complete characterization of abovementioned environmental and food matrices. Integrated bio/chemosensoristic devices can also allow an “holistic approach” that may prove to be more suitable for diagnostics of environmental and food real matrices, where the copresence of more bioactive substances is frequent. Therefore, this approach can be focused on the determination of net effect (mixture effect) of bioactive substances present in real matrices.

Environmental Health and Food Safety: Scenario and Needs

Over the last few years, the abiotic contaminants levels in the environmental compartments and food increased to the point where they can cause potential human health effects due to exposure to chemical toxic substances. In particular, the interactions between environment and food supply chain that mainly occur at primary production level (including harvesting, milking and farmed animal production prior to slaughter, hunting and fishing, and harvesting of wild products) can cause serious both short- and long-term detrimental effects on human health.

Environmental and food safety remains a major global challenge, in particular in developing countries, where socioeconomic status predisposes a large share of the population to a direct environmental-origin contamination and/or consumption of contaminated food products.

To minimize the negative and dramatic impacts (especially in developing countries) of chemical toxic substances of anthropic origin on environmental and human health, several focused actions are needed. For instance, promoting a sustainable use of chemicals and agrochemicals (e.g., pesticides, veterinary antibiotics, and food additives); the development of toxicovigilance practices and systems (1) and more effective primary prevention strategies (raising users’ awareness and promoting the use of good practice codes); moreover, implementation of legislative regulations; and the development of proper infrastructures and effective protocols for safely recycle and dispose of hazardous wastes are necessary.

Against this background, such critical issues have produced a great demand for simplified, sensitive, and rapid screening methods (2, 3), without (or with reduced) sample pretreatments, suitable for environmental monitoring and surveillance at critical control points throughout the entire agro-food supply chain. In fact, recent progress and challenges in the field of analytical chemistry are focused on improving analytical methods with reduced environmental impact and developing new analytical sensoristic devices for continuous monitoring and diagnostics oriented toward environmental health and food safety.

On-site, cost-effective sensoristic devices capable of routine, sensitive, and selective detection of a range of targeted contaminants present in the environment and foods can be employed, for instance, to overcome time limitations and to reduce costs of sample collection and transport to laboratories, thus providing benefits for a rapid diagnostics and early corrective actions.

Emerging Role of Portable Bio/Chemosensoristic Devices in Environment and Agro-Food Supply Chain Monitoring

For on-site diagnostics and environmental/food monitoring purposes, the application of standard and traditional analytical techniques (very sensitive and selective techniques, but costly, time consuming and requiring trained personnel with technical skills to perform the analysis) is in contrast with the current need of rapid, cheap, easy-to-use, and portable devices (4).

For these purposes, chemosensoristic and biosensoristic devices (herein collectively referred to as bio/chemosensoristic devices) are promising tools with interesting analytical features, which can be potentially exploited for on-site real-time applications, diagnostics, and screening for both environmental and food matrices. Such devices could be employed, e.g., to overcome existing limitations in measurements currently used in environmental and agro-food fields. While those measurements are mainly focused on the independent analyses of various parameters and analytes, complexity of environmental and food matrices requires a new holistic-like approach (4).

For instance, regarding nutritional and toxicology characterization of foods, a broader modern vision based on the concept of “whole food” is taking off (4, 5). Environmental and food matrices are complex mixtures of bioactive molecules, whose complex interactions between individual components could eventually produce different and hardly theoretically predictable “net effects.” “Net effect” is necessarily different from single effects of each individual substance, and it could be additive (when it is equal to the sum of contributions of individual substances), synergistic (when it is greater than the sum of contributions of the individual substances), and antagonistic (when it is less than the sum of contributions of individual substances).

For a more proper and complete characterization of food matrices (and environmental ones) “as a whole,” integrated analyses of physical, chemical, and biological parameters through sensoristic devices could be more suitable. The multichannel platform BEST (6) is a HACCP-like monitoring system that follows this approach for the generation of integrated analytical information. More specifically, BEST focuses on identification, control, simultaneous, and non-stop monitoring of anomalous variations throughout agro-zootechnical productions, developed to allow simultaneous collection and analysis of multiple signals. Such signals are produced from a battery of selected analogical and/or digital bio/chemosensoristic devices (or probes), integrated with each other and functioning simultaneously. The simultaneous acquisition of multiparameters and integrated information can be useful in determining correlations and relationships among different data (through multivariate data analysis), and it can constitute a flexible grid of indexes and multiple markers in series. Such integrated analytical approach helps to define a “fingerprint” and to identify new marker indexes of food matrices. A field validation of BEST prototype is taking place in a farm in the Lazio region (Italy) within project ALERT (7). This project, funded by the Italian Ministry of Economic Development under the Call Industria 2015 New technologies for Made in Italy (www.alert2015.it), aims at developing the BEST prototype for industrial-scale production. Another new interesting approach for innovative monitoring and diagnostics of the environment and the agro-food supply chain is provided by a recent patented physicochemical sensing device called SNOOP (8). SNOOP is a multiparameter and multisignal sensoristic device that uses advanced and appropriately designed sensitive materials. Such sensitive materials can be both biological materials (e.g., whole cells, enzymes, and aptamers) and chemical materials (newly synthesized and/or functionalized inorganic and organic materials), whose one of the main features is the specific interaction with the target analyte(s) present in real and complex matrices. Such interactions can produce specific or aspecific physicochemical (electric or optic) responses, and the simultaneous use of different sensitive materials and the combination/integration of outgoing signals can significantly increase the screening ability of SNOOP.

Electrochemical Bio/Chemosensoristic Devices

The field of electrochemical and optical bio/chemosensoristic devices has grown rapidly in the past few years. Thanks to advantages provided by intrinsic analytical features and the development of new advanced sensitive materials, the employment of these devices has proved to be very useful for chemical contaminants detection in environmental and food matrices (Tables S1 and S2 in Supplementary Material). In particular, biosensors or biosensoristic devices (an integrated receptor–transducer device, which is capable of providing selective quantitative or semiquantitative analytical information using a biological recognition element) (9) hold promise to be relatively cheap and portable devices for in situ detection of environmental and food contaminants (10, 11).

A recent review on key research interests in the development of biosensors in South Africa has highlight a particular interest on the development of electrochemical (amperometric, impedimetric, potentiometric, and voltammetric) biosensor due to low fabrication and analytical equipment costs, in particular, for pesticides and heavy metals detection. Other research areas include nanotechnology, identification and validation of biomarkers, and development of biorecognition agents (antibodies and aptamers) and new biosensor design approaches (e.g., development of new materials) (12).

In recent literature, enzymes (1315) and whole cells (16) seem to have been replaced with antibodies (1723) and aptamers (2429) as recognition elements in electrochemical and optical biosensoristic devices.

Regarding electrochemical devices, they possess unique features to address the challenges of field and on-site analytical chemistry: possibility of miniaturization and portability, sensitivity, selectivity, a wide linear range of detection, minimal power requirement, and cost-effective instrumentation. Voltammetry is one of the most widely used electroanalytical techniques for electrochemical detection in bio/chemosensoric devices (see Tables S1 and S2 in Supplementary Material). In fact, various voltammetric techniques possess intrinsic analytical advantages and features and included excellent sensitivity, rapid analysis times, and possibility of simultaneous determination of different analytes. In voltammetric pulse techniques, through different modulation of the applied potential, a higher speed of measurement and sensitivity (useful for determination of species at trace levels) can be achieved. In particular, differential pulse voltammetry and square-wave voltammetry have been extensively described in the recent literature for detection of various chemical contaminants in environmental samples (24, 25, 3035). Other widely used electrochemical techniques includes cyclic voltammetry (for studies on redox behavior of analytes) (31, 36, 37) and stripping techniques (characterized by preconcentration step of the analyte onto or into the working electrode to achieve a greater sensitivity) (3842). In addition, the latter are commonly applied for determination of metal speciation (chemical form can influence bioavailability of metals) useful for environmental risk assessment of metal pollution (43). Amperometry is another widely used electrochemical technique in bio/chemosensoristics (1317, 4451). Together with voltammetric techniques, electrochemical impedance spectroscopy is an extremely useful technique for a broad range of applications, including characterization of materials and detecting interaction between recognition elements (e.g., antibodies and aptamers) of sensoristic devices and analyte, through measures of changes in electrical surface properties of electrodes (26, 28). To improve analytical features and performances of electrochemical techniques, the last decades have witnessed a tremendous development of innovative sensitive materials for surface functionalization of electrodes. Several advances in the development of bio/chemosensors (in particular for electrochemical devices) have been achieved through the employment of (modified) electrodes (Tables S1 and S2 in Supplementary Material). Traditional mercury-based electrodes (39, 41) have gradually been replaced (because of low mechanical stability and toxicity of mercury) by other electrodes made of better suitable materials. As replacement of mercury, alternative materials (with similar or better analytical features) have been employed and/or developed: bismuth (a non-toxic element with high hydrogen overpotential and good mechanical stability) (52), boron-doped diamond (with a wider electrochemical potential window and reduced fouling compared to traditional materials) (34), nitrogen-doped graphene (doping converts an excellent conductor as graphene into a p- or n-type semiconductor) (49), and single and multiwalled carbon nanotubes and nanoparticles. Looking at recent literature, a considerable attention has been paid to the development and exploitation of nanostructured materials (nanoparticles, nanowires, or nanotubes) for sensoristic purposes: carbon-based (e.g., single-walled and multiwalled carbon nanotubes) (17, 31, 35, 42, 45, 50) and nanoparticles with different chemical composition (13, 24, 2630, 32, 40, 50, 51). These nanomaterials (also functionalized) can modify surface architectures and functions of electrodes by, for instance, (i) enlarging active surface (e.g., increasing of docking sites for biological recognition elements) and (ii) enhancing electron transfer or electrical properties and amplify signals in general. Another interesting supramolecular-based approach to develop innovative materials for electrode modification is the synthesis of molecular and ion imprinted polymers. These are synthetic polymers able to mimicking biological recognition elements, like antibodies and aptamers, useful for the design of high-specificity sensoristic devices (30, 35, 38, 50, 53). Basically, these polymers are obtained from a copolymerization process of suitable monomers in the presence of a molecular or ionic template (the target analyte); the successive removal of the template leaves in the polymer structure binding sites that can re-host the analyte. Although they bring several advantages in terms of durability and cost-effective production (compared to aptamers and antibodies), it is still necessary to solve some problems related to heterogeneity of binding sites that can bring to non-specific bindings.

Optical Biosensoristic Devices and Lab-on-a-Chip (LOC)

The recent interest in optical biosensoristic devices for food analysis, with fluorescent, bioluminescent or chemiluminescent labels for detection, as well as the direct (label-free) detection (i.e., no reporter elements to generate a signal are needed) (54, 55), is increasing. The development of label-free technologies and in particular label-free surface plasmon resonance (SPR) has become the greatest example of employment of the technology as a routine analytical method in such fields (56).

Actually, biosensoristic devices based on SPR are ideal platforms for the label-free detection of molecular monolayers as they allow for qualitative and quantitative multiplexing measurements of biomolecular interactions in real-time without requiring a labeling procedure in the framework of food safety (57, 58).

Indeed, by using SPR-based immunosensors, one can obtain robust and quantitative results with narrow- or broad-spectrum specificity in relatively short time. In the case of milk, SPR circumvents the issues related to turbidity and protein fouling (both are generally limiting factors for optical-based biosensoristic devices application for milk testing) by measuring the refractive index modulation on the reverse side of the metal film where the biological selective element is immobilized (18, 59).

Since late 90s, SPR biosensoristic devices have become the main tool for the study of biomolecular interactions in life science, with successful applications in the field of food safety (58).

Although there are great advantages of the SPR technology, some disadvantages are evident: high cost of the readout instrumentation and a still high cost of the consumables (sensing chip and reagents) and large instrumentation footprint.

In recent years, nanoplasmonics (e.g., noble metal nanoparticles, nanometallic gratings, or a combination of metallic nanocavities organized in nanogratings) has shown a great potential in overcoming the technological/commercial limits of SPR (60) and for developing nanoplasmonic detection platforms. Even though important technological effort is still to be done to be competitive with point-of-care screening technologies, the integration on the same disposable and miniaturized platform of low-cost photonics devices with multiplexing nanoplasmonic and advanced microfluidics system can be considered as new and non-disruptive technology able to ensure competitiveness from both the economic and the detection point of views.

The development of advanced photonic biosensoristic devices has to be brought beyond the state of the art of the point-of-care-diagnostic systems by the synergetic integration of the different technological building blocks with consequent improvement of the single-component outputs. Moreover, the introduction of outperforming light-excitation/detection scheme allows for unraveling the potentiality of the sensor in terms for disposability, reliability, miniaturization, and multiplexing while providing laboratory quality analysis (61).

Within the current point-of-care diagnostic market, there is a limited number of systems that operate without the requirement for a dedicated desktop reader, and there are no quantitative, portable diagnostic platforms with multiple detection methods. The components from existing laboratory equipment are too bulky, fragile, and expensive and require too much mechanical integration to be consolidated into a point-of-care device (55).

Miniaturization (from microelectrodes/nanosensors to microfluidic platforms) is an increasing trend as a response to these needs to develop new miniature and portable analytical devices for environmental and food monitoring and diagnostics.

In this scenario, LOC devices have shown themselves to be highly effective for laboratory-based research, where their superior analytical performance has established them as efficient tools for complex tasks and a promising tools for a number of environmental monitoring applications, i.e., continuous surveillance of selected parameters and contaminant concentrations (62) and for agricultural and food safety (63). Referring to the state of the art in the recently developed LOC methods (64), it can be observed that they are based on nucleic acid amplification, biosensoristic devices, flow cytometry, spectrometry techniques, and multisensors systems.

However, to date, they have not been well suited to point-of-care or in-the-field applications: although the chips themselves are cheap and small, they must generally be used in conjunction with bulky optical detectors, which are needed to identify or quantify the analytes or reagents present. Furthermore, most existing detectors are limited to analysis of a single analyte at a predetermined location on the chip. The lack of an integrated, multiplexing, and fast detection scheme (one which is miniaturized, integrated, and able to monitor multiple locations on the chip) is a major obstacle to the deployment of diagnostic devices in the field. This issue has prevented the development of more complex tests where rapid, kinetic, or multipoint analysis is required.

Conclusion

Development of improved electrochemical and optical bio/chemosensoristic devices represents a technological challenge to broaden boundaries of field diagnostics and monitoring environmental and food samples. In particular, specific improved features of integration, portability (e.g., sensors equipped with built-in reading systems), cheapness, simplification of experimental protocols (less time- and labor-demanding protocols), and development of efficient high-throughput approaches are required. Concerning LOC devices, fast detection scheme and the ability to monitor at multiple locations on the chip could ensure a high selectivity and sensitivity for the analyte of interest. All these devices could be employed for the identification of new quality and safety marker indexes in real matrices as well as for the determination of mixture effects of bioactive substances.

Author Contributions

All authors have made equal contributions in the writing and revising of this mini-review.

Conflict of Interest Statement

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Supplementary Material

The Supplementary Material for this article can be found online at http://journal.frontiersin.org/article/10.3389/fpubh.2017.00080/full#supplementary-material.

References

1. Bertrand PG, Ahmed HAM, Ngwafor R, Frazzoli C. Toxicovigilance systems and practices in Africa. Toxics (2016) 4(3):13. doi: 10.3390/toxics4030013

CrossRef Full Text | Google Scholar

2. Appicciafuoco B, Dragone R, Frazzoli C, Bolzoni G, Mantovani A, Ferrini AM. Microbial screening for quinolones residues in cow milk by bio-optical method. J Pharm Biomed Anal (2015) 106:179–85. doi:10.1016/j.jpba.2014.11.037

PubMed Abstract | CrossRef Full Text | Google Scholar

3. Kalunke RM, Grasso G, D’Ovidio R, Dragone R, Frazzoli C. Detection of ciprofloxacin residues in cow milk: a novel and rapid optical β-galactosidase-based screening assay. Microchem J (2016). doi:10.1016/j.microc.2016.12.014

CrossRef Full Text | Google Scholar

4. Dragone R, Grasso G. Biosensoristic devices: monitoring and diagnostics in agro-zootechnical productions. In: Frazzoli C, Asongalem AE, Orisakwe OE, editors. Cameroon-Nigeria-Italy Scientific Cooperation: Veterinary Public Health and Sustainable Food Safety to Promote “One Health/One Prevention”. (Vol. 12). Rome: Istituto Superiore di Sanità (2012). 70 p.

Google Scholar

5. Dragone R, Ermilov L, Grasso G, Maggioni S, Mantovani A, Frazzoli C. Antioxidant power as biochemical endpoint in bread for screening and early managing quality and toxicant-related safety anomalies in food production. Food Chem Toxicol (2016) 94:31–8. doi:10.1016/j.fct.2016.04.028

PubMed Abstract | CrossRef Full Text | Google Scholar

6. Frazzoli C, Mantovani A, Campanella L, Dragone R. System (Best-Integrated Toxicity (Bio) Sensor’s System for Hazard Analysis and Management in the Food Chain and the Environment) for Environment Diagnostic and Monitoring and Self-Control of Food Chain, Including Primary Production, and Relevant Method for Quick Management of Hazard. International Publication number EP2304428 A1 (2011).

Google Scholar

7. Frazzoli C, Mantovani A, Dragone R. Local role of food producers’ communities for a Global One-Health framework: the experience of translational research in an Italian dairy chain. J Agric Chem Environ (2014) 3(02):14. doi:10.4236/jacen.2014.32B003

CrossRef Full Text | Google Scholar

8. Dragone R, Frazzoli C, Monacelli F. Chemical-Physical Sensing Device for Chemical – Toxicological Diagnostics in Real Matrices. International Publication number WO2012140685 A1 (2012).

Google Scholar

9. Thévenot DR, Toth K, Durst RA, Wilson GS. Electrochemical biosensors: recommended definitions and classification. Pure Appl Chem (1999) 71(12):2333–48. doi:10.1351/pac199971122333

PubMed Abstract | CrossRef Full Text | Google Scholar

10. Narsaiah K, Jha SN, Bhardwaj R, Sharma R, Kumar R. Optical biosensors for food quality and safety assurance – a review. J Food Sci Technol (2012) 49(4):383–406. doi:10.1007/s13197-011-0437-6

CrossRef Full Text | Google Scholar

11. Rotariu L, Lagarde F, Jaffrezic-Renault N, Bala C. Electrochemical biosensors for fast detection of food contaminants – trends and perspective. TrAC Trends Anal Chem (2016) 79:80–7. doi:10.1016/j.trac.2015.12.017

CrossRef Full Text | Google Scholar

12. Fogel R, Limson J. Developing biosensors in developing countries: South Africa as a case study. Biosensors (2016) 6(1):5. doi:10.3390/bios6010005

CrossRef Full Text | Google Scholar

13. Chauhan N, Pundir CS. An amperometric biosensor based on acetylcholinesterase immobilized onto iron oxide nanoparticles/multi-walled carbon nanotubes modified gold electrode for measurement of organophosphorus insecticides. Anal Chim Acta (2011) 701(1):66–74. doi:10.1016/j.aca.2011.06.014

PubMed Abstract | CrossRef Full Text | Google Scholar

14. Fusco M, Federico R, Boffi A, Macone A, Favero G, Mazzei F. Characterization and application of a diamine oxidase from as component of an electrochemical biosensor for the determination of biogenic amines in wine and beer. Anal Bioanal Chem (2011) 2(401):707–16. doi:10.1007/s00216-011-5131-z

CrossRef Full Text | Google Scholar

15. Sanllorente-Méndez S, Domínguez-Renedo O, Arcos-Martínez MJ. Immobilization of acetylcholinesterase on screen-printed electrodes. Application to the determination of arsenic (III). Sensors (Basel) (2010) 10(3):2119–28. doi:10.3390/s100302119

PubMed Abstract | CrossRef Full Text | Google Scholar

16. Cortés-Salazar F, Beggah S, van der Meer JR, Girault HH. Electrochemical As(III) whole-cell based biochip sensor. Biosens Bioelectron (2013) 47:237–42. doi:10.1016/j.bios.2013.03.011

PubMed Abstract | CrossRef Full Text | Google Scholar

17. Belkhamssa N, Justino CI, Santos PS, Cardoso S, Lopes I, Duarte AC, et al. Label-free disposable immunosensor for detection of atrazine. Talanta (2016) 146:430–4. doi:10.1016/j.talanta.2015.09.015

PubMed Abstract | CrossRef Full Text | Google Scholar

18. Fernández F, Hegnerová K, Piliarik M, Sanchez-Baeza F, Homola J, Marco MP. A label-free and portable multichannel surface plasmon resonance immunosensor for on site analysis of antibiotics in milk samples. Biosens Bioelectron (2010) 26(4):1231–8. doi:10.1016/j.bios.2010.06.012

CrossRef Full Text | Google Scholar

19. Hirakawa Y, Yamasaki T, Watanabe E, Okazaki F, Murakami-Yamaguchi Y, Oda M, et al. Development of an immunosensor for determination of the fungicide chlorothalonil in vegetables, using surface plasmon resonance. J Agric Food Chem (2015) 63(28):6325–30. doi:10.1021/acs.jafc.5b01980

PubMed Abstract | CrossRef Full Text | Google Scholar

20. Mauriz E, Calle A, Abad A, Montoya A, Hildebrandt A, Barceló D, et al. Determination of carbaryl in natural water samples by a surface plasmon resonance flow-through immunosensor. Biosens Bioelectron (2006) 21(11):2129–36. doi:10.1016/j.bios.2005.10.013

PubMed Abstract | CrossRef Full Text | Google Scholar

21. Sharma A, Catanante G, Hayat A, Istamboulie G, Rejeb IB, Bhand S, et al. Development of structure switching aptamer assay for detection of aflatoxin M1 in milk sample. Talanta (2016) 158:35–41. doi:10.1016/j.talanta.2016.05.043

PubMed Abstract | CrossRef Full Text | Google Scholar

22. Tomassetti M, Martini E, Campanella L, Favero G, Sanzó G, Mazzei F. A new surface plasmon resonance immunosensor for triazine pesticide determination in bovine milk: a comparison with conventional amperometric and screen-printed immunodevices. Sensors (Basel) (2015) 15(5):10255–70. doi:10.3390/s150510255

PubMed Abstract | CrossRef Full Text | Google Scholar

23. Wu H, Li H, Chua FZ, Li SFY. Rapid detection of melamine based on immunoassay using portable surface plasmon resonance biosensor. Sens Actuators B Chem (2013) 178:541–6. doi:10.1016/j.snb.2012.12.089

CrossRef Full Text | Google Scholar

24. Chen Z, Li L, Mu X, Zhao H, Guo L. Electrochemical aptasensor for detection of copper based on a reagentless signal-on architecture and amplification by gold nanoparticles. Talanta (2011) 85:730–5. doi:10.1016/j.talanta.2011.04.056

PubMed Abstract | CrossRef Full Text | Google Scholar

25. Cui L, Wu J, Ju H. Label-free signal-on aptasensor for sensitive electrochemical detection of arsenite. Biosens Bioelectron (2016) 79:861–5. doi:10.1016/j.bios.2016.01.010

PubMed Abstract | CrossRef Full Text | Google Scholar

26. Fan L, Zhao G, Shi H, Liu M, Li Z. A highly selective electrochemical impedance spectroscopy-based aptasensor for sensitive detection of acetamiprid. Biosens Bioelectron (2013) 43:12–8. doi:10.1016/j.bios.2012.11.033

PubMed Abstract | CrossRef Full Text | Google Scholar

27. Fan L, Zhao G, Shi H, Liu M, Wang Y, Ke H. A femtomolar level and highly selective 17β-estradiol photoelectrochemical aptasensor applied in environmental water samples analysis. Environ Sci Technol (2014) 48(10):5754–61. doi:10.1021/es405685y

CrossRef Full Text | Google Scholar

28. Rivas L, Mayorga-Martinez CC, Quesada-González D, Zamora-Gálvez A, de la Escosura-Muñiz A, Merkoçi A. Label-free impedimetric aptasensor for ochratoxin-A detection using iridium oxide nanoparticles. Anal Chem (2015) 87(10):5167–72. doi:10.1021/acs.analchem.5b00890

PubMed Abstract | CrossRef Full Text | Google Scholar

29. Rhouati A, Yang C, Hayat A, Marty JL. Aptamers: a promising tool for ochratoxin A detection in food analysis. Toxins (Basel) (2013) 5(11):1988–2008. doi:10.3390/toxins5111988

CrossRef Full Text | Google Scholar

30. Chen H, Zhang Z, Cai R, Rao W, Long F. Molecularly imprinted electrochemical sensor based on nickel nanoparticles-graphene nanocomposites modified electrode for determination of tetrabromobisphenol A. Electrochim Acta (2014) 117:385–92. doi:10.1016/j.electacta.2013.11.185

CrossRef Full Text | Google Scholar

31. Fayemi OE, Adekunle AS, Ebenso EE. A sensor for the determination of lindane using PANI/Zn, Fe (III) oxides and nylon 6, 6/MWCNT/Zn, Fe (III) oxides nanofibers modified glassy carbon electrode. J Nanomater (2016) 2016:1–10. doi:10.1155/2016/4049730

CrossRef Full Text | Google Scholar

32. Karuppiah C, Muthupandi K, Chen SM, Ali MA, Palanisamy S, Rajan A, et al. Green synthesized silver nanoparticles decorated on reduced graphene oxide for enhanced electrochemical sensing of nitrobenzene in waste water samples. RSC Adv (2015) 5(39):31139–46. doi:10.1039/C5RA00992H

CrossRef Full Text | Google Scholar

33. Rather JA, De Wael K. Fullerene-C 60 sensor for ultra-high sensitive detection of bisphenol-A and its treatment by green technology. Sens Actuators B Chem (2013) 176:110–7. doi:10.1016/j.snb.2012.08.081

CrossRef Full Text | Google Scholar

34. Švorc Ľ, Rievaj M, Bustin D. Green electrochemical sensor for environmental monitoring of pesticides: determination of atrazine in river waters using a boron-doped diamond electrode. Sens Actuators B Chem (2013) 181:294–300. doi:10.1016/j.snb.2013.02.036

CrossRef Full Text | Google Scholar

35. Wong A, Foguel MV, Khan S, de Oliveira FM, Tarley CRT, Sotomayor MD. Development of an electrochemical sensor modified with MWCNT-COOH and MIP for detection of diuron. Electrochim Acta (2015) 182:122–30. doi:10.1016/j.electacta.2015.09.054

CrossRef Full Text | Google Scholar

36. Cinti S, Talarico D, Palleschi G, Moscone D, Arduini F. Novel reagentless paper-based screen-printed electrochemical sensor to detect phosphate. Anal Chim Acta (2016) 919:78–84. doi:10.1016/j.aca.2016.03.011

PubMed Abstract | CrossRef Full Text | Google Scholar

37. Masikini M, Waryo TT, Baker PG, Ngqongwa LV, Williams AR, Iwuoha EI. Hydroxy-iron/β-cyclodextrin-film amperometric sensor for the endocrine disruptor substance bisphenol-A in an aqueous medium with reduced fouling effects. Anal Lett (2011) 44(11):2047–60. doi:10.1080/00032719.2010.539741

CrossRef Full Text | Google Scholar

38. Alizadeh T, Amjadi S. Preparation of nano-sized Pb2+ imprinted polymer and its application as the chemical interface of an electrochemical sensor for toxic lead determination in different real samples. J Hazard Mater (2011) 190(1):451–9. doi:10.1016/j.jhazmat.2011.03.067

PubMed Abstract | CrossRef Full Text | Google Scholar

39. Illuminati S, Annibaldi A, Truzzi C, Finale C, Scarponi G. Square-wave anodic-stripping voltammetric determination of Cd, Pb and Cu in wine: set-up and optimization of sample pre-treatment and instrumental parameters. Electrochim Acta (2013) 104:148–61. doi:10.1016/j.electacta.2013.04.001

CrossRef Full Text | Google Scholar

40. Niu P, Fernández-Sánchez C, Gich M, Ayora C, Roig A. Electroanalytical assessment of heavy metals in waters with bismuth nanoparticle-porous carbon paste electrodes. Electrochim Acta (2015) 165:155–61. doi:10.1016/j.electacta.2015.03.001

CrossRef Full Text | Google Scholar

41. Shahbazi Y, Ahmadi F, Fakhari F. Voltammetric determination of Pb, Cd, Zn, Cu and Se in milk and dairy products collected from Iran: an emphasis on permissible limits and risk assessment of exposure to heavy metals. Food Chem (2016) 192:1060–7. doi:10.1016/j.foodchem.2015.07.123

CrossRef Full Text | Google Scholar

42. Wong A, Scontri M, Materon EM, Lanza MR, Sotomayor MD. Development and application of an electrochemical sensor modified with multi-walled carbon nanotubes and graphene oxide for the sensitive and selective detection of tetracycline. J Electroanal Chem (2015) 757:250–7. doi:10.1016/j.jelechem.2015.10.001

CrossRef Full Text | Google Scholar

43. Janssen CR, Heijerick DG, De Schamphelaere KAC, Allen HE. Environmental risk assessment of metals: tools for incorporating bioavailability. Environ Int (2003) 28(8):793–800. doi:10.1016/S0160-4120(02)00126-5

PubMed Abstract | CrossRef Full Text | Google Scholar

44. Amati L, Campanella L, Dragone R, Nuccilli A, Tomassetti M, Vecchio S. New investigation of the isothermal oxidation of extra virgin olive oil: determination of free radicals, total polyphenols, total antioxidant capacity, and kinetic data. J Agric Food Chem (2008) 56(18):8287–95. doi:10.1021/jf8011695

PubMed Abstract | CrossRef Full Text | Google Scholar

45. Cámara-Martos F, da Costa J, Justino CI, Cardoso S, Duarte AC, Rocha-Santos T. Disposable biosensor for detection of iron (III) in wines. Talanta (2016) 154:80–4. doi:10.1016/j.talanta.2016.03.057

PubMed Abstract | CrossRef Full Text | Google Scholar

46. Dragone R, Frazzoli C, Grappelli C, Campanella L. A new respirometric endpoint-based biosensor to assess the relative toxicity of chemicals on immobilized human cells. Ecotoxicol Environ Saf (2009) 72(1):273–9. doi:10.1016/j.ecoenv.2008.02.011

PubMed Abstract | CrossRef Full Text | Google Scholar

47. Dragone R, Frazzoli C, Grasso G, Rossi G. Sensor with intact or modified yeast cells as rapid device for toxicological test of chemicals. J Agric Chem Environ (2014) 3(02):35. doi:10.4236/jacen.2014.32B006

CrossRef Full Text | Google Scholar

48. Dragone R, Cheng R, Grasso G, Frazzoli C. Diuron in water: functional toxicity and intracellular detoxification patterns of active concentrations assayed in tandem by a yeast-based probe. Int J Environ Res Public Health (2015) 12(4):3731–40. doi:10.3390/ijerph120403731

PubMed Abstract | CrossRef Full Text | Google Scholar

49. Fan H, Li Y, Wu D, Ma H, Mao K, Fan D, et al. Electrochemical bisphenol A sensor based on N-doped graphene sheets. Anal Chim Acta (2012) 711:24–8. doi:10.1016/j.aca.2011.10.051

PubMed Abstract | CrossRef Full Text | Google Scholar

50. Huang J, Zhang X, Lin Q, He X, Xing X, Huai H, et al. Electrochemical sensor based on imprinted sol-gel and nanomaterials for sensitive determination of bisphenol A. Food Control (2011) 22(5):786–91. doi:10.1016/j.foodcont.2010.11.017

CrossRef Full Text | Google Scholar

51. Yin H, Cui L, Chen Q, Shi W, Ai S, Zhu L, et al. Amperometric determination of bisphenol A in milk using PAMAM–Fe 3 O 4 modified glassy carbon electrode. Food Chem (2011) 125(3):1097–103. doi:10.1016/j.foodchem.2010.09.098

CrossRef Full Text | Google Scholar

52. Arduini F, Calvo JQ, Palleschi G, Moscone D, Amine A. Bismuth-modified electrodes for lead detection. TrAC Trends Anal Chem (2010) 29(11):1295–304. doi:10.1016/j.trac.2010.08.003

CrossRef Full Text | Google Scholar

53. Bougrini M, Florea A, Cristea C, Sandulescu R, Vocanson F, Errachid A, et al. Development of a novel sensitive molecularly imprinted polymer sensor based on electropolymerization of a microporous-metal-organic framework for tetracycline detection in honey. Food Control (2016) 59:424–9. doi:10.1016/j.foodcont.2015.06.002

CrossRef Full Text | Google Scholar

54. Morrow KJ. Label-free biosensors show early promise – optical and nonoptical platforms are on the heels of conventional ELISA. Genet Eng Biotechnol News (2007) 27:4.

Google Scholar

55. Capitan-Vallvey LF, Palma AJ. Recent developments in handheld and portable optosensing – a review. Anal Chim Acta (2011) 696(1):27–46. doi:10.1016/j.aca.2011.04.005

PubMed Abstract | CrossRef Full Text | Google Scholar

56. Bottazzi B, Fornasari L, Frangolho A, Giudicatti S, Mantovani A, Marabelli F, et al. Multiplexed label-free optical biosensor for medical diagnostics. J Biomed Opt (2014) 19(1):017006. doi:10.1117/1.JBO.19.1.017006

PubMed Abstract | CrossRef Full Text | Google Scholar

57. Kim SJ, Gobi KV, Harada R, Shankaran DR, Miura N. Miniaturized portable surface plasmon resonance immunosensor applicable for on-site detection of low-molecular-weight analytes. Sens Actuators B Chem (2006) 115(1):349–56. doi:10.1016/j.snb.2005.09.025

CrossRef Full Text | Google Scholar

58. Piliarik M, Vala M, Tichý I, Homola J. Compact and low-cost biosensor based on novel approach to spectroscopy of surface plasmons. Biosens Bioelectron (2009) 24(12):3430–5. doi:10.1016/j.bios.2008.11.003

PubMed Abstract | CrossRef Full Text | Google Scholar

59. Mazumdar SD, Hartmann M, Kämpfer P, Keusgen M. Rapid method for detection of Salmonella in milk by surface plasmon resonance (SPR). Biosens Bioelectron (2007) 22(9):2040–6. doi:10.1016/j.bios.2006.09.004

PubMed Abstract | CrossRef Full Text | Google Scholar

60. Estevez MC, Otte MA, Sepulveda B, Lechuga LM. Trends and challenges of refractometric nanoplasmonic biosensors: a review. Anal Chim Acta (2014) 806:55–73. doi:10.1016/j.aca.2013.10.048

PubMed Abstract | CrossRef Full Text | Google Scholar

61. Ryu G, Huang J, Hofmann O, Walshe CA, Sze JY, McClean GD, et al. Highly sensitive fluorescence detection system for microfluidic lab-on-a-chip. Lab Chip (2011) 11(9):1664–70. doi:10.1039/c0lc00586j

PubMed Abstract | CrossRef Full Text | Google Scholar

62. Campos CD, da Silva JA. Applications of autonomous microfluidic systems in environmental monitoring. RSC Adv (2013) 3(40):18216–27. doi:10.1039/c3ra41561a

CrossRef Full Text | Google Scholar

63. Guo L, Feng J, Fang Z, Xu J, Lu X. Application of microfluidic “lab-on-a-chip” for the detection of mycotoxins in foods. Trends Food Sci Technol (2015) 46(2):252–63. doi:10.1016/j.tifs.2015.09.005

CrossRef Full Text | Google Scholar

64. Lafleur JP, Jönsson A, Senkbeil S, Kutter JP. Recent advances in lab-on-a-chip for biosensing applications. Biosens Bioelectron (2016) 76:213–33. doi:10.1016/j.bios.2015.08.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: agro-food supply chain, milk, on-site diagnostics, electroanalytical methods, biosensoristic devices, surface plasmon resonance, lab-on-a-chip

Citation: Dragone R, Grasso G, Muccini M and Toffanin S (2017) Portable Bio/Chemosensoristic Devices: Innovative Systems for Environmental Health and Food Safety Diagnostics. Front. Public Health 5:80. doi: 10.3389/fpubh.2017.00080

Received: 14 November 2016; Accepted: 29 March 2017;
Published: 05 May 2017

Edited by:

Mohiuddin Md. Taimur Khan, Washington State University, USA

Reviewed by:

Dr. Md. Asaduzzaman, University of Dhaka, Bangladesh
Tanvir Ahmed, Massachusetts Institute of Technology, USA

Copyright: © 2017 Dragone, Grasso, Muccini and Toffanin. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) or licensor are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Gerardo Grasso, gerardo.grasso@ismn.cnr.it

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.