Next Article in Journal
A Fluorescent Thermometer Based on a Pyrene-Labeled Thermoresponsive Polymer
Previous Article in Journal
Flexible Temperature Sensors on Fibers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Trends in Monitoring of European Water Framework Directive Priority Substances Using Micro-Sensors: A 2007–2009 Review

by
Philippe Namour
1,2,*,
Mathieu Lepot
3 and
Nicole Jaffrezic-Renault
1
1
Université de Lyon, Laboratory of Analytical Sciences, UMR CNRS 5180, 43 boulevard 11 novembre 1918, F-69622, Villeurbanne cedex, France
2
Cemagref, UR MALY, CP 220, F-69336, Lyon cedex 09, France
3
Université de Lyon, INSA Lyon, LGCIE, 34 Avenue des arts, F-69621 Villeurbanne Cedex, France
*
Author to whom correspondence should be addressed.
Sensors 2010, 10(9), 7947-7978; https://doi.org/10.3390/s100907947
Submission received: 12 June 2010 / Revised: 21 July 2010 / Accepted: 9 August 2010 / Published: 26 August 2010
(This article belongs to the Section Chemical Sensors)

Abstract

: This review discusses from a critical perspective the development of new sensors for the measurement of priority pollutants targeted in the E.U. Water Framework Directive. Significant advances are reported in the paper and their advantages and limitations are also discussed. Future perspectives in this area are also pointed out in the conclusions. This review covers publications appeared since December 2006 (the publication date of the Swift report). Among priority substances, sensors for monitoring the four WFD metals represent 81% of published papers. None of analyzed publications present a micro-sensor totally validated in laboratory, ready for tests under real conditions in the field. The researches are mainly focused on the sensing part of the micro-sensors. Nevertheless, the main factor limiting micro-sensor applications in the environment is the ruggedness of the receptor towards environmental conditions. This point constitutes the first technological obstacle to be overcome for any long-term field tests.

1. Introduction

The Water Framework Directive (WFD) governs European water policy. WFD has been in place as the main European regulation for the protection of water resources and the water environment since 2000 [1]. One of its principal objectives is to achieve good chemical and ecological status and to restore water bodies to a ‘good status’ by 2015. WFD requires management of water bodies so that the water quality does not affect their ecological services. Chemical status refers to specific pollutants (e.g., priority substances or priority hazardous substances) for which environmental quality standards (EQS) are proposed and defined for pollutants as minimum requirements [2].

As the WFD implementation gradually comes into effect in European countries, the environmental metrology market is bound to increase over the coming years. In view of the high cost of the laboratory analyses required and the potential artifacts that may be introduced during the conventional sampling protocol: “send a technician, take an isolated sample, send it to the laboratory and analyze it” with at best, in conventional water chemistry, a 24-hours flow proportional composite sample, a new type of field analysis must be designed. An average sample obscures the essentially dynamic character of a polluting event and average contents are devoid of any ecological realism. Biocenoses in rivers are never exposed to average contents, which have no actual existence as far as they are concerned. They are in reality exposed to changes in their physicochemical ambience. A first monitoring of dissolved copper, lead and cadmium with a submersible voltametric flow cell for periods of four days in coastal waters shows that the potentially most toxic forms of metals may vary in concentration on a time scale of less than one hour [3]. This first data confirm the poor ecological relevance of the conventional sampling protocol and the necessity of a continuous measurement. The greater and more sudden these changes, the more disturbance they cause. In terms of toxicology, fluctuation is a more important parameter than the average level, and in the present case the concentration peak reached by the pollutant is more important than the average concentration. The ecological pertinence of such a procedure is more than problematic [4]. New environmental monitoring strategies must be designed. A possible avenue that merits further exploration involves the deployment of low-cost instrumentation allowing massive data logging, as well as tools for subsequent data validation, management and interpretation. However, while this new type of instrumentation is possible, and even desirable, such deployment cannot at the present time cover all the WFD’s parameters and there is little probability of this situation changing before 2015, or even 2021. Consequently, faced with the magnitude of this metrological challenge and the urgency of the situation, a paradigm shift is required in order to imagine a new approach to the problem of water monitoring. Given this situation, current research on micro-sensors is leading to the emergence of many measuring principles. The Swift report ( http://www.swift-wfd.com), published in December 2006, lists a wide range of monitoring methods currently available or under development for supporting the WFD. This review covers publications after December 2006, the publication date of the Swift report. It then covers the period extending from January, 2007 to December 2009, and concerns micro-sensors of so called priority substances from the list quoted in appendix II of the Common Decision nº3/2008 of December 20th, 2007 [6]. This review, although not being able to aspire to the exhaustiveness, constitutes a relatively exact panorama of the recent research efforts in the field of the measurement of priority substances in the water using micro-sensors.

For practical reasons, with the analyte constituting the access key for chemical analysts, micro-sensors were firstly classified according to the measured analyte and then, for one analyte, according to the measurement principle or transduction mode: electrochemical and optical. The papers will then be presented according to the following classification: first the measured analyte then, when several technologies are used, they were grouped by type. Finally we tried to focus this review mainly on the micro-sensors at the field validation stage, that is the systems having passed lab tests and susceptible to field validation. The criteria of publication analysis are: the type of aquatic environment studied, the detection limit and the reproducibility.

Finally, with a concern for sustainable development and the eco-design of instruments installed in natural environment, this review does not take into account publications presenting methods involving toxic compounds, such as “heavy metals”, in particular mercury impregnation or films, even if the quantities used are relatively low. Indeed, substances such as cadmium or mercury have been classified as ‘priority hazardous substances’ in the Decision nº2455/2001/EC [5] and Directive 2008/32/CE [6], for which Member States should implement necessary measures with the aim of ceasing or phasing out emissions, discharges and losses into water of those priority hazardous substances which derive from human activities. So it is preferable to banish these hazardous substances from our devices, rather than quibble over low or negligible implemented quantities, and to be vigilant about the potential toxicity of any new substances used in our devices.

Furthermore, development of new devices using priority hazardous substances leads to a commercial dead-end, and a waste of time and money: they cannot be used in Europe, and even in the other parts of the world. Indeed, mercury is recognized as a chemical of global concern. US EPA’s Roadmap for Mercury (5 July 2006) promotes reducing mercury in processes and products, even where cost-effective substitutes do not exist. The overall goal of the Global Mercury Partnership of United Nations Environment Programme (Governing Council Decision 25/5, Nairobi, Kenya, 16–20 February 2009) is to reduce and eventually eliminate mercury use in products and processes, and raising awareness of mercury-free alternatives. Among these products electric and electronic devices are targeted.

2. Reviews & Books

Some recent reviews were published during the years 2007–2009. Johnson et al. [7] refocus problems posed by the punctual sampling of dynamic systems such as rivers. The authors insist on the necessity of high frequency and continuous measures in time, as well as on the obligation of taking into account spatial variations. They also regret the current absence, for the most part, of simple, selective and sensitive enough sensors for the chemical elements of interest to the geochemist operating without drift over long periods. Pejcic et al. [8] surveyed some of the work that has been undertaken using sensors to detect hydrocarbons. Of the various transducers examined in this review the optical-based sensor appears to be the most promising in terms of water monitoring. The Bosch et al. review [9] in the journal Sensors, is about the use of optical fibre micro-biosensors, mainly in biological applications. Let us note a section dedicated to the measurement of metals via cells or immobilized enzymes and micro-biosensors detecting PCB or pesticides via antibodies immobilized on silica fibres. Also let us quote the Lieberzeit and Dickert review [10] dedicated to the sensors applicable to the environmental chemistry, operational in the field, during these last five years. The authors end their paper with the very scarce literature published on actually operational sensors, albeit with some breakthroughs in the metal measurement domain. The development of simple and strong optical or electrochemical prototypes adapted to the field conditions, will permit one to envisage actual field monitoring in a near future. Clare Reimers’ review [11] is more particularly centred on the usage of glass microelectrodes in oceanography. A section is dedicated to the study of metal speciation in the marine environment with minimal sampling artefacts.

Two articles review the challenges associated with metal speciation studies, and focus on voltammetric techniques for the in situ study of metal speciation. More specifically, they summarize the specific conceptual, analytical, and technical criteria that must be considered to develop rugged, field deployable, non-perturbing sensors and probes [12,13]. Lange et al. [14] overviewed surface acoustic wave-based biosensor technology, and a small section was devoted to atrazine and PAH detection. The review of Moreno-Garrido [15] draws a general picture of the advances in immobilization techniques and biotechnology, using freshwater and marine micro-algae, in environmental aquatic research for the toxicity measurement of substance or effluent. Jones and Compton’s review [16] summarizes works using boron-doped diamond (BDD) electrodes for stripping voltammetry, particularly for measurements of toxic metals such as cadmium or lead. BDD does not seem to be a suitable mercury substitute in all cases, but is one of the more widely applicable alternative electrode materials in comparison to glassy carbon or solid metal electrodes. Jaffrezic-Renault and Dzyadevych [17] presented the application of the conductometric measurement method to environmental monitoring and specifications obtained for the detection of different pesticides, herbicides and heavy metal ions, based on enzyme inhibition.

In a detailed review, Seidel and Niessner [18] described microarray techniques for the simultaneous detection of multiple analytes and presented some environmental applications such as pesticides, toxins and endocrine-disrupting compound detection in water. Recent advances in the development and applications of nucleic acid-based biosensors for environmental application are reviewed by Palchetti and Mascini [19], with special emphasis on functional nucleic acid elements and the detection of chemically-induced DNA damage caused by genotoxic pollutants as some PAH, pesticides or solvents. A chapter of book “Antiterrorism and Homeland Defense: Polymers and Materials” is devoted to porphyrin-enzyme complexes for the detection of organophosphates using evanescent wave absorbance spectroscopy (EWAS) [20]. Nolan and Lippard [21] provided a very comprehensive account of progress in the design and application of optical sensors for mercury, in the environmental and biological media, up until July 2007.

Diamond et al. [22] present the concept of ‘wireless sensor networks’. A part of their review is dedicated to remote environmental monitoring and highlights limitations with the current manifestations of these platforms, particularly with respect to integration of chemo-/biosensing capabilities. In the same way, Strobl and Robillard [23] give insight into the strategies of water quality monitoring network design for surface freshwaters, along with their weaknesses and shortcomings. They propose nine recommendations for the successful development of a methodology for designing monitoring networks. A section of Bogues’ article [24] gives a brief overview of the environmental sensing technologies in water quality monitoring and considers briefly the research effort.

Jaffrezic-Renault [25] presents in Actualité Chimique different kinds of electrochemical micro-sensors allowing the in situ monitoring of pollutants in waters: heavy metals, ammonium, total organic carbon, organophosphate pesticides. The review of Baruah and Dutta [26] on nanotechnology applications in pollution sensing in agriculture and environment contains an important section on the measurement of pollutants by means of nano-structured materials. This review is far from exhaustive concerning the topic of micro-sensors and certain references are more than 10 years old. Pesavento et al., [27] presented the electrochemical methods for metal speciation in natural waters, mainly by means of voltammetry. It is noteworthy that a section is dedicated to in situ measurements, especially focused on practical needs. The single micro-sensor presented is a gel electrode containing mercury. The second part of this review deals with alternative procedures as separation techniques based on ion exchange and complexing resins, and micro separation methods as Donnan membrane technique, diffusive gradients in thin-film gels and permeation liquid membrane. Reardon et al.’s review [28] focuses on photoluminescence-based biosensors of potential use in the environmental applications. They give examples of optical biosensors for organics (aromatic solvents, PAH, organophosphate or halogenated pesticides) and metal determinations (Hg, Cd, Pb and Ni), however these references are previous to 2007. Leray and Value [29], after a summary of the main classes of fluorescence molecular receptors for the detection of toxic metal ions (Photoinduced Electron or Charge Transfer, Excimer Formation or Disappearance and Energy Transfer), present different calixarene-based fluoroionophores for the detection of caesium, mercury, lead and cadmium. A review of Selid et al. [30] examines different useful micro- and nano-sensors for mercury determination, mainly in healthcare and in a lesser extent, in the environmental field. Finally, David et al. [31] published a review on alkylphenols in marine and coastal environment, mainly centred on the environment contamination, however a section is dedicated to their detection in environmental matrices by biological tests on cells but not strictly involving micro-sensors as we understand them.

3. Scientific Articles

3.1. Metals

The WFD targets four priority metals: cadmium, mercury, nickel and lead [1]. Table 1 summarizes the various target values for freshwater water bodies. The strongest constraints are given by the SEQ-Eau, a water quality evaluation system developed by French Water Agencies in order to maintain the water biological balances in an optimal way. The detection limits of micro-sensors should be equal or lower than SEQ-Eau target values, although they do not represent statutory values, as the limit values of the Order 2001-1220 are only imperative for drinking water.

The analysis of the publications from 2007 to 2009 in the field of micro-sensors, shows that research efforts mainly concern the determination of WFD metals in waters (126 papers), the detection of WFD organic substances being the subject of less attention (24 papers), then 83% of the papers concern only 17% of priority substances. The sensors for metal detection divide up roughly equally between electrochemical (56%) and optical (44%) techniques, with strong disparities according to the metal concerned. As a general rule the electrochemical methods are preferred for the detection in water of cadmium (86%), nickel (67%) and lead (86%). On the other hand optical methods are proposed for mercury detection in 73% of the papers.

Approximately twenty articles deal with the continuous determination of cadmium in water, mainly (81%) by electrochemical techniques (Table 2). Only five articles about optical techniques for cadmium determination were published from 2007 to 2009 (Table 3).

Around 50% of electrochemical sensor display a detection limit for cadmium below 1 μg/L [3441] on modified carbon electrodes. Other micro-sensors give detection limits ranging from 1 μg/L [4246] to 100–120 μg/L [60] and 500 μg/L [62] on enzymatic biosensors, 3.6 μg/L using anodic stripping voltammetry on screen printed electrodes doped with bismuth nanoparticles [54] and 20 μg/L by differential pulse voltammetry through polyvinylchloride-2-nitrophenylethyl ether gel [58].

Optical sensors are two or three magnitude order less sensitive than electrochemical sensors. The most sensitive optical sensor displays a detection limit equal to 60 μg/L [63], then are following: 130 μg/L for fluorescent sensor [64] and 560 μg/L for the virtual sensor [65].

Even the detection limits of electrochemical sensors are still too high for water monitoring. The EQS values recommended by the WFD in freshwater range from 0.08 μg/L to 0.25 μg/L according to water hardness; INERIS publishes for cadmium, a PNEC equal to 0.21 μg/L [68] and the SEQ-Eau target value for an optimal water quality goes from 1 ng/L to 9 ng/L, according to water hardness [32]. Finally none of these micro-sensors was validated in a natural environment.

3.1.1. Mercury

Contrary to cadmium determination, for which electrochemistry is widely used, researchers seem to have preferred optical methods for mercury determination in water. Fifty-five articles deal with micro-sensors measuring mercury in water:

  • Forty-three articles suggest measuring mercury by an optical means: molecular absorption (Table 4); and fluorescence (Table 5) or by chemiluminescence (two sensors). Among the optical methods, let us note the indirect method of Dittman et al. [69] based on the correlation between 254 nm absorbance of hydrophobic acid fraction of dissolved organic matter and mercury concentration. This method was successfully applied to waters of three different forested watersheds in New Hampshire, Vermont, and the State of New York (USA). It gives determination coefficients equal to, respectively, 0.96, 0.99 and 0.98 and its detection limit would be around 0.5 ng/L, corresponding to 0.05 absorbance unit. The authors attribute this correlation to the high content of reduced sulphur sites in hydrophobic acid fraction of organic matter;

  • Fifteen articles propose an electrochemical method (Table 6).

Among the mixed optical methods, let us point out two techniques:
  • Lee and Mirkin’s method, associating DNA arrays using DNA-modified gold nanoparticle probes and a conventional flatbed scanner to measure scattered light, after silver amplification, gives a detection limit of 2 μg/L [70];

  • The virtual sensor of Wang et al. [65] combining together nine cross-reactive sensing fluorescent elements on micro-plate with pattern recognition software in order to discriminate and quantify Hg in water. The sensor arrays are excited with a broadband UV lamp and four filters are used for emission detection: 380–500 nm λmax= 435 nm; 480–600 nm λmax = 525 nm; 523 nm and 580 nm. The array capacity was tested at different ranges of pH and at different cation concentrations using linear discriminant analysis. Quantitative analysis can be achieved with 93% accuracy in the concentration ≥20 mg/L [65].

Two sensors are based on luminescence, the first one is based on bioluminescence on recombinant E. Coli immobilised on optical fibres [109]; and the second one is based on chemoluminescence of an iminophenol trimers [110]. Their detection limits are respectively 2.6 μg/L and 10 μg/L.

If we put aside the detection limit of Dittman et al.’s method [69] which would demand a validation on waters different to waters from forested watersheds (nitrate interference), the detection limits of the majority of these micro-sensors are still too high for an environmental application, but not too far from the “0.05 μg/L target”. Certain optical micro-sensors reach the detection limit of 0.2 μg/L by spectrometry [71,72,74] or fluorimetry [87,88], and electrochemical micro-sensors come down to the limit detection of 0.1 μg/L [38,111].

Only two micro-sensors reach the “0.05 μg/L target”, the first one with a detection limit 0.002 μg/L at pH 1 on diamond like carbon electrode [112] and the second one with a detection limit 0.03 μg/L but on acetate buffer at pH 4.2 on screen printed paste carbon electrode [113]. These micro-sensors are ready for the targeted or recommended values for water pollution control: WFD fixes a target value at 0.05 μg/L, but they are still no enough sensitive for reaching the SEQ-Eau target value: 0.007 μg/L [32]. Testing of these sensors for in situ mercury determination still remains to be done. None of these micro-sensors was evaluated in the field; only Dittman et al.’s method [69] should be applicable in the field with a commercial UV probe, of which spectral deconvolution algorithms should be fitted, and with a specific calibration.

3.1.2. Nickel

Only six articles (Table 7), among which two come from the same team, deal with nickel detection. Among these six papers, four use electrochemical methods: cyclic voltammetry [113,124], or conductometry [42,62], and two teams use optical methods – fluorimetry [65] or colorimetry [125].

Electrochemical methods give lower detection limits than optical methods. Only the electrochemical micro-sensors, once validated in laboratory, should be useful for nickel monitoring within the WFD framework (AA-EQS: 20 μg/L). Only the bismuth-film electrode micro-sensor [124] covers completely the INERIS or SEQ-Eau requirements, their target values being, respectively, 0.5 μg/L and 0.25 μg/L.

3.1.3. Lead

Thirty-six articles describe micro-sensors measuring lead in waters, the majority of them propose electrochemical methods (Table 8), only five articles present optical methods. The majority of the published electrochemical methods are voltammetric methods, mainly on modified carbon electrodes (paste or glassy carbon, graphite) with detection limits below the SEQ-Eau target value: 0.21 μg/L [34,40,126,127,128]. Turek et al., develop a virtual sensor founded on fuzzy logic data analysis of combined potentiometric signals from Ag-, Cu- and Pb-selective chalcogenide glass sensors (AgIAsSe, CuAgAsSe, PbI2Ag2SAs2S3). This method recognizes an unknown metal in solutions with 100% probability, for heavy metal concentrations higher than 210 μg/L [129].

Predicted Non Effect Concentration (PNEC) for lead, advocated by INERIS (PNEC: 5 μg/L), is only achieved using voltammetric sensors ([35,39,43,44,49,51,52,54,130132] and one conductometric biosensor based on Chlorella vulgaris [42]. However, this sensor shows a poor specificity. Indeed, heavy-metal ions acting as algal alkaline phosphatase inhibitors, so cadmium, cobalt, nickel and lead exhibit the same detection limit (1 μg/L).

Finally among published optical methods (Table 9), the most sensitive micro-sensor uses DNA probes and fluorescence quenching or turn-off. Pb(II) is measured by ADNzyme activation and DNA cleavage of two fluorescent cyanine probes: Cy3 (550/570 nm) for the detection and Cy5 (650/670 nm) for the reference [135]. Liu et al. [90] use a thrombin-binding aptamer probe labelled with donor carboxyfluorescein and quencher 4-([4-(dimethylamino)phenyl]azo)benzoic acid at its 5′ and 3′ termini, respectively. These two optical methods give acceptable detection limits for water monitoring, with respectively: 2 μg/L [135] and 0.1 μg/L [90]. Two other optical methods are colorimetric ones and present too high detection limits: 18 μg/L [136] and 1.3 mg/L [138].

All the aforementioned sensors are adapted to lead monitoring in water according to the WFD criterion: target value at 7 μg/L. However none of these micro-sensors is still at the evaluation stage and ruggedness and hardiness will have to guide selection of the micro-sensors fit for field validation.

3.1.4. Conclusions on metal determination

Mainly because of environmental regulations, the current research on sensors for metal determinations is under a kind of double bind: mercury elimination from devices and keeping the detection limits of mercury electrodes. This particularly stimulating situation will lead to new greener micro-sensors. These new generations of sensors are still in their infancy and many approaches are being explored. The electrochemical approach is the privileged technique by the four fifths of the authors, at least for Cd, Ni and Pb. For the mercury titration there is an inverse proportion with dominance in 80% of the optical approach. It is worthy noting that representing about 60% of published papers, the research efforts in the WFD metal determination seem at present to be to concentrate on mercury detection. The activity seems more particularly active in the research for new chromophores more specific and more sensitive, and of original couplings between chromophore and ionophore. Finally, some articles investigate the potential of virtual sensors, otherwise known as “soft sensors” or “smart sensors”, in heavy metal monitoring. These sensors consist of an array of simple and reliable sensors that are not analyte-specific but can be linked by a computer programmed to process certain sample features and build a proxy of the “unsensed” metal [65,129]. Multivariate calibration, a statistically inspired modification of partial least square method, is applied to DPASV voltammograms acquired on carbon screen-printed electrodes. This data processing leads to better overall root mean square error of prediction values for Cd and Pb. The authors propose to investigate the application of artificial neural networks to the voltammograms in order to reduce the errors and to improve the number of target metals [139]. Until now, their detection limits are too high, and research should be carried out in order to make them operational tools in the field.

4. Organic Substances

Among priority substances listed in Appendix II of the Common Decision taken on 20 December, 2007 by the Council of Europe [2], 36 are organic substances, the five remaining being metals (Cd, Ni, Pb and Hg) and an organotin. Organic substances can be grouped in main chemical or usage groups: 17 pesticides, 10 chlorinated solvents, eight PAH and miscellaneous organic substances as alkylphenols, detergents, brominated diphenyl ether, aryl-halides and benzene. Table 10 summarises the superior limit concentrations recommended or statutory in freshwater.

However the scientific publications concerning priority WFD organic substance determinations by means of micro-sensors represent a more modest proportion, than that dedicated to four WFD metals. During years 2007–2009 we counted only two dozens of papers, mainly on detection of organophosphates and of some carbamates (Table 11). Among these, twelve articles present electrochemical methods on modified electrodes, among which seven enzymatic biosensors based on choline esterase activity. Other micro-sensors are intended to measure atrazine [140142], isoproturon [143], paraquat [144] or DDT-like compounds [145]. Three articles propose sensors for solvent determination in water: two based on infrared spectroscopy and one on impedance spectrometry.

4.1. Organophosphates and Carbamates

Among organophosphate pesticides (Table 11) only chlorpyrifos appears among the thirty-three WFD substances, with an AA-EQS of 0.03 μg/L. However the proposed colorimetric method on gold nanoparticles presents high detection limits of several orders of magnitude for the titration of chlorpyrifos and malathion (20 μg/L and 100 μg/L, respectively). [146]. Acetylcholine esterase activity monitoring methods using voltammetry developed by Zhao et al., [147] and by Yin et al., [148] allow to reach lower concentrations, respectively: 0.5 ng/L for paraoxon and 1 ng/L for parathion [147], 6 ng/L for paraoxon methyl and 7 ng/L for carbofuran [148]. These methods would deserve to be tested on both WFD organophosphates: chlorpyrifos and chlorfenvinphos. Nevertheless these biosensors determine the total concentration of organophosphates and carbamate pesticides, and not a specific compound. Now, the definition of good chemical status in the WFD imposes the measurement of priority organic substance concentrations and not the measurement of toxic effects.

4.2. Other pesticides

Among other methods proposed for pesticide determination (Table 12), three methods are proposed for triazine measurement. Pardieu et al. [140] presented an electrochemical micro-sensor, based on a molecularly imprinted conducting polymer, deposited on a platinum electrode, able to detect, by cyclic voltammetry, triazines (atrazine, terbutylazine and simazine) and to measure atrazine with a detection limit of 4.5 μg/L. An immunoassay based on a commercially available surface plasmon resonance (SPR) biosensor was proposed. Measurement can be performed in 25 minutes, including the regeneration cycle, and its detection limit in laboratory is 20 ng/L and 26 ng/L in a river sample [141]. A microbioreactor and integrated sensors for the online measurement of oxygen production was used to measure benzalkonium chloride and atrazine effects on the macrophyte Elodea canadensis. No observable adverse effect concentration (NOAEC) for atrazine is 10 μg/L [142]. Gouzy et al. [143] developed an immunosensor based on SPR for the measurement of isoproturon. The rat monoclonal anti-isoproturon antibody was reversibly immobilized through the use of a capture mouse anti-rat monoclonal antibody, which was covalently immobilized on the sensor chip surface. The detection limit could have reach 0.1 μg/L, but the operational range goes from 1.3 μg/L to 16.3 μg/L (inhibition from 20 to 80%). Lee et al [144] measured morphology and resonance characteristics of human hepatoma cell line cultured onto an indium tin oxide layer on the surface of a quartz crystal modified with a collagen film, for the study of the effect of paraquat (2.7 μg/μL) through resonance frequency and resonance resistance responses. The system gives a qualitative response and allows the visualization of the action of paraquat on cells. Finally, an immunosensor based on commercially available SPR has been developed for the monitoring of DDT, its metabolites and analogues in real water samples. Low detection limits are attained for DDT-selective: 15 ng/L; and DDT group-selective immunoassays: 31 ng/L [145].

4.3. Solvents

The restricted number of solvent sensors developed (Table 13) present detection limits higher than the WFD and SEQ-Eau’s target values (Table 10), respectively 10 μg/L and 0.5 μg/L for benzene and 2.5 μg/L and 1.2 μg/L for chloroform.

Finally, Kurup [162] describes a virtual sensor constituted of an array of seven SnO2 micro-sensors doped with different impurities (as palladium or platinum) and equipped with a gas sampling membrane and integrated into a probe used for the volatile organic pollutant detection in underground environment. This technology supplies at the moment qualitative results (three levels: low, average and high), which must be then confirmed by conventional analyses in laboratory.

5. Conclusions

As noticed above, 81% of the publications appeared in 2007–2009 concern only 12% of the priority substances included in the WFD list, namely metals (e.g., Cd, Hg, Ni, Pb). The origin of this disparity already noticed in the production is doubtlessly attributable to the heterogeneousness of chemical structures listed in WFD list which exceeds the financial capacity for sensor development or imposes one weakly profitable development of specific micro-sensor for each priority substance. The fact that only the organophosphates and carbamates are subject of a relatively important number of publications among those dedicated to organic substances (57%), tends to confirm this origin. Indeed these sensors are based on acetylcholinesterase inhibition by organophosphate and carbamates. So, the conception of micro-sensors measuring biological effects rather than concentrations could constitute an innovative way allowing the release from constraint of the development of substance specific micro-sensors. However the notion of “Environmental Quality Standards” in the WFD sense imposes the measurement of specific concentrations and not some effects. This normative aspect of the Directive establishes a strong constraint in the research for new analytical devices.

None of articles published during 2007–2009 presents a micro-sensor totally validated in the laboratory and ready for tests under real field conditions. Research is mainly focused on the sensitive part of the micro-sensors: the receptor. Now, even if progress still remains to achieve, in particular to lower the limit of detection, the principles of detection do not seem to constitute the main limiting factor the development of environmental micro-sensors. The sensitivities of the best cadmium or mercury sensors exhibit detection limits respectively five and two folds higher than SQE target values. It is too high for in situ measurement in natural waters, but these sensors could be right now used for the industrial wastewater control (e.g., electroplating plants). During the wait for a better sensitivity, these mercury-free sensors could be associated to concentration devices in lab-on-a-chips, in order to lower the detection limits.

On the other hand, receptor ruggedness to environmental conditions must be improved, in particular concerning their fouling resistance; technologies such as surface passivation, or ultrasound cleaning of the sensitive surface, should allow progress in this domain. This point constitutes the first technological obstacle to be overcome, before going into long-term tests in real environments.

Other scientific and technological obstacles to overcome in order to build operational micro-sensor networks are at the transducer and the transmitter levels. Setup up a micro-sensor in the natural environment requires an infrastructure capable of fuelling it with energy, watching it and collecting data. This means that skills outside the fields of chemistry, such as physics, electronics and computing, which will be required in order to build a complete in situ measurement chain. The main axes of progress must be centred on the following points:

  • – Miniaturization: pursue the integration of the various modules (receptor, transducer and transmitter) on the same chip. It leads to a decrease of the energy consumption and seems to have, at least on the conductometric electrodes, a beneficial effect on the sensitivity. This miniaturization requires for skills in micro-technology;

  • – Communication: assure the communication between the micro-sensors and the readout station and between the micro-sensors. The network should have the possibility of positioning these sensors (search for micro-sensors scattered after a flood in particular), of verifying the state of its communications with the readout station and its neighbour micro-sensors, of validating the quality of its data before their transmission, of alerting in case of abnormal situation. Two wireless communication modes should be studying: ultrasound-based or radio frequencies;

  • – Autonomy: decrease the energy consumption or develop possibilities of in situ energy supply. Sedimentary microbial fuel cells or marine turbines would be technologies that should be investigated;

  • – Eco-design: integrate the possible impact of the chosen technology on the environment, as well as life cycle of the micro-sensor, and principles of measurement. Indeed in the eventuality of a loss of the material in the environment, the micro-sensor composition has to banish the use of well-known toxic or dangerous matter for the environment. This point concerns all devices introduced into the environment: we have to minimize as much as possible any toxic impact of our devices. Indeed, the WFD [1] demands, for priority hazardous substances from human activities (as cadmium or mercury), the cessation or phasing out of discharges, emissions and losses into water before 2020. This regulation coming into force in the Member States, the use of mercury is banished in the short-term. Also, development of new devices using priority hazardous substances leads to a commercial dead-end, and a waste of time and money: it cannot be used in Europe. Here and now, the regulatory requirements must be included in the design of new monitoring devices and we have to minimise the potential toxic impact of all components used to build efficient monitoring sensor network platform. Clearly, there is plenty of room for progression.

Finally it is advisable to mention the necessity of developing data processing technologies to manage in an optimal way the considerable data flows produced by a micro-sensor network. There are very significant challenges for the micro-sensor research community for delivering sensing platforms that are appropriate for integration into scaled-up deployments in terms of sustainability, cost, and reliability. Calls for proposals should centre on the conception of complete micro-sensor networks (material and software) and on the emergence of projects having as a global objective an operational monitoring network, from the receptors to the management computer.

Acknowledgments

Authors thank ONEMA (The French National Agency for Water and Aquatic Environment) for its financial support.

References

  1. JOL Directive 2000/60/CE. du Parlement européen et du Conseil du 23 octobre 2000 établissant un cadre pour une politique communautaire dans le domaine de l’eau. Directive 2000/60/CE; du Parlement européen et du Conseil: Luxembourg, Germany, 2000.
  2. JOL Directive 2008/32/CE. du Parlement européen et du Conseil du 11 mars 2008 modifiant la directive 2000/60/CE établissant un cadre pour une politique communautaire dans le domaine de l’eau; en ce qui concerne les compétences d’exécution conférées à la Commission. Directive 2008/32/CE; du Parlement européen et du Conseil: Luxembourg, Germany, 2008; pp. 60–61.
  3. Braungardt, CB; Achterberg, EP; Axelsson, B; Buffle, J; Graziottin, F; Howell, KA; Illuminati, S; Scarponi, G; Tappin, AD; Tercier-Waeber, ML; Turner, D. Analysis of dissolved metal fractions in coastal waters: An inter-comparison of five voltammetric in situ profiling (VIP) systems. Mar. Chem 2009, 114, 47–55. [Google Scholar]
  4. Greenwood, R; Mills, GA; Roig, B. Introduction to emerging tools and their use in water monitoring. TrAC Trends Anal. Chem 2007, 26, 263–267. [Google Scholar]
  5. JOL Décision nº 2455/2001/CE du Parlement européen et du Conseil du 20 novembre 2001 établissant la liste des substances prioritaires dans le domaine de l’eau et modifiant la directive 2000/60/CE. Décision nº 2455/2001/EC; du Parlement européen et du Conseil: Luxembourg, Germany, 2001; pp. 1–5.
  6. JOC Position Commune (CE) nº 3/2008 du 20 décembre 2007 arrêtée par le Conseil, statuant conformément à la procédure visée à l’article 251 du traité instituant la Communauté européenne, en vue de l’adoption d’une directive du Parlement européen et du Conseil établissant des normes de qualité environnementale dans le domaine de l’eau et modifiant les directives 82/176/CEE, 83/513/CEE, 84/156/CEE, 84/491/CEE, 86/280/CEE et 2000/60/CE. Position Commune nº 3/2008; du Parlement européen et du Conseil: Luxembourg, Germany, 2008; pp. 1–15.
  7. Johnson, KS; Needoba, JA; Riser, SC; Showers, WJ. Chemical sensor networks for the aquatic environment. Chem. Rev 2007, 107, 623–640. [Google Scholar]
  8. Pejcic, B; Eadington, P; Ross, A. Environmental monitoring of hydrocarbons: A chemical sensor perspective. Environ. Sci. Technol 2007, 41, 6333–6342. [Google Scholar]
  9. Bosch, ME; Sanchez, AJR; Rojas, FS; Ojeda, CB. Recent development in optical fiber biosensors. Sensors 2007, 7, 797–859. [Google Scholar]
  10. Lieberzeit, PA; Dickert, FL. Sensor technology and its application in environmental analysis. Anal. Bioanal. Chem 2007, 387, 237–247. [Google Scholar]
  11. Reimers, CE. Applications of microelectrodes to problems in chemical oceanography. Chem. Rev 2007, 107, 590–600. [Google Scholar]
  12. Tercier-Waeber, ML; Taillefert, M. Remote in situ voltammetric techniques to characterize the biogeochemical cycling of trace metals in aquatic systems. J. Environ. Monitor 2008, 10, 30–54. [Google Scholar]
  13. Tercier-Waeber, ML; Confalonieri, F; Koudelka-Hep, M; Dessureault-Rompre, J; Graziottin, F; Buffle, J. Gel-integrated voltammetric microsensors and submersible probes as reliable tools for environmental trace metal analysis and speciation. Electroanal 2008, 20, 240–258. [Google Scholar]
  14. Lange, K; Rapp, BE; Rapp, M. Surface acoustic wave biosensors: A review. Anal. Bioanal. Chem 2008, 391, 1509–1519. [Google Scholar]
  15. Moreno-Garrido, I. Microalgae immobilization: Current techniques and uses. Bioresource Technol 2008, 99, 3949–3964. [Google Scholar]
  16. Jones, SEW; Compton, RG. Stripping analysis using boron-doped diamond electrodes. Curr. Anal. Chem 2008, 4, 170–176. [Google Scholar]
  17. Jaffrezic-Renault, N; Dzyadevych, SV. Conductometric microbiosensors for environmental monitoring. Sensors 2008, 8, 2569–2588. [Google Scholar]
  18. Seidel, M; Niessner, R. Automated analytical microarrays: A critical review. Anal. Bioanal. Chem 2008, 391, 1521–1544. [Google Scholar]
  19. Palchetti, I; Mascini, M. Nucleic acid biosensors for environmental pollution monitoring. Analyst 2008, 133, 846–854. [Google Scholar]
  20. Johnson-White, B; Harmon, HJ. Optical Enzyme-Based Sensors for Reagentless Detection of Chemical Analytes. In Antiterrorism & Homeland Defense: Polymers & Materials; Reynolds, JG, Lawson, GE, Koester, CJ, Eds.; American Chemical Society: Washington, DC, USA, 2007; pp. 57–70. [Google Scholar]
  21. Nolan, EM; Lippard, SJ. Tools and tactics for the optical detection of mercuric ion. Chem. Rev 2008, 108, 3443–3480. [Google Scholar]
  22. Diamond, D; Coyle, S; Scarmagnani, S; Hayes, J. Wireless sensor networks and chemo-/biosensing. Chem. Rev 2008, 108, 652–679. [Google Scholar]
  23. Strobl, RO; Robillard, PD. Network design for water quality monitoring of surface freshwaters: A review. J. Environ. Manage 2008, 87, 639–648. [Google Scholar]
  24. Bogue, R. Environmental sensing: Strategies, technologies and applications. Sensor Rev 2008, 28, 275–282. [Google Scholar]
  25. Jaffrezic-Renault, N. Analytical electrochemistry for environment: electrochemical microsensors for the in situ monitoring of pollutants. Actual Chimique 2009, 52–55. [Google Scholar]
  26. Baruah, S; Dutta, J. Nanotechnology applications in pollution sensing and degradation in agriculture: A review. Environ. Chem. Lett 2009, 7, 191–204. [Google Scholar]
  27. Pesavento, M; Alberti, G; Biesuz, R. Analytical methods for determination of free metal ion concentration, labile species fraction and metal complexation capacity of environmental waters: A review. Anal. Chim. Acta 2009, 631, 129–141. [Google Scholar]
  28. Reardon, KF; Zhong, Z; Lear, KL. Environmental Applications of Photoluminescence-Based Biosensors. In Optical Sensor Systems in Biotechnology; Springer-Verlag: Berlin, Germany, 2009; Volume 116, pp. 99–123. [Google Scholar]
  29. Leray, I; Valeur, B. Calixarene-based fluorescent molecular sensors for toxic metals. Eur J Inorg Chem 2009, 3525–3535. [Google Scholar]
  30. Selid, PD; Xu, HY; Collins, EM; Face-Collins, MS; Zhao, JX. Sensing mercury for biomedical and environmental monitoring. Sensors 2009, 9, 5446–5459. [Google Scholar]
  31. David, A; Fenet, H; Gomez, E. Alkylphenols in marine environments: Distribution monitoring strategies and detection considerations. Mar. Pollut. Bull 2009, 58, 953–960. [Google Scholar]
  32. RMC. SEQ-Eau. système d’évaluation de la qualité de l’eau des cours d’eau (Version 2) 2003, 2003;. 40.
  33. 20381–20399.
  34. Deng, WF; Tan, YM; Fang, ZF; Xie, QJ; Li, YY; Liang, XL; Yao, SZ. ABTS-Multiwalled carbon nanotubes nanocomposite/Bi film electrode for sensitive determination of Cd and Pb by differential pulse stripping voltammetry. Electroanal 2009, 21, 2477–2485. [Google Scholar]
  35. El Tall, O; Jaffrezic-Renault, N; Sigaud, M; Vittori, O. Anodic stripping voltammetry of heavy metals at nanocrystalline boron-doped diamond electrode. Electroanal 2007, 19, 1152–1159. [Google Scholar]
  36. Rehacek, V; Hotovy, I; Vojs, M; Mika, F. Bismuth film electrodes for heavy metals determination. Microsyst. Technol 2008, 14, 491–498. [Google Scholar]
  37. El Mhammedi, MA; Achak, M; Najih, R; Bakasse, M; Chtaini, A. Micro-extraction and trace determination of cadmium by square wave voltammetry at the carbon paste electrode impregnated with Ca10(PO4)6(OH)2. Mater. Chem. Phys 2009, 115, 567–571. [Google Scholar]
  38. Pal, P; Bhattacharyay, D; Mukhopadhyay, A; Sarkar, P. The detection of mercury, cadmium, and arsenic by the deactivation of urease on rhodinized carbon. Environ. Eng. Sci 2009, 26, 25–32. [Google Scholar]
  39. Wu, Y; Li, NB; Luo, HQ. Simultaneous measurement of Pb, Cd and Zn using differential pulse anodic stripping voltammetry at a bismuth/poly(p-aminobenzene sulfonic acid) film electrode. Sensor. Actuat. B-Chem 2008, 133, 677–681. [Google Scholar]
  40. Tesarova, E; Baldrianova, L; Hocevar, SB; Svancara, I; Vytras, K; Ogorevc, B. Anodic stripping voltammetric measurement of trace heavy metals at antimony film carbon paste electrode. Electrochim. Acta 2009, 54, 1506–1510. [Google Scholar]
  41. Tian, YQ; Li, NB; Luo, HQ. Simultaneous determination of trace zinc (ii) and cadmium (ii) by differential pulse anodic stripping voltammetry using a MWCNTs-NaDBS modified stannum film electrode. Electroanal 2009, 21, 2584–2589. [Google Scholar]
  42. Berezhetskyy, AL; Durrieu, C; Nguyen-Ngoc, H; Chovelon, JM; Dzyadevich, S; Tran-Minh, C. Conductometric biosensor based on whole-cell microalgae for assessment of heavy metal in wastewater. Biopolym. Cell 2007, 23, 511–518. [Google Scholar]
  43. Kokkinos, C; Economou, A; Raptis, I; Efstathiou, CE. Lithographically fabricated disposable bismuth-film electrodes for the trace determination of Pb(II) and Cd(II) by anodic stripping voltammetry. Electrochim. Acta 2008, 53, 5294–5299. [Google Scholar]
  44. Senthilkumar, S; Saraswathi, R. Electrochemical sensing of cadmium and lead ions at zeolite-modified electrodes: Optimization and field measurements. Sensor. Actuat. B-Chem 2009, 141, 65–75. [Google Scholar]
  45. Vedhi, C; Selvanathan, G; Arumugam, P; Manisankar, P. Electrochemical sensors of heavy metals using novel polymer-modified glassy carbon electrodes. Ionics 2009, 15, 377–383. [Google Scholar]
  46. Guedri, H; Durrieu, C. A Nafion/BSA based conductometric algal whole cell biosensor for heavy metal monitoring. Environ Sci Technol 2009, II, 623–628. [Google Scholar]
  47. Guedri, H; Durrieu, C. A self-assembled monolayers based conductometric algal whole cell biosensor for water monitoring. Microchim. Acta 2008, 163, 179–184. [Google Scholar]
  48. Cugnet, C; Zaouak, O; Rene, A; Pecheyran, C; Potin-Gautier, M; Authier, L. A novel microelectrode array combining screen-printing and femtosecond laser ablation technologies: Development, characterization and application to cadmium detection. Sensor. Actuat. B-Chem 2009, 143, 158–163. [Google Scholar]
  49. Zheng, H; Yan, Z; Dong, H; Ye, B. Simultaneous determination of lead and cadmium at a glassy carbon electrode modified with Langmuir-Blodgett film of p-tert-butylthiacalix[4]arene. Sensor. Actuat. B-Chem 2007, 120, 603–609. [Google Scholar]
  50. Toghill, KE; Wildgoose, GG; Moshar, A; Mulcahy, C; Compton, RG. The fabrication and characterization of a bismuth nanoparticle modified boron doped diamond electrode and its application to the simultaneous determination of cadmium(II) and lead(II). Electroanal 2008, 20, 1731–1737. [Google Scholar]
  51. Yantasee, W; Charnhattakorn, B; Fryxell, GE; Lin, Y; Timchalk, C; Addleman, RS. Detection of Cd, Pb, and Cu in non-pretreated natural waters and urine with thiol functionalized mesoporous silica and Nafion composite electrodes. Anal. Chim. Acta 2008, 620, 55–63. [Google Scholar]
  52. Guell, R; Aragay, G; Fontas, C; Antico, E; Merkoci, A. Sensitive and stable monitoring of lead and cadmium in seawater using screen-printed electrode and electrochemical stripping analysis. Anal. Chim. Acta 2008, 627, 219–224. [Google Scholar]
  53. Bagal-Kestwal, D; Karve, MS; Kakade, B; Pillai, VK. Invertase inhibition based electrochemical sensor for the detection of heavy metal ions in aqueous system: Application of ultra-microlelectrode to enhance sucrose biosensor’s sensitivity. Biosens. Bioelectron 2008, 24, 657–664. [Google Scholar]
  54. Rico, MAG; Olivares-Marin, M; Gil, EP. Modification of carbon screen-printed electrodes by adsorption of chemically synthesized Bi nanoparticles for the voltammetric stripping detection of Zn(II), Cd(II) and Pb(II). Talanta 2009, 80, 631–635. [Google Scholar]
  55. Cooper, J; Bolbot, JA; Saini, S; Setford, SJ. Electrochemical method for the rapid on site screening of cadmium and lead in soil and water samples. Water Air Soil Poll 2007, 179, 183–195. [Google Scholar]
  56. Zou, ZW; Jang, A; MacKnight, E; Wu, PM; Do, J; Bishop, PL; Ahn, CH. Environmentally friendly disposable sensors with microfabricated on-chip planar bismuth electrode for in situ heavy metal ions measurement. Sensor. Actuat. B-Chem 2008, 134, 18–24. [Google Scholar]
  57. Kadara, RO; Tothill, IE. Development of disposable bulk-modified screen-printed electrode based on bismuth oxide for stripping chronopotentiometric analysis of lead(II) and cadmium(II) in soil and water samples. Anal. Chim. Acta 2008, 623, 76–81. [Google Scholar]
  58. Lee, HJ; Lagger, G; Pereira, CM; Silva, AF; Girault, HH. Amperometric tape ion sensors for cadmium(II) ion analysis. Talanta 2009, 78, 66–70. [Google Scholar]
  59. Toghill, KE; Xiao, L; Wildgoose, GG; Compton, RG. Electroanalytical Determination of Cadmium(II) and Lead(II) Using an Antimony Nanoparticle Modified Boron-Doped Diamond Electrode. Electroanal 2009, 21, 1113–1118. [Google Scholar]
  60. Ensafi, AA; Meghdadi, S; Sedighi, S. Sensitive cadmium potentiometric sensor based on 4-hydroxy salophen as a fast tool for water samples analysis. Desalination 2009, 242, 336–345. [Google Scholar]
  61. Gupta, VK; Jain, AK; Ludwig, R; Maheshwari, G. Electroanalytical studies on cadmium(II) selective potentiometric sensors based on t-butyl thiacalix[4]arene and thiacalix[4]arene in poly(vinyl chloride). Electrochim. Acta 2008, 53, 2362–2368. [Google Scholar]
  62. Berezhetskyy, AL; Sosovska, OF; Durrieu, C; Chovelon, JM; Dzyadevych, SV; Tran-Minh, C. Alkaline phosphatase conductometric biosensor for heavy-metal ions determination. Innov. Tech. Biol. Med. RBM 2008, 29, 136–140. [Google Scholar]
  63. Ensafi, AA; Meghdadi, S; Fooladgar, E. Development of a new selective optical sensor for Cd(II) Ions based on 4-Hydroxy Salophen. IEEE Sens. J 2008, 8, 1794–1800. [Google Scholar]
  64. Luo, HY; Jiang, JH; Zhang, XB; Li, CY; Shen, GL; Yu, RQ. Synthesis of porphyrin-appended terpyridine as a chemosensor for cadmium based on fluorescent enhancement. Talanta 2007, 72, 575–581. [Google Scholar]
  65. Wang, Z; Palacios, MA; Anzenbacher, P. Fluorescence sensor array for metal ion detection based on various coordination chemistries: General performance and potential application. Anal. Chem 2008, 80, 7451–7459. [Google Scholar]
  66. Banerjee, S; Kara, S; Santra, S. A simple strategy for quantum dot assisted selective detection of cadmium ions. Chem Commun 2008, 3037–3039. [Google Scholar]
  67. Tang, XL; Peng, XH; Dou, W; Mao, J; Zheng, JR; Qin, WW; Liu, WS; Chang, J; Yao, XJ. Design of a semirigid molecule as a selective fluorescent chemosensor for recognition of Cd(II). Org. Lett 2008, 10, 3653–3656. [Google Scholar]
  68. INERIS. Cadmium et ses dérivés; INERIS: 22/02/2005,; 2005; p. 60. [Google Scholar]
  69. Dittman, JA; Shanley, JB; Driscoll, CT; Aiken, GR; Chalmers, AT; Towse, JE. Ultraviolet absorbance as a proxy for total dissolved mercury in streams. Environ. Pollut 2009, 157, 1953–1956. [Google Scholar]
  70. Lee, JS; Mirkin, CA. Chip-based scanometric detection of mercuric ion using DNA-functionalized gold nanoparticles. Anal. Chem 2008, 80, 6805–6808. [Google Scholar]
  71. Amini, MK; Khezri, B; Firooz, AR. Development of a highly sensitive and selective optical chemical sensor for batch and flow-through determination of mercury ion. Sensor. Actuat. B-Chem 2008, 131, 470–478. [Google Scholar]
  72. Yari, A; Papi, F. Highly selective sensing of mercury(II) by development and characterization of a PVC-based optical sensor. Sensor. Actuat. B-Chem 2009, 138, 467–473. [Google Scholar]
  73. Khezri, B; Amini, MK; Firooz, AR. An optical chemical sensor for mercury ion based on 2-mercaptopyrimidine in PVC membrane. Anal Bioanal Chem 2008, 1–8. [Google Scholar]
  74. Li, JS; Yao, JJ; Zhong, WW. Membrane blotting for rapid detection of mercury(II) in water. Chem. Commun 2009, 33, 4962–4964. [Google Scholar]
  75. Kalyan, Y; Pandey, AK; Bhagat, PR; Acharya, R; Natarajan, V; Naidu, GRK; Reddy, AVR. Membrane optode for mercury(II) determination in aqueous samples. J. Hazard. Mater 2009, 166, 377–382. [Google Scholar]
  76. El-Safty, SA. Organic-inorganic hybrid mesoporous monoliths for selective discrimination and sensitive removal of toxic mercury ions. J. Mater. Sci 2009, 44, 6764–6774. [Google Scholar]
  77. Shunmugam, R; Gabriel, GJ; Smith, CE; Aamer, KA; Tew, GN. A highly selective colorimetric aqueous sensor for mercury. Chem.Eur. J 2008, 14, 3904–3907. [Google Scholar]
  78. Shao, N; Gao, X; Wang, H; Yang, RH; Chan, WH. Spiropyran-based optical approaches for mercury ion sensing: Improving sensitivity and selectivity via cooperative ligation interactions using cysteine. Anal. Chim. Acta 2009, 655, 1–7. [Google Scholar]
  79. Ensafi, AA; Katiraei Far, A; Meghdadi, S. Highly selective optical sensor for mercury assay based on covalent immobilization of 4-hydroxy salophen on a triacetylcellulose membrane. Sensor. Actuat. B-Chem 2008, 133, 84–90. [Google Scholar]
  80. Ensafi, AA; Fouladgar, M. A sensitive and selective bulk optode for determination of Hg(II) based on hexathiacyclooctadecane and chromoionophore V. Sensor. Actuat. B-Chem 2009, 136, 326–331. [Google Scholar]
  81. Ros-Lis, JV; Casasús, R; Comes, M; Coll, C; Marcos, MD; Martínez-Máñez, R; Sancenón, F; Soto, J; Amorós, P; El Haskouri, J; Garró, N; Rurack, K. A Mesoporous 3D Hybrid Material with Dual Functionality for Hg2+ Detection and Adsorption. Chem.Eur. J 2008, 14, 8267–8278. [Google Scholar]
  82. Ensafi, AA; Fouladgar, M. Development of a spectrophotometric optode for the determination of Hg(II). IEEE Sens. J 2008, 8, 347–353. [Google Scholar]
  83. Leng, B; Zou, L; Jiang, J; Tian, H. Colorimetric detection of mercuric ion (Hg2+) in aqueous media using chemodosimeter-functionalized gold nanoparticles. Sensor. Actuat. B-Chem 2009, 140, 162–169. [Google Scholar]
  84. Nuriman; Kuswandi, B; Verboom, W. Selective chemosensor for Hg(II) ions based on tris[2-(4-phenyldiazenyl)phenylaminoethoxy]cyclotriveratrylene in aqueous samples. Anal. Chim. Acta 2009, 655, 75–79. [Google Scholar]
  85. Pérez-Hernández, J; Albero, J; Llobet, E; Correig, X; Matías, IR; Arregui, FJ; Palomares, E. Mercury optical fibre probe based on a modified cladding of sensitised Al2O3 nano-particles. Sensor. Actuat. B-Chem 2009, 143, 103–110. [Google Scholar]
  86. Lee, SJ; Lee, JE; Seo, J; Jeong, IY; Lee, SS; Jung, JH. Optical sensor based on nanomaterial for the selective detection of toxic metal ions. Adv. Funct. Mater 2007, 17, 3441–3446. [Google Scholar]
  87. Zheng, AF; Chen, JL; Wu, GN; Wei, HP; He, CY; Kai, XM; Wu, GH; Chen, YC. Optimization of a sensitive method for the “switch-on” determination of mercury(II) in waters using Rhodamine B capped gold nanoparticles as a fluorescence sensor. Microchim. Acta 2009, 164, 17–27. [Google Scholar]
  88. Ye, BC; Yin, BC. Highly Sensitive Detection of Mercury(II) Ions by Fluorescence Polarization Enhanced by Gold Nanoparticles. Angew. Chem. Int. Edit 2008, 47, 8386–8389. [Google Scholar]
  89. Huang, W; Zhou, P; Yan, WB; He, C; Xiong, LQ; Li, FY; Duan, CY. A bright water-compatible sugar-rhodamine fluorescence sensor for selective detection of Hg2+ in natural water and living cells. J. Environ. Monitor 2009, 11, 330–335. [Google Scholar]
  90. Liu, CW; Huang, CC; Chang, HT. Highly Selective DNA-Based Sensor for Lead(II) and Mercury(II) Ions. Anal. Chem 2009, 81, 2383–2387. [Google Scholar]
  91. He, CS; Zhu, WP; Xu, YF; Chen, T; Qian, XH. Trace mercury(II) detection and separation in serum and water samples using a reusable bifunctional fluorescent sensor. Anal. Chim. Acta 2009, 651, 227–233. [Google Scholar]
  92. Yu, CJ; Cheng, TL; Tseng, WL. Effects of Mn2+ on oligonucleotide-gold nanoparticle hybrids for colorimetric sensing of Hg2+: Improving colorimetric sensitivity and accelerating color change. Biosens. Bioelectron 2009, 25, 204–210. [Google Scholar]
  93. Guo, LQ; Hu, H; Sun, RQ; Chen, GA. Highly sensitive fluorescent sensor for mercury ion based on photoinduced charge transfer between fluorophore and pi-stacked T-Hg(II)-T base pairs. Talanta 2009, 79, 775–779. [Google Scholar]
  94. Han, ZX; Luo, HY; Zhang, XB; Kong, RM; Shen, GL; Yu, RQ. A ratiometric chemosensor for fluorescent determination of Hg2+ based on a new porphyrin-quinoline dyad. Spectrochim. Acta A 2009, 72, 1084–1088. [Google Scholar]
  95. Ren, X; Xu, QH. Highly sensitive and selective detection of mercury ions by using oligonucleotides, DNA intercalators, and conjugated polymers. Langmuir 2009, 25, 29–31. [Google Scholar]
  96. Long, F; He, M; Shi, HC; Zhu, AN. Development of evanescent wave all-fiber immunosensor for environmental water analysis. Biosens. Bioelectron 2008, 23, 952–958. [Google Scholar]
  97. Yang, Y; Jiang, J; Shen, G; Yu, R. An optical sensor for mercury ion based on the fluorescence quenching of tetra(p-dimethylaminophenyl)porphyrin. Anal. Chim. Acta 2009, 636, 83–88. [Google Scholar]
  98. Liu, X; Tang, Y; Wang, L; Zhang, J; Song, S; Fan, C; Wang, S. Optical detection of mercury(II) in aqueous solutions by using conjugated polymers and label-free oligonucleotides. Adv. Mater 2007, 19, 1471–1474. [Google Scholar]
  99. Liu, CW; Huang, CC; Chang, HT. Control over surface DNA density on gold nanoparticles allows selective and sensitive detection of mercury(II). Langmuir 2008, 24, 8346–8350. [Google Scholar]
  100. Nolan, EM; Lippard, SJ. Turn-on and ratiometric mercury sensing in water with a red-emitting probe. J. Am. Chem. Soc 2007, 129, 5910–5918. [Google Scholar]
  101. Lee, DN; Kim, GJ; Kim, HJ. A Fluorescent coumarinylalkyne probe for the selective detection of mercury(II) ion in water. Tetrahedron Lett 2009, 50, 4766–4768. [Google Scholar]
  102. Wanichacheva, N; Siriprumpoonthum, M; Kamkaew, A; Grudpan, K. Dual optical detection of a novel selective mercury sensor based on 7-nitrobenzo-2-oxa-1,3-diazolyl subunits. Tetrahedron Lett 2009, 50, 1783–1786. [Google Scholar]
  103. Li, CY; Zhang, XB; Qiao, L; Zhao, Y; He, CM; Huan, SY; Lu, LM; Jian, LX; Shen, GL; Yu, RQ. Naphthalimide-Porphyrin hybrid based ratiometric bioimaging probe for Hg2+: well-resolved emission spectra and unique specificity. Anal. Chem 2009, 81, 9993–10001. [Google Scholar]
  104. Li, H; Yan, H. Ratiometric fluorescent mercuric sensor based on thiourea-thiadiazole- pyridine linked organic nanoparticles. J. Phys. Chem. C 2009, 113, 7526–7530. [Google Scholar]
  105. Shang, ZB; Wang, Y; Jin, WJ. Triethanolamine-capped CdSe quantum dots as fluorescent sensors for reciprocal recognition of mercury (II) and iodide in aqueous solution. Talanta 2009, 78, 364–369. [Google Scholar]
  106. Liu, Y; Yu, M; Chen, Y; Zhang, N. Convenient and highly effective fluorescence sensing for Hg2+ in aqueous solution and thin film. Bioorg. Med. Chem 2009, 17, 3887–3891. [Google Scholar]
  107. Wang, J; Huang, L; Xue, M; Liu, L; Wang, Y; Gao, L; Zhu, J; Zou, Z. Developing a novel fluorescence chemosensor by self-assembly of Bis-Schiff base within the channel of mesoporous SBA-15 for sensitive detecting of Hg2+ ions. Appl. Surf. Sci 2008, 254, 5329–5335. [Google Scholar]
  108. Vaswani, KG; Keranen, MD. Detection of aqueous mercuric ion with a structurally simple 8-hydroxyquinoline derived ON-OFF fluorosensor. Inorg. Chem 2009, 48, 5797–5800. [Google Scholar]
  109. Ivask, A; Green, T; Polyak, B; Mor, A; Kahru, A; Virta, M; Marks, R. Fibre-optic bacterial biosensors and their application for the analysis of bioavailable Hg and As in soils and sediments from Aznalcollar mining area in Spain. Biosens. Bioelectron 2007, 22, 1396–1402. [Google Scholar]
  110. Kanan, SM; Abu-Yousef, IA; Hassouneh, N; Malkawi, A; Abdo, N; Kanan, MC. A highly selective luminescent sensor for detecting mercuric ions in water. Aust. J. Chem 2009, 62, 1593–1599. [Google Scholar]
  111. Kong, RM; Zhang, XB; Zhang, LL; Jin, XY; Huan, SY; Shen, GL; Yu, RQ. An ultrasensitive electrochemical “turn-on” label-free biosensor for Hg2+ with AuNP-functionalized reporter DNA as a signal amplifier. Chem Commun 2009, 5633–5635. [Google Scholar]
  112. Seehra, MS; Ranganathan, S; Manivannan, A. Electrochemical quantification of mercury in solutions using boron-doped diamond electrodes: Electrode regeneration and role of gold and impurities. Anal. Lett 2008, 41, 2162–2170. [Google Scholar]
  113. Khadro, B; Vittori, O; Jaffrezic-Renault, N. A new differential pulse anodic stripping voltammetry miniaturized sensor system for detection of nickel and mercury in waters. Sens. Lett 2009, 7, 829–832. [Google Scholar]
  114. Kim, TH; Lee, J; Hong, S. Highly selective environmental nanosensors based on anomalous response of carbon nanotube conductance to mercury ions. J. Phys Chem-C 2009, 113, 19393–19396. [Google Scholar]
  115. Gupta, VK; Singh, AK; Al Khayat, A; Gupta, B. Neutral carriers based polymeric membrane electrodes for selective determination of mercury(II). Anal. Chim. Acta 2007, 590, 81–90. [Google Scholar]
  116. Şar, E; Berber, H; Aşçı, B; Cankitrtaran, H. Determination of some heavy metal ions with a carbon paste electrode modified by poly(glycidylmethacrylate-methylmethacrylate-divinylbenzene) microspheres functionalized by 2-aminothiazole. Electroanal 2008, 20, 1533–1541. [Google Scholar]
  117. Javanbakht, M; Divsar, F; Badiei, A; Ganjali, MR; Norouzi, P; Ziarani, GM; Chaloosi, M; Jahangir, AA. Potentiometric detection of Mercury(II) ions using a carbon paste electrode modified with substituted thiourea-functionalized highly ordered nanoporous Silica. Anal. Sci 2009, 25, 789–794. [Google Scholar]
  118. Han, D; Kim, YR; Oh, JW; Kim, TH; Mahajan, RK; Kim, JS; Kim, H. A regenerative electrochemical sensor based on oligonucleotide for the selective determination of mercury(II). Analyst 2009, 134, 1857–1862. [Google Scholar]
  119. Ensafi, AA; Meghdadi, S; Allafchian, AR. Highly selective potentiometric membrane sensor for Hg(II) based on bis(benzoyl acetone) diethylene triamine. IEEE Sens. J 2008, 8, 248–254. [Google Scholar]
  120. Abu-Shawish, HM. A mercury(II) selective sensor based on N,N′-bis(salicylaldehyde)-phenylenediamine as neutral carrier for potentiometric analysis in water samples. J. Hazard. Mater 2009, 167, 602–608. [Google Scholar]
  121. Khun, NW; Liu, E. Linear sweep anodic stripping voltammetry of heavy metals from nitrogen doped tetrahedral amorphous carbon thin films. Electrochim. Acta 2009, 54, 2890–2898. [Google Scholar]
  122. Marcolino-Junior, LH; Janegitz, BC; Lourenção, BC; Fatibello-Filho, O. Anodic stripping voltammetric determination of mercury in water using a chitosan-modified carbon paste electrode. Anal. Lett 2007, 40, 3119–3128. [Google Scholar]
  123. Guascito, MR; Malitesta, C; Mazzotta, E; Turco, A. Inhibitive determination of metal ions by an amperometric glucose oxidase biosensor—Study of the effect of hydrogen peroxide decomposition. Sensor. Actuat. B-Chem 2008, 131, 394–402. [Google Scholar]
  124. Kokkinos, C; Economou, A; Raptis, I; Speliotis, T. Disposable mercury-free cell-on-a-chip devices with integrated microfabricated electrodes for the determination of trace nickel(II) by adsorptive stripping voltammetry. Anal. Chim. Acta 2008, 622, 111–118. [Google Scholar]
  125. Li, HB; Cui, ZM; Han, CP. Glutathione-stabilized silver nanoparticles as colorimetric sensor for Ni2+ ion. Sensor. Actuat. B-Chem 2009, 143, 87–92. [Google Scholar]
  126. El Mhammedi, MA; Achak, M; Chtaini, A. Ca10(PO4)6(OH)2 modified carbon-paste electrode for the determination of trace lead(II) by square-wave voltammetry. J. Hazard. Mater 2009, 161, 55–61. [Google Scholar]
  127. Nasraoui, R; Floner, D; Geneste, F. Analytical performances of a flow electrochemical sensor for preconcentration and stripping voltammetry of metal ions. J. Electroanal. Chem 2009, 629, 30–34. [Google Scholar]
  128. Faridbod, F; Ganjali, MR; Larijani, B; Hosseini, M; Alizadeh, K; Norouzi, P. Highly selective and sensitive asymmetric lead microsensor based on 5,5,dithiobis(2-nitrobenzoic acid) as an excellent hydrophobic neutral carrier for nano level monitoring of lead in real samples. Int. J. Electrochem. Sc 2009, 4, 1528–1540. [Google Scholar]
  129. Turek, M; Heiden, W; Riesen, A; Chhabda, TA; Schubert, J; Zander, W; Krüger, P; Keusgen, M; Schöning, MJ. Artificial intelligence/fuzzy logic method for analysis of combined signals from heavy metal chemical sensors. Electrochim. Acta 2009, 54, 6082–6088. [Google Scholar]
  130. Hassan, RYA; Habib, IHI; Hassan, HNA. Voltammetric determination of lead (II) in medical lotion and biological samples using chitosan-carbon paste electrode. Int. J. Electrochem. Sc 2008, 3, 935–945. [Google Scholar]
  131. Vedhi, C; Selvanathan, G; Arumugam, P; Manisankar, P. Electrochemical sensors of heavy metals using novel polymer-modified glassy carbon electrodes. Ionics 2008, 1–7. [Google Scholar]
  132. Yantasee, W; Hongsirikarn, K; Warner, CL; Choi, D; Sangvanich, T; Toloczko, MB; Warner, MG; Fryxell, GE; Addleman, RS; Timchalk, C. Direct detection of Pb in urine and Cd, Pb, Cu, and Ag in natural waters using electrochemical sensors immobilized with DMSA functionalized magnetic nanoparticles. Analyst 2008, 133, 348–355. [Google Scholar]
  133. Li, XG; Ma, XL; Huang, MR. Lead(II) ion-selective electrode based on polyaminoanthraquinone particles with intrinsic conductivity. Talanta 2009, 78, 498–505. [Google Scholar]
  134. Kazemi, SY; Shamsipur, M; Sharghi, H. Lead-selective poly(vinyl chloride) electrodes based on some synthesized benzo-substituted macrocyclic diamides. J. Hazard. Mater 2009, 172, 68–73. [Google Scholar]
  135. Zuo, P; Yin, BC; Ye, BC. DNAzyme-based microarray for highly sensitive determination of metal ions. Biosens. Bioelectron 2009, 25, 935–939. [Google Scholar]
  136. Ensafi, AA; Far, AK; Meghdadi, S. Highly selective optical-sensing film for lead(II) determination in water samples. J. Hazard. Mater 2009, 172, 1069–1075. [Google Scholar]
  137. Yu, M; He, F; Tang, Y; Wang, S; Li, Y; Zhu, D. Non-ionic water-soluble crown-ether-substituted polyfluorene as fluorescent probe for lead ion assays. Macromol. Rapid Comm 2007, 28, 1333–1338. [Google Scholar]
  138. Ensafi, AA; Fouladgar, M. Development a simple PVC membrane bulk optode for determination of lead ions in water samples. Sens. Lett 2009, 7, 177–184. [Google Scholar]
  139. Cauchi, M; Bessant, C; Setford, S. Simultaneous Quantitative Determination of cadmium, lead, and copper on carbon-ink screen-printed electrodes by differential pulse anodic stripping voltammetry and partial least squares regression. Electroanal 2008, 20, 2571–2577. [Google Scholar]
  140. Pardieu, E; Cheap, H; Vedrine, C; Lazerges, M; Lattach, Y; Garnier, F; Ramita, S; Pernelle, C. Molecularly imprinted conducting polymer based electrochemical sensor for detection of atrazine. Anal. Chim. Acta 2009, 649, 236–245. [Google Scholar]
  141. Farré, M; Martinez, E; Ramon, J; Navarro, A; Radjenovic, J; Mauriz, E; Lechuga, L; Marco, MP; Barcelo, D. Part per trillion determination of atrazine in natural water samples by a surface plasmon resonance immunosensor. Anal. Bioanal. Chem 2007, 388, 207–214. [Google Scholar]
  142. Vervllet-Scheebaurn, M; Ritzenthaler, R; Normann, J; Wagner, E. Short-term effects of benzalkonium chloride and atrazine on Elodea canadensis using a miniaturised microbioreactor system for an online monitoring of physiologic parameters. Ecotox. Environ. Safe 2008, 69, 254–262. [Google Scholar]
  143. Gouzy, MF; Keß, M; Krämer, PM. A SPR-based immunosensor for the detection of isoproturon. Biosens. Bioelectron 2009, 24, 1563–1568. [Google Scholar]
  144. Lee, DY; Kang, HW; Kaneko, S; Kwon, YS; Muramatsu, H. Direct monitoring of paraquat induced cell death using quartz crystal sensor. Thin Solid Films 2009, 518, 707–710. [Google Scholar]
  145. Mauriz, E; Calle, A; Manclus, JJ; Montoya, A; Hildebrandt, A; Barcelo, D; Lechuga, LM. Optical immunosensor for fast and sensitive detection of DDT and related compounds in river water samples. Biosens. Bioelectron 2007, 22, 1410–1418. [Google Scholar]
  146. Lisha, KP; Anshup; Pradeep, T. Enhanced visual detection of pesticides using gold nanoparticles. J. Environ. Sci. Heal. B 2009, 44, 697–705. [Google Scholar]
  147. Zhao, W; Ge, PY; Xu, JJ; Chen, HY. Selective detection of hypertoxic organophosphates pesticides via PDMS composite based acetylcholinesterase-inhibition biosensor. Environ. Sci. Technol 2009, 43, 6724–6729. [Google Scholar]
  148. Yin, HS; Ai, SY; Xu, J; Shi, WJ; Zhu, LS. Amperometric biosensor based on immobilized acetylcholinesterase on gold nanoparticles and silk fibroin modified platinum electrode for detection of methyl paraoxon, carbofuran and phoxim. J. Electroanal. Chem 2009, 637, 21–27. [Google Scholar]
  149. Mensah-Brown, AK; Wenzel, MJ; Josse, FJ; Yaz, EE. Near real-time monitoring of organophosphate pesticides in the aqueous-phase using SH-SAW sensors Including Estimation-Based Signal Analysis. IEEE Sens. J 2009, 9, 1817–1824. [Google Scholar]
  150. Khenifi, A; Derriche, Z; Forano, C; Prevot, V; Mousty, C; Scavetta, E; Ballarin, B; Guadagnini, L; Tonelli, D. Glyphosate and glufosinate detection at electrogenerated NiAl-LDH thin films. Anal. Chim. Acta 2009, 654, 97–102. [Google Scholar]
  151. Hildebrandt, A; Ribas, J; Bragos, R; Marty, JL; Tresanchez, M; Lacorte, S. Development of a portable biosensor for screening neurotoxic agents in water samples. Talanta 2008, 75, 1208–1213. [Google Scholar]
  152. Hildebrandt, A; Bragos, R; Lacorte, S; Marty, JL. Performance of a portable biosensor for the analysis of organophosphorus and carbamate insecticides in water and food. Sensor. Actuat. B-Chem 2008, 133, 195–201. [Google Scholar]
  153. Wang, ZG. Electrochemical Investigation of Methyl Parathion at a Poly-L-cysteine Film-modified Glassy Carbon Electrode. Chem. Anal. (Warsaw, Pol.) 2009, 54, 403–414. [Google Scholar]
  154. Gong, JM; Liu, T; Song, DD; Zhang, XB; Zhang, LZ. One-step fabrication of three-dimensional porous calcium carbonate-chitosan composite film as the immobilization matrix of acetylcholinesterase and its biosensing on pesticide. Electrochem. Commun 2009, 11, 1873–1876. [Google Scholar]
  155. Shimomura, T; Itoh, T; Sumiya, T; Mizukami, F; Ono, M. Amperometric biosensor based on enzymes immobilized in hybrid mesoporous membranes for the determination of acetylcholine. Enzyme Microb. Tech 2009, 45, 443–448. [Google Scholar]
  156. Upadhyay, S; Rao, GR; Sharma, MK; Bhattacharya, BK; Rao, VK; Vijayaraghavan, R. Immobilization of acetylcholineesterase-choline oxidase on a gold-platinum bimetallic nanoparticles modified glassy carbon electrode for the sensitive detection of organophosphate pesticides, carbamates and nerve agents. Biosens. Bioelectron 2009, 25, 832–838. [Google Scholar]
  157. Yang, XF; Lin, L; Sun, D. electrochemical behaviors and analysis of parathion at nano-alumina film modified electrode. Chin. J. Anal. Chem 2009, 37, 1057–1060. [Google Scholar]
  158. Du, D; Wang, J; Smith, JN; Timchalk, C; Lin, YH. biomonitoring of organophosphorus agent exposure by reactivation of cholinesterase enzyme based on carbon nanotube-enhanced f low-injection amperometric detection. Anal. Chem 2009, 81, 9314–9320. [Google Scholar]
  159. Lima, KMG; Raimundo, IM, Jr; Pimentel, MF. Improving the detection limits of near infrared spectroscopy in the determination of aromatic hydrocarbons in water employing a silicone sensing phase. Sensor. Actuat. B-Chem 2007, 125, 229–233. [Google Scholar]
  160. Silva, AMS; Pimentel, MF; Raimundo, IM, Jr; Almeida, YMB. A PVC sensing phase for determination of BTEX in water employing mid-infrared spectroscopy. Vib. Spectrosc 2008, 46, 39–44. [Google Scholar]
  161. Carvalho, ER; Correa, AA; Filho, NC; Oliveira, ON, Jr; Gomes, HL; Mattoso, LHC; Martin-Neto, L. Detection of chloroform with a sensor array consisting of electrochemically deposited polythiophenes films: Processes governing the electrical response. Sens. Lett 2007, 5, 374–379. [Google Scholar]
  162. Kurup, PU. Novel technologies for sniffing soil and ground water contaminants. Curr. Sci. India 2009, 97, 1212–1219. [Google Scholar]
Table 1. Summary of the superior limit concentrations (μg/L) of metallic elements, recommended or statutory in fresh water.
Table 1. Summary of the superior limit concentrations (μg/L) of metallic elements, recommended or statutory in fresh water.
Surface water (good chemical status)Drinking water
WFD [2] AA-EQSINERIS** PNECSEQ-Eau [32]Order in Council 2001-1220 [33]SEQ Underground water
Cd0.08–0.250.210.00151
Hg0.050.240.00710.5
Ni200.50.252010
Pb7.250.21105

*AA-EQS: Environmental quality standards expressed as an annual average value;**INERIS values are PNEC (Previsible Non Effect Concentration) from the following reports: Cd [29]; Hg [30]; Ni [31]; Pb [32]. SEQ-Eau: Système d’Evaluation de la Qualité de l’eau (water quality evaluation system) developed by the French Water Agencies.

Table 2. Analytical characteristics of electrochemical sensors for cadmium determination in water, published between 2007 and 2009.
Table 2. Analytical characteristics of electrochemical sensors for cadmium determination in water, published between 2007 and 2009.
MethodTested sampleLoD (μg/L)RSD (%)Ref.
DPASV on multiwalled carbon nanotubes/Bi film modified glassy carbonTap, river & spring waters0.24.6[34]
DPASV on boron-doped diamondAcetate buffer0.4nr[35]
DPASV on bismuth film/graphite electrodeAcetate pH 40.5nr[36]
SWASV on hydroxyapatite modified carbon paste electrodeTap water0.53.8[37]
Inhibition of urease immobilized on screen-printed 10% rhodinized carbon electrodeStandards0.6nr[38]
DPASV on bismuth/poly(p-aminobenzene sulfonic acid) film electrodeTap water0.63.4[39]
SWASV on antimony film carbon paste electrodeSpiked lake water0.83.8[40]
DPASV on multiwalled carbon nanotubes-sodium dodecyl benzene sulfonate modified stannum film electrodeTap water0.94.5[41]
Conductometric measurement of alkaline phosphatase activity of the microalgae Chlorella vulgaris on two gold interdigitated electrodesWastewater1.0nr[42]
SWASV on bismuth-film electrodeRiver water1.04.4[43]
LASV on a zeolite NH4-Y modified carbon paste electrodeGround & wastewaters1.07[44]
SWASV on polymer-modified glassy carbon electrodePlating wastewater1.02.5[45]
Conductometric measurement of alkaline phosphatase activity of the microalgae Chlorella vulgaris on two platinum interdigitated electrodesTris-HCl buffer1.07[46, 47]
SWASV on square array of 8 × 8 screen-printed carbon-based microelectrodesRiver1.318[48]
DPASV on glassy carbon electrode modified with a thiacalix[4]arene filmRiver & lake waters2.23.2[49]
SWASV on bismuth nanoparticle modified boron doped diamondStandard pH 1.22.3nr[50]
SWASV on glassy-carbon electrodes functionalized with composite of Nafion and thiol self-assembled monolayers on mesoporous supportsRiver water & seawater2.55[51]
SWASV on screen-printed graphite electrodeSeawater2.92.3[52]
Cyclic voltammetry of invertase-glucose oxidase activities on platinum microelectrodePhosphate buffer3.0nr[53]
SWASV on carbon disk screen-printed electrode with Bi ionsWastewaters Spiked tap water3.610[54]
SWASV on nitrogen doped diamond-like carbon microelectrode arrayAcetate pH 44.6nr[36]
DPASV on screen printed carbon electrodeMine water94.5[55]
ASV on bismuth electrodePore & ground water9.32[56]
ASV on screen-printed bismuth oxide/graphite electrodeWastewater169.1[57]
DPASV on tape ion sensor based on metal ion transfer reactions at the water/gel micro-interfaceStandards20nr[58]
LSASV on antimony nanoparticle modified boron doped diamondStandard pH 138nr[59]
Potentiometry on ion-selective electrodes with PVC/4-hydroxy salophen membraneRiver water & wastewater1001[60]
Potentiometry on ion-selective electrodes with PVC/thiacalix[4]arene membranesRiver water120nr[61]
Conductometric measurement of alkaline phosphatase activity in enzyme membranes on gold interdigitated electrodesTris-nitrate buffer5004[62]

*LoD: Limit of Detection; RSD: Relative Standard Deviation; nr: not reported; DPASV: Differential Pulse Anodic Stripping Voltammetry; SWASV: Square-Wave Anodic Stripping Voltammetry; LSASV: linear-scan anodic stripping voltammetry.

Table 3. Analytical characteristics of optical sensors for cadmium determination in water, published between 2007 and 2009.
Table 3. Analytical characteristics of optical sensors for cadmium determination in water, published between 2007 and 2009.
MethodTested sampleLoD (mg/L)RSD (%)Ref.
Absorbance of 4-hydroxy salophen on triacetyl cellulose membrane at 431 nmWastewater & river water0.062.9[63]
Fluorescence enhancement of porphyrin-terpyridine complexRiver waters0.134.1[64]
Nine cross-reactive sensing fluorescent elements on micro-plate & data processed by pattern recognitionMineral waters0.56nr[65]
Photoluminescence enhancement of CdS:Mn/ ZnS Qdots in 1,10-diaza-18-crown-6 shellStandard11nr[66]
Fluorescence enhancement of 8-hydroxyquinoline (319/410 nm)Standard3,000nr[67]

*LoD: Limit of Detection; RSD: Relative Standard Deviation; nr: not reported; DPASV: Differential Pulse Anodic Stripping Voltammetry; SWASV: Square-Wave Anodic Stripping Voltammetry; LSASV: linear-scan anodic stripping voltammetry.

Table 4. Analytical characteristics of colorimetric sensors for mercury determination in water, published between 2007 and 2009.
Table 4. Analytical characteristics of colorimetric sensors for mercury determination in water, published between 2007 and 2009.
λab (nm)ChromophoreTested sampleLoD (μg/L)RSD (%)Ref.
6502-Mercapto-2-thiazoline and chromo-ionophore/PVC membraneRiver waters0.010.76[71]
5804-Phenyl-2,6-bis(2,3,5,6-tetrahydrobenzo[b] [1,4,7]-trioxononin-9-yl)pyrylium perchlorate/PVC membrane, (turn-on)Shaft & fountain waters0.020.04[72]
6632-Mercaptopyrimidine in PVC membrane (turn-off)River waters0.080.42[73]
4053,30,5,50-Tetramethylbenzidine on DNA probe (turn off)Spiked creek waters0.2nr[74]
5204(2-Pyridylazo)resorcinolSpiked water0.225[75]
450Tetraphenylporphinetetrasulfonic acid on mesoporous monolithBuffer pH 91.20.4[76]
557Tetrapyridine on “dip-sticks (turn-off)Standards5nr[77]
400/530Spiropyran probe (turn on).Spiked tap water20nr[78]
4294-HydroxysalophenTap & river water261.8[79]
665HexathiacyclooctadecaneTap & river water402.4[80]
642Squaraine dye derivativesStandards100nr[81]
5601-[2-Pyridylazo]-2-naphthol in triacetyl celluloseRiver & waste waters1603.1[82]
254Hydrophobic acid fraction in DOC used as proxy of dissolved HgRiver & lake water300[69]
520/300Naphthalimide derivative on gold nanoparticlesStandard500nr[83]
490Cyclotriveratrylene derivativeRiver water1,0000.9[84]
530Sensitised alumina cladding and PLS modelStandards1,000[85]
542Azo-coupled macrocyclic dye, on a silica nanotubeStandards20,000nr[86]

*LoD: Limit of Detection; RSD: Relative Standard Deviation; nr: not reported.

Table 5. Analytical characteristics of fluorimetric sensors for mercury determination in water, published between 2007 and 2009.
Table 5. Analytical characteristics of fluorimetric sensors for mercury determination in water, published between 2007 and 2009.
λex/ λem (nm)ChromophoreTested sampleLoD (μg/L)RSD (%)Ref.
510/581Rhodamine on thioglycolic acid modified gold nanoparticles (−)River water0.011.2[87]
485/535Polarization on DNA functionalized gold nanoparticles (+)Spiked river water0.2nr[88]
500/550Rhodamine derivatives (+)Spiked waters1nr[89]
475/510carboxyfluorescein on 4-([4-(dimethylamino)phenyl]azo) benzoic acid binding on DNA probesPond water1nr[90]
401/538Naphthalimide derivative of 2,6-bis(aminomethyl)pyridineTap waters1.42.3[91]
520/600Polythymine oligonucleotide T33, citrate-capped gold nanoparticlesPond water28[92]
273/519–527T–T mismatch base pairs, (−)Spiked river water4[93]
425/646–603Prophirin-quinoline (turn-off at 646 nm and turn on at 603 nm)Spiked river waters4.4nr[94]
490/510Oligonucleotides, DNA intercalators, and conjugated polymersStandards6.4nr[95]
228–280/330CdS-encapsulated DNA nanocomposite (−)Wastewater8[96]
438/673Porphyrin/PVC membrane (−)Spiked river water84[97]
273/519–527T–T mismatch base pairs (+)Buffer8.4[98]
480/525DNA-functionalized gold nanoparticles & OliGreenSpiked pond water10[99]
499/624Seminaphthofluorescein chromophore. (+)River waters104.5[100]
365/488Coumarinylalkyne (−)Buffer pH 7.420[101]
458/5342-[(Aminoethylthio)propylthio]ethanamine on 7-nitro-benzo-2-oxa-1,3-diazolyl moieties (+)Acetonitrile/water20[102]
415/ratio 525/650Naphthalimide-porphyrin hybrid (+)River waters204.4[103]
633/ratio 501/403Polymer thiourea-thiadiazole-pyridine (+)Water/Tetrahydrofuran22.6nr[104]
490/591–520CdSe quantum dots surface-modified with triethanolamineStandards36nr[105]
352/500Quinolinocyclodextrin derivative (−)Standards60nr[106]
246/309N,N′-(1,4-phenylenedimethylidyne) bis-1,4-benzene-diamine (−)Ethanol/water 9:1120nr[107]
303/4878-Benzyloxyquinolinebased ester (−)Standards520nr[108]
300–400/Nine cross-reactive sensing fluorescent elements on micro-plate and data processed by pattern recognition (−)Mineral waters20,000nr[65]

*LoD: Limit of Detection; RSD: Relative Standard Deviation; nr: not reported; (−) turn off, (+) turn on

Table 6. Analytical characteristics of electrochemical sensors for mercury determination in water, published between 2007 and 2009.
Table 6. Analytical characteristics of electrochemical sensors for mercury determination in water, published between 2007 and 2009.
MethodTested sampleLoD (μg/L)RSD (%)Ref.
DPASV on Boron-doped diamond like carbon electrodeStandard pH 10.00210[112]
DPASV on screen-printed carbon paste microelectrode chipStandard solutions0.03nr[113]
Amperometric measure of ammonia from urease on rhodinized carbon electrodeStandards0.1nr[38]
DPSV on labeled DNA gold nanoparticlesTap & river waters0.14[111]
Cyclic voltammetry on invertase-glucose oxidase modified Pt electrode (inhibition)Spiked waters2nr[53]
Field Effect Transistor on single walled carbon nanotube sensorStandards2nr[114]
Potentiometry on 2-amino-6-purinethiol or 5-amino-1,3,4-thiadiazole-2-thiol/PVC membraneNatural waters8.83[115]
SWASV on 2-aminothiazole modified carbon paste electrodeTap & waste waters1210[116]
Potentiometry on substituted thiourae-functionalized nanopourous silicaWastewater144.5[117]
DPSV on labeled DNA with a ferrocene on a gold electrode surfaceBuffers20nr[118]
Potentiometry on bis(benzoyl acetone) diethylene triamine/PVC membraneNatural & waste waters200.5[119]
Potentiometry on salophen modified carbon paste electrodeNatural waters300.8[120]
LASV on Nitrogen doped tetrahedral amorphous carbon thin films/siliconeStandard pH 1200nr[121]
ASV on chitosan modified carbon paste electrodeSpiked waters1305[122]
Chrono-amperometry on glucose oxidase in poly-o-phenylenediamine film (inhibition)Standard5006[123]

*LoD: Limit of Detection; RSD: Relative Standard Deviation; nr: not reported; DPASV: Differential Pulse Anodic Stripping Voltammetry; SWASV: Square-Wave Anodic Stripping Voltammetry; LSASV: linear-scan anodic stripping voltammetry.

Table 7. Analytical characteristics of sensors for nickel determination in water published between 2007 and 2009.
Table 7. Analytical characteristics of sensors for nickel determination in water published between 2007 and 2009.
MethodTested sampleLoD (μg/L)RSD (%)Ref.
SWASV on bismuth-film electrode after complexation with dimethylglyoximeRiver water0.12.3[124]
DPASV on carbon screen-printed micro-electrode chipAcetate buffer0.5nr[113]
Conductometric measurement of alkaline phosphatase activity of the microalgae Chlorella vulgaris on two gold interdigitated electrodesWastewater1nr[42]
Colorimetry (ratio 540 nm/396 nm) of glutathione-stabilized silver nanoparticlesStandard600nr[125]
Conductometric measurement of alkaline phosphatase activity in enzyme membrane on gold interdigitated electrodesTris-nitrate buffer5,0004[62]
Nine cross-reactive sensing fluorescent elements on micro-plate & data processed by pattern recognitionMineral waters6,000nr[65]

*LoD: Limit of Detection; RSD: Relative Standard Deviation; nr: not reported; DPASV: Differential Pulse Anodic Stripping Voltammetry; SWASV: Square-Wave Anodic Stripping Voltammetry; LSASV: linear-scan anodic stripping voltammetry.

Table 8. Analytical characteristics of electrochemical sensors for lead determination in water, published between 2007 and 2009.
Table 8. Analytical characteristics of electrochemical sensors for lead determination in water, published between 2007 and 2009.
MethodTested samplesLoD (μg/L)RSD (%)Ref.
DPASV on multiwalled carbon nanotubes/Bi film modified glassy carbon electrodeTap & river water0.14.1[34]
Potentiometry on 5,5,dithiobis(2-nitrobenzoic acid)/PVC membraneMine water0.1220[128]
SWASV on hydroxyapatite modified carbon paste electrodeTap & wastewater0.164.1[126]
SWASV on antimony film carbon paste electrodeLake water0.21.2[40]
LASV on graphite felt electrodeStandards0.215[127]
DPASV on carbon paste electrode modified with biomolecular chitosanTap water0.33.5[130]
SWASV on bismuth-film electrodesRiver water0.54.4[43]
DPASV on bismuth film/graphite electrodeAcetate pH 40.5nr[36]
DPASV on bismuth/poly(p-aminobenzene sulfonic acid) film electrodeTap water0.803.9[39]
SWASV on carbon disk screen-printed electrode with Bi ionsWastewater0.97.4[54]
Conductometric measurement of alkaline phosphatase activity of the microalgae Chlorella vulgaris on two gold interdigitated electrodesWastewater1nr[42]
SWASV on polymer-modified glassy carbon electrodePlating wastewater12.5[131]
DPASV on boron-doped diamondAcetate buffer1.1nr[35]
DPASV on glassy carbon electrode modified with p-tert-butylthiacalix[4]areneLake & tap waters1.72.9[49]
SWASV on screen-printed graphite electrodeSeawater1.84.9[52]
SWASV on bismuth nanoparticle modified boron doped diamondStandard pH 1.21.9nr[50]
SWASV on glassy-carbon electrodes modified with Nafion & thiol monolayer composite on mesoporous supportsRiver & ground waters2.75[51, 132]
LASV on a zeolite NH4-Y modified carbon paste electrodeGroundwater & industrial wastewater3.67[44]
SWASV on carbon paste electrode modified with 2-aminothiazoleTap water4.510[116]
SWASV on nitrogen doped diamond-like carbon microelectrode arrayAcetate buffer pH 44.6nr[36]
Cyclic voltammetry of invertase–glucose oxidase activities on platinum microelectrodePhosphate buffer6.2nr[53]
ASV on screen-printed Bismuth oxide/graphite electrodeWastewater85.6[57]
SWASV on bismuth electrodePore & groundwater815[56]
LSASV on antimony nanoparticle modified boron doped diamondStandard pH 118.5nr[59]
Potentiometry on polyaminoanthraquinone microparticle/PVC membraneSpiked rain & tap waters160nr[133]
LASV on Nitrogen doped tetrahedral amorphous carbon thin films/siliconeStandard pH 1210nr[121]
SWASV on polymer-modified glassy carbon electrodePlating wastewater4002.8[45]
Potentiometry on substituted macrocyclic diamides/PVC membraneStandards410[134]
Conductometric measurement of alkaline phosphatase activity in enzyme membrane on gold interdigitated electrodesStandards40,0004[62]

*LoD: Limit of Detection; RSD: Relative Standard Deviation; nr: not reported; DPASV: Differential Pulse Anodic Stripping Voltammetry; SWASV: Square-Wave Anodic Stripping Voltammetry; ; LSASV: linear-scan anodic stripping voltammetry.

Table 9. Analytical characteristics of optical sensors for lead determination in water published between 2007 and 2009.
Table 9. Analytical characteristics of optical sensors for lead determination in water published between 2007 and 2009.
MethodTested sampleLoD (μg/L)RSD (%)Ref.
Fluorimetry (475/518nm) on thrombin-binding aptamer probe labeled with carboxyfluorescein & 4-([4-(dimethylamino)phenyl]azo)benzoic acid (turn on)Pond water0.11.3[90]
Fluorimetric scanning of metal dependent DNAzymes microarraySpiked river water26.7[135]
Colorimetry (434 nm) of 4-hydroxysalophen on cellulose membraneRiver & wastewater182.1[136]
Fluorimetry (365/418 nm) on polyfluorene with two benzo-18-crown-6 side chains (turn off)Standards1,000nr[137]
Colorimetry (530 nm) on diphenylcarbazone/PVC membraneNatural & wastewater1,300nr[138]

*LoD: Limit of Detection; RSD: Relative Standard Deviation; nr: not reported; DPASV: Differential Pulse Anodic Stripping Voltammetry; SWASV: Square-Wave Anodic Stripping Voltammetry; LSASV: linear-scan anodic stripping voltammetry.

Table 10. Summary of the superior limit concentrations (μg/L) in organic substances, recommended or statutory in freshwater.
Table 10. Summary of the superior limit concentrations (μg/L) in organic substances, recommended or statutory in freshwater.
Superficial WaterDrinking Water

WFD [2] AA-EQSSEQ-Eaux [32]Order in Council 2001-1220 [33]SEQ Underground water

Benzene100.51.00.5
Chloroform2.51.21005
Atrazine0.60.020.10.05
Isoproturon0.30.020.10.05
Chlorfenvinphos0.10.00030.10.05
Chlorpyrifos0.030.000050.10.05
DDTpp’0.010.00020.10.05

*AA-EQS: Environmental quality standards expressed as an annual average value. SEQ-Eau : Système d’Evaluation de la Qualité de l’eau (Water Quality Evaluation System) developed by French Water Agencies.

Table 11. Micro-sensors for organophosphate and carbamates (in italics) determination in water. WFD priority organic substance in bold type.
Table 11. Micro-sensors for organophosphate and carbamates (in italics) determination in water. WFD priority organic substance in bold type.
MethodCompoundLoDRef.
Colorimetry on gold nanoparticlesChlorpyrifos
Malathion
20 μg/L
100 μg/L
[146]
Guided shear horizontal surface acoustic wave devices on LiTaO3Phosmet
Parathion
5 mg/L
2 mg/L
[149]
Voltammetry on clay modified electrodes containing Ni2Al-NO3Glyphosate
Glufosinate
169 μg/L
905 μg/L
[150]
Amperometric activity measurement of acetylcholinesterase immobilized on screen-printed graphite electrodeChlorpyrifos2 μg/L[151,152]
Voltammetry on glassy carbon electrode modified with a poly-L-cysteine filmMethyl parathion1.7 μg/L[153]
Amperometric activity measurement of acetylcholinasterase immobilized on gold nanoparticles and silk fibroin modified platinum electrodeMethyl paraoxon
Carbofura
Phoxim
6 ng/L
7 ng/L
0.6 μg/L
[148]
Amperometric activity measurement of acetylcholinesterase immobilized on calcium carbonate-chitosan composite filmMethyl parathion1 μg/L[154]
Amperometric activity measurement of acetylcholinesterase and choline oxidase within separate hybrid mesoporous silica membranesDiazinon-oxon0.37 μg/L[155]
Amperometric activity measurement of acetylcholinesterase on gold-platinum bimetallic nanoparticles onto 3-amino-propyltriethoxy silane modified glassy carbon electrodeParaoxon ethyl
Sarin
Aldicarb
9.5 μg/L
6 μg/L
48 μg/L
[156]
Amperometric activity measurement of acetylcholinesterase on poly(dimethosiloxane)-poly(diallydimethylammonium)/gold nanoparticles composite filmParaoxon
Parathion
0.5 ng/L
1.0 ng/L
[147]
Voltammetry on nano-alumine film modified electrodeParathion0.3 μg/L[157]
Amperometric reactivation measurement of cholinesterase from organophosphorus-inhibited rat saliva on carbon nanotube modified screen printed carbon electrodeParaoxon0.14 μg/L[158]

*LoD: Limit of Detection.

Table 12. Micro-sensors for pesticide determination in water, other than organophosphate and carbamates. WFD priority organic substances in bold type.
Table 12. Micro-sensors for pesticide determination in water, other than organophosphate and carbamates. WFD priority organic substances in bold type.
MethodCompoundLoDRef.
Voltammetry on conducting MIP deposited on an platinum electrodeAtrazine4.5 μg/L[140]
Surface plasmon resonance (SPR) on monoclonal anti-bodies immobilized on a gold sensor surfaceAtrazine20 ng/L[141]
Oxygen production and fluorescence measures on the algae Elodea canadensis cultureAtrazine10 μg/L[142]
Surface plasmon resonance (SPR) on monoclonal anti-bodies immobilized on sensor chip surfaceIsoproturon0.1 μg/L[143]
Resonance characteristics hepatoma cultured on an indium tin oxide surface of quartz crystalParaquat2.7 mg/L[144]
Surface plasmon resonance (SPR) on monoclonal antibodies on gold-thin layerDDT15 ng/L[145]

*LoD: Limit of Detection.

Table 13. Micro-sensors for solvent determination in water.
Table 13. Micro-sensors for solvent determination in water.
MethodCompoundLoDRef.
Infrared spectroscopy on polydimethylsiloxane (PDMS) membraneBenzene0.6 mg/L[159]
Mid-infrared spectroscopy on 2-(2-hydroxy-5-tert-octylphenyl)benzotriazol/PVC membraneBenzene5.0 mg/L[160]
Impedance spectrometry on sensor array made up with electrodeposited polythiophene films onto interdigitated gold electrodesChloroform0.1 mg/L[161]

*LoD: Limit of Detection

Share and Cite

MDPI and ACS Style

Namour, P.; Lepot, M.; Jaffrezic-Renault, N. Recent Trends in Monitoring of European Water Framework Directive Priority Substances Using Micro-Sensors: A 2007–2009 Review. Sensors 2010, 10, 7947-7978. https://doi.org/10.3390/s100907947

AMA Style

Namour P, Lepot M, Jaffrezic-Renault N. Recent Trends in Monitoring of European Water Framework Directive Priority Substances Using Micro-Sensors: A 2007–2009 Review. Sensors. 2010; 10(9):7947-7978. https://doi.org/10.3390/s100907947

Chicago/Turabian Style

Namour, Philippe, Mathieu Lepot, and Nicole Jaffrezic-Renault. 2010. "Recent Trends in Monitoring of European Water Framework Directive Priority Substances Using Micro-Sensors: A 2007–2009 Review" Sensors 10, no. 9: 7947-7978. https://doi.org/10.3390/s100907947

Article Metrics

Back to TopTop