Anda di halaman 1dari 15

1987

Review

Ion-Selective Electrode Potentiometry in Environmental Analysis


Roland De Marco,a* Graeme Clarke,a Bobby Pejcicb
a

Nanochemistry Research Institute, Department of Applied Chemistry, Curtin University of Technology, GPO Box U1987, Perth, Western Australia, 6845, Australia *e-mail: r.demarco@exchange.curtin.edu.au b CSIRO Petroleum, ARRC, PO Box 1130, Bentley, Western Australia, 6102, Australia Received: May 16, 2007 Accepted: May 29, 2007 Abstract This review will illustrate how it is possible to develop ion-selective electrode (ISE) methodologies that meet the stringent requirements (i.e., high selectivities and very low detection limits) for the analysis of important analytes in the environment, and will present a variety of examples on the application of ISEs in environmental analysis. Despite the experimental biases that have limited the analytical performance of ISEs through apparently high detection limits and modest selectivities, there has been a plethora of research in the application of ISEs in the monitoring of environmentally important trace metals and anions in natural waters and soils. Most popular has been the analysis of free metals in natural waters, as this parameter is known to be a master variable in the uptake and toxicology of trace metals on aquatic biota reflecting the bioavailability of trace metals in the environment. Furthermore, as copper is a major trace metal in coastal waters due to its extensive use in antifouling paints on sea vessels and structures, there are many reports in the literature on the use of the copper ISE in assays of either free copper or the copper complexing capacity of natural waters and soil peats. Moreover, there have been a variety of studies showing a strong correlation between free copper levels and the toxicity of copper on a variety of marine and fresh water organisms. Nevertheless, there are numerous reports in the literature that have used ISEs to monitor important anions such as fluoride, phosphate, sulfate, nitrate, nitrite, chloride, cyanide, etc., as well as other important cations such as ammonium and chromium(VI) in waste and natural waters. In conclusion, this review will present new and interesting perspectives on the application of ISEs in environmental analysis using approaches such as real-time remote monitoring of water quality, shipboard monitoring of environmentally important analytes using flow analysis instrumentation, the use of robust all-solid-state ISEs in submersible instruments for long-term deployment in the field, and innovative analytical approaches such as backside calibration and switchtrodes that avoid standard addition analysis and the concomitant perturbation in analyte speciation in natural samples. Keywords: Ion-selective electrode, Trace metals, Metal speciation, Natural waters, Field monitoring DOI: 10.1002/elan.200703916

1. Introduction
Despite the outstanding potential of ion-selective electrodes (ISEs) for the analysis of environmentally important analytes such as trace metals, phosphate, nitrite, nitrate, etc., it is perceived that ISE devices generally lack the sensitivity and selectivity needed for the analysis of trace analytes in complex and challenging samples such as seawater, estuarine waters, rivers, lakes, soils, etc. Notwithstanding, there were several excellent papers in the 1970s and 1980s [1 4] demonstrating that it is possible to use a crystalline membrane copper ISE in the analysis of nanomolar levels of copper in natural waters, as long as the ISE is handled correctly, so as to minimize the dissolution of the ISE and the concomitant carry-over of copper from sample to sample. Clearly, these seminal papers were ahead of their time, and laid the foundations for the use of crystalline membrane ISEs in environmental analysis. With regard to polymeric membrane ISEs, Sokalski et al. [5] presented a landmark paper in 1997 on the influence of
Electroanalysis 19, 2007, No. 19-20, 1987 2001

transmembrane fluxes on the detection limits and selectivities of ionophore-based ISEs and demonstrated that this uncompensated experimental bias was responsible for the apparently mediocre sensitivity and selectivity of this class of ISEs. Clearly, this research paved the way for the creation of new and improved polymeric ISEs with vastly improved detection limits and selectivities, enabling the ISE analysis of trace constituents in the environment. Accordingly, there have been several papers on the application of this new class of polymeric ISE in the analysis of trace metals in the environment [6 8]. Regarding the analysis of trace metals in the environment, the authors view is that the great virtue of ISEs is their ability to sense the free metal ion activity, which is widely recognized as a master variable responsible for the uptake and toxicity of metals by biota [9] (see Fig. 1), thereby providing an analytical technique capable of monitoring the impact of trace metal inputs in the environment. Notably, the free metal content of environmental waters is regulated by the metal buffering ability of the natural water [9], as
 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1988

R. De Marco et al.

Fig. 1. Speciation model for metals in the environment (redrawn from Buffle et al. [9]).

dictated by the inorganic and organic speciation of metals in the sample, and this may also be estimated using the socalled metal complexation capacity that is determined by monitoring metal titrations of natural waters using ISEs. Clearly, the speciation of a trace metal with the inorganic (i.e., OH, HCO3, CO32, Cl and SO42) and organic (i.e., HzLn where n denotes the possibility of 1, 2, 3, etc. different chelating strength organic ligands depending on the sample) ligands in natural waters is crucial in regulating the level of free metal, and the concomitant bioavailability of the trace metal. The speciation of trace metals may be generalized satisfactorily using the following reaction scheme [10]: H OH ! H2O H HCO3 ! H2CO3 H CO32 ! HCO3 Mg2 CO32 ! MgCO3 Ca2 CO32 ! CaCO3 Mn zCO32 ! M(CO3)zn2z M zHCO ! M(HCO )
n 3 nz 3 z

the metal-ligand formation equilibrium constants and concentrations of ligands via ISE metal titrations of the natural medium. Either way, an ISE provides a useful analytical tool for investigating the trace metal speciation of environmental waters, which is crucial in determining the bioavailability of metals in the environment. Most significantly, these equilibria also demonstrate that acidification of a natural water sample is expected to breakdown the inorganic and organic metal complexes in the sample, thereby ensuring that an ISE is capable of determining the concentration of total metal in a natural sample, if it has been acidified to a low enough pH (viz., pH < 2). Consequently, ISEs are a powerful research tool in environmental monitoring, as they permit a simultaneous monitoring of metal inputs, as reflected by the level of total metal, and the bioavailability of these metal inputs as ascertained from the level of free metal. Like many electrochemical sensors, ISEs are extremely attractive in environmental science since their simple measurement principles and portability make them suitable for on-site, shipboard and/or in-situ field analyses of environmentally important analytes. Significantly, an ISE potentiometric sensor only requires a measurement of the electrode potential at near zero current, thereby ensuring that simple and inexpensive instrumentation is required in field studies, and this is a highly desirable feature of ISEs in environmental analysis. Furthermore, there are ISEs for virtually every known cation and anion, so it is possible to assemble an array of ISEs that is capable of covering a broad range of environmentally important trace species, both cationic and anionic, providing the environmental scientist with a monitoring dataset that is rich in chemical information. Notwithstanding, the ISE electroanalysis of complex environmental samples like lakes, rivers, seawater, soils, etc. is a challenging task due to a wide range of potential problems, as discussed elsewhere by De Marco et al. [11]: i) electrode fouling by the sample matrix components (e.g., chloride, hydroxide, organic ligands, other cation or anion interferences, etc.) causing erroneous response characteristics; electrode drift leading to poor reproducibility as the sensor surface is altered on continuous exposure to real samples; electrode dissolution causing a high surface excess of analyte that degrades the detection limit of the ISE beyond the realms of trace analyses in the environment; electrode carry-over of the analyte which can cause cross-contamination of samples with the trace analyte; electrode instability as passivation of the ISE in real samples causes the response characteristics to become erratic and non-representative of the sample assayed.

ii)

iii)

yMn zOH ! My(OH)zynz Mn zCl ! MClznz Mn zSO42 ! M(SO4)zn2z Mn HzLn ! MLnnz zH In essence, the aforesaid speciation equilibria can be probed reliably by using an ISE to measure either the concentration of free metal under equilibrium conditions, or by elucidating

iv) v)

The problems of electrode drift and electrode dissolution can be ameliorated by using an ISE that has been conditioned in the sample matrix, typically overnight, to produce a stable, reproducible and fast responding ISE, also minimizing sample contact times and the concomitant ISE release of analyte into the sample. The cross-contamination of samples

Electroanalysis 19, 2007, No. 19-20, 1987 2001 www.electroanalysis.wiley-vch.de  2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Ion-Selective Electrode Potentiometry in Environmental Analysis Table 1. Examples of applications of ISEs in environmental analysis. ISE Cu ISE Cu ISE Fe ISE Cd ISE Hg ISE F ISE Pb ISE Cr ISE CN ISE UO22 ISE Anion, cation & gas ISEs H2PO4 & Sm3 ISE Mg ISE Cu ISE TFP ISE Anion ISEs Environmental application Analysis of Cu2 in natural waters Copper complexation capacity in natural waters and soil peats Analysis of Fe3 in natural waters Analysis of Cd2 in natural waters Analysis of Hg2 in natural waters Analysis of F in natural waters Analysis of Pb2 in natural waters and culture media Analysis of HCrO4 in natural waters Analysis of CN in natural waters Analysis of UO22 in tap water and seawater Analysis of CN, NO2, F, NH4 and NH3 in natural waters Analysis of PO43 directly or indirectly by titration in natural waters Analysis of Mg2 in seawater Determination of water hardness Back titration of excess 2-aminoperimindium ion for SO42 in seawater Analysis of corrosive anions (e.g., Cl, NO3 and SO42) on electronic equipment or in reinforced concrete structures References

1989

[3, 19, 27, 33, 43] [69 74] [75, 76] [77 80] [81 83] [84 87] [6, 7] [88, 89] [90 92] [93] [94 96] [97, 98] [99] [100] [101] [102, 103]

via electrode carry-over is easily overcome by employing a measurement protocol in which the electrode is cleansed in a low analyte bearing solution (e.g., a sacrificial calibration standard comprising ultratrace levels of the analyte below those expected in the natural sample) as well as a sample washing regime where the ISE is pre-conditioned in multiple batches of the sample to be analyzed until a comparable response is obtained between wash samples indicating that the analyte has not been carried over into the sample. Frequent calibration and consequent rejuvenation of the ISE is needed if electrode fouling and electrode passivation is evident upon continuous exposure of the ISE to real samples, and for crystalline membrane ISEs this would comprise frequent repolishing of the sensor surface, while for polymeric ISEs this may necessitate the adoption of a fresh ISE membrane. Nevertheless, these electrode fouling and electrode passivation problems may be minimized by using an ISE in flow-injection analysis (FIA) or continuous flow analysis (CFA) where a shearing of adsorbed species from the electrode surface under the influence of hydrodynamic flow can cause the desorption of foulants and the prevention of electrode fouling [12]. Despite a preponderance of research on the monitoring of trace copper in the environment using a crystalline membrane ISE, there is a wide range of analytical applications of ISEs in environmental analysis including the analysis of other important trace metals such as iron(III), mercury(II), cadmium(II), lead(II) and chromium(VI), as well as important anions such as fluoride, phosphate, sulfate, nitrate, nitrite, chloride and cyanide. Table 1 summarizes the analytical applications of ISEs in environmental science. Notwithstanding, this review will necessarily emphasize the ISE electroanalysis of copper in the environment, as this represents about two thirds of the research undertaken in this field, but other environmental analyses with ISEs will also be discussed since they demonstrate the general utility and great potential of ISEs for environmental analysis.

As there are excellent general review articles on crystalline and polymeric membrane ISEs [13 18], the present review will focus specifically on the application of ISEs in environmental analysis. It is important to note that this topic has not been addressed adequately since the suitability of ISEs for environmental analysis has only been realized in recent times.

2. Copper in Natural Samples


Copper is a major trace metal in the environment due to its extensive use in antifouling paints, and poses serious environmental hazards due to its strong toxicity at excessive levels [19]. Moreover, copper is an essential trace metal nutrient at ambient levels [20, 21], and it has been demonstrated using a variety of analytical techniques and aquatic organisms [20 24] that the inhibition of growth or toxicity of copper on phytoplankton is related to free copper(II) or Cu2, and not the concentration of total dissolved copper. This is the single most important feature that makes the copper ISE so attractive in environmental science leading to considerable research in the development of reliable copper ISE methods for the analysis of natural samples (i.e., seawater, lakes, rivers, soils, etc.) [1 4, 11, 12, 19 22, 25 68].

2.1. Free and Total Copper It is significant to note that early and seminal papers using the crystalline membrane copper ISE demonstrated the utility of this ISE for the analysis of trace copper in the environment [1 4]. Notably, Smith and Manahan [2] demonstrated that standard addition potentiometry with the Cu ISE can be used to measure accurately Cu2 levels down to 108 M in tap water and natural water samples, as

Electroanalysis 19, 2007, No. 19-20, 1987 2001 www.electroanalysis.wiley-vch.de  2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1990 long as the sample containers are cleaned rigorously in concentrated acids and a dilute solution of EDTA, ultrapure reagents are used, and the electrode release of copper is minimized through the use of a complexing antioxidant buffer. Similarly, Jasinski et al. [3] showed that the copper ISE can be used to analyze residual amounts of copper as low as 3 109 M in artificial and real seawater, if precautions are taken to minimize the contamination of samples using clean techniques and reagents, as well as employing highly stirred solutions to facilitate the dispersal of electrode released copper out of the electrodes diffusion layer into the bulk sample. Most significantly, Jasinski et al. [3] showed that a standard addition analysis of seawater at pH 8 yielded a characteristic titration shaped curve and a concomitant super-Nernstian response for the Cu ISE, noting that this effect is indicative of copper complexation with ubiquitous seawater ligands, but theoretical Nernstian response was evident in the same seawater sample at pH 3 due to a negation of the complexation capacity of seawater ligands due to protonation of their metal coordinating functional groups. Notably, the research of Barica [25] in Canadian Prairies at pH 7.8 to 8.6 also demonstrated a double Nernstian response to added copper(II), but this anomalous response was found to be highly reproducible and useable when the ISE was calibrated under these analytical conditions. The pioneering work by Zirino and coworkers [1, 4] demonstrated that the potential of the Cu ISE in seawater, indicative of free copper, correlated beautifully with anodic stripping voltammetry (ASV) determinations of total copper, and that the copper ISE is susceptible to oxidation by oxygen, but its effect is minimal, and the authors subsequently adopted this combined ISE and ASV approach in the shipboard monitoring of free and total copper in samples taken along the coast from San Diego USA to Pisco Peru. Essentially, the outstanding potential of the copper(II) ISE for the analysis of Cu2 in natural waters was demonstrated in this very early research, and provided analytical chemists with an excellent foundation for the development of reliable ISE methods for environmental analysis. Camusso et al. [26] reported an interesting study on variations in the activity of Cu2 in Lake Orta Italy as a function of depth, as monitored using a conductivitytemperature-depth (CTD) device fitted with a jalpaite copper(II) ISE, along with pH, oxygen and conductivity sensors. In this important paper, it was shown that respiration below the surface layer is responsible for a diminution in dissolved oxygen levels and a concomitant decrease in solution pH due to nitrate formation via the oxidation of ammonium ions, and the Cu2 levels in the lake mirrored the expected changes in speciation as acidification led to dissociation of natural copper(II) complexes into free copper(II). In essence, this study highlighted the tremendous potential of the copper(II) ISE for in-situ field studies of copper in the environment using submersible instrumentation. In a landmark paper, Belli and Zirino [19] studied the response of the crystalline membrane copper ISE in real and

R. De Marco et al.

artificial seawater. It was found that a standard addition analysis of real and artificial seawater at pH 8 yielded a super-Nernstian response to added copper, but the theoretical Nernstian response was obtained in both media at pH 2. Conversely, the response of the copper ISE in the spiked artificial seawater at pH 8 yielded the theoretical Nernstian response when plotted against the MICROQL calculated free copper levels obtained using the well-known copperinorganic speciation of artificial seawater. The authors used simulations for copper speciation in seawater to prove that the apparent super-Nernstian response of the copper ISE in standard addition analyses of seawater is due to forcing the ISE data to fit the incorrect parameter in total copper since the copper partitioning factor that dictates the level of free copper in seawater is variable at different levels of added copper. Last but not least, Belli and Zirino [19] noted that the response of the copper ISE in chloride-based copper buffers was independent of the salinity at the concentrations of seawater, clearly demonstrating that the electrode is free of the usual chloride interference that is problematic during the analysis of saline solutions containing high concentrations of total copper. Clearly, this research demonstrated that the copper ISE can be used in determinations of total copper in acidified seawater. De Marco [27] undertook an evaluation of the response of crystalline membrane copper ISEs in seawater media, and found that all commercially available ISEs yielded a Nernstian response in saline copper buffers in the range 1015 to 109 M Cu2. Notwithstanding, it was found that the electrode employing a crystalline membrane of jalpaite (i.e., Ag1.5Cu0.5S) was the only electrode capable of providing Nernstian response that was collinear with the response obtained in chloride-free unbuffered standards in the range 106 to 101 M Cu2. Clearly, this outcome demonstrated that the response of the jalpaite copper ISE in saline calibration buffers is free of the usual chloride interference effect that is evident at 106 M Cu2 or higher, and De Marco [27] argued that this is due to kinetic limitations of the chloride-induced membrane surface reactions that are contingent on the activity of copper, which is at ultra-trace levels below 109 M. Most significantly, the response of the copper ISE in artificial seawater comprising low levels of total copper (i.e., 3 109 and 3.6 108 M) and 106 M glycine yielded ISE determined free copper levels within 0.2 log(aCu2) units of those calculated using the wellknown complex formation and ionization constants for copper(II)-glycine, copper(II)-hydroxy, copper(II)-carbonate, etc. complexes and their associated ligands. Without doubt, the work of De Marco [27] demonstrated that the copper ISE is capable of detecting free copper in seawater comprising nanomolar concentrations of total copper. De Marco and co-workers [12, 28, 29] conducted important mechanistic studies on the electrode kinetic and surface chemical behavior of the jalpaite copper(II) ISE in seawater. In contaminated and uncontaminated San Diego Bay samples, Zirino et al. [12] noted a surface excess and lower limit of detection of 2 108 M total copper with a static ISE, with the influence of hydrodynamic flow at either a rotating

Electroanalysis 19, 2007, No. 19-20, 1987 2001 www.electroanalysis.wiley-vch.de  2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Ion-Selective Electrode Potentiometry in Environmental Analysis

1991 cleaning protocol entailing exposure of the ISE to a sacrificial copper(II) buffer comprising low activities of Cu2, then it is possible to conduct reliable determinations of Cu2 in seawater media. The aforesaid mechanistic research on the response mechanism of the copper(II) ISE in seawater inspired a number of subsequent studies by Zirino and co-workers [11, 32, 33] into the analysis of copper(II) in seawater using the jalpaite copper(II) ISE. First, De Marco et al. [11] demonstrated that overnight conditioning of the electrode in seawater is needed to stabilize its response in seawater enabling fast response times and short electrode contact times with the sample, thereby minimizing contamination of the sample through electrode dissolution. Furthermore, these authors showed [11] that ISE determined free copper(II) levels in organic-free or UV photooxidized seawater are commensurate with those calculated using the well-known inorganic copper(II) speciation in seawater, and the analysis of seawater at pH 2 using a standard addition method gave ISE total copper levels consistent with those measured using graphite furnace atomic absorption spectrometry (GFAAS). Next, Zirino et al. [32] respectively measured free, total and labile copper(II) levels at various sites in San Diego Bay using ISE, GFAAS and ASV, and found that the data is reconcilable with an equilibrium model for the binding of copper to natural organic ligands, thereby providing stability constants for seawater copper(II)-organo complexes that are comparable with previously published findings. Last, Zirino et al. [33] used a RDE jalpaite copper(II) ISE in the analysis of free copper(II) activities in San Diego Bay demonstrating that the electrode release of copper is diminished substantially at rotation speeds in excess of 4,800 revolutions per minute, thereby allowing a reliable determination of the true thermodynamic activity of Cu2. Most significantly, the authors [33] showed that, between measurements, storage of the ISE in a sacrificial copper(II) buffer in complete darkness minimized electrode carry-over of Cu2 and undesirable photochemical effects with this sensor, thereby providing a substantially improved reproducibility of approximately 0.06 log(aCu2) units. Furthermore, Zirino et al. [33] also showed that continuous bathing of a copper(II) ISE in natural seawater progressively titrated the ubiquitous organic ligands with electrode released Cu2 yielding a breakthrough point in the potentiometric response of the ISE, as evidenced by a sharp rise in the electrode potential that is commensurate with Nernstian response to the GFAAS determined values of released copper(II), signifying that it is possible to utilize the ISE corrosion-based release of copper in autotitrations of the copper complexing capacity. Further credence for this copper ISE autotitration method was provided by the EIS electrode kinetic data in seawater [29] showing that seawater autotitrated by the electrode release of copper(II) correlates with a relative constancy in ISE potential up to the breakthrough point in the autotitration, at which point the copper(II) ISE kinetics are consistent with a response to elevated levels of Cu2. Clearly, these papers provided

disc electrode (RDE) copper(II) ISE or static copper(II) ISE residing in a flowing stream in the wall-jet cell of a CFA or FIA analyzer allowing dispersal of electrode released copper into the bulk solution and lowering the ISE detection limit to approximately 1010 M. In a separate study, De Marco [28] used X-ray photoelectron spectroscopy (XPS) to demonstrate that continuous aging of the membrane in artificial seawater leads to a modification of the sensor surface to a copper deficient sulfide phase together with electrode fouling due to silver chloride formation, and adsorption of the natural organic ligands in seawater alleviates this chloride interference by promoting the peptization of deposited silver chloride, thereby minimizing the seawater-induced fouling of the membrane. A subsequent electrode kinetic study of the jalpaite copper(II) ISE in seawater using electrochemical impedance spectroscopy (EIS) [29] demonstrated that the behavior of the sensor in seawater is compliant fully with the generally accepted response mechanism involving the reduction of copper(II) and concomitant copper(I) ion-exchange at copper sulfide sites. Significantly, the EIS charge transfer kinetics of the ISE in the presence of varying levels of commercial humic acid [29] showed that natural organic ligands exert a mild interference effect on the response of the copper(II) ISE, as evidenced by the ligand-induced facilitation of the copper(II) reduction process, but potentiometric data revealed that the typical level of humic acid in seawater (i.e., 1 mg L1) is tolerable leading to an error of about 0.1 log(aCu2) units. Also noteworthy was the observation that a RDE jalpaite copper(II) ISE yielded a more substantial change in the influence of rotation speed in natural seawater, as predicted using the Levich equation, noting that this observation is symbolic of a hydrodynamic-induced desorption of natural organic ligands from the surface of the jalpaite copper(II) ISE. In any event, an interesting FIA study with a solid-state chalcogenide copper(II) ISE in unbuffered saline copper standards at concentrations above 106 M Cu2 [30] demonstrated that the usual chloride interference associated with complexation of copper(I) generated through the reductive ion-exchange of copper(II) is eliminated by the phenomenon of kinetic discrimination in which the copper(II) and chloride response processes are time resolved in the FIA transient due to significant differences in the kinetics or non-steady state signals of the ISE towards these species. Accordingly, contrary to popular belief that organic ligands interfere strongly with the response of the copper(II) ISE in natural waters [17, 18], this mechanistic research demonstrated that ubiquitous seawater organic ligands exert a mild interference, and can in fact protect the sensor against the chloride-induced fouling of the sensor during continuous exposure to seawater together with the possibility of compensating for these interference effects under the influence of hydrodynamic flow at an RDE or in CFA and/or FIA. In a brief review on the use and misuse of crystalline copper(II) ISEs in seawater, Mackey and De Marco [31] showed that a jalpaite copper(II) ISE pre-equilibrated in real seawater, if handled using clean techniques and a

Electroanalysis 19, 2007, No. 19-20, 1987 2001 www.electroanalysis.wiley-vch.de  2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1992 unequivocal evidence of the utility of the jalpaite copper(II) ISE in precise and accurate determinations of free and total copper in seawater. There are other interesting papers on the use of a jalpaite copper(II) ISE in the analysis of Cu2 in natural waters [34 42]. For example, Stella and Ganzeril-Valenti [34] used a copper(II) ISE to monitor variations in free copper(II) as a function of pH in the Ticino and Po Rivers, and the authors used this data in the elucidation of the inorganic complexation of these natural waters. Significantly, Cabaniss and Shuman [35] as well as Holm and co-workers [36, 37] adopted a combined copper(II) ISE potentiometry and fluorescence quenching (FQ) method in studies of the copper(II) binding of natural organic matter in seawater, rivers and lakes, and these workers demonstrated that both methods gave comparable values for free copper(II) in samples containing low levels of total copper(II) (i.e., < 107 M). Similarly, Rozan et al. [38] used a combination of ASV and ISE to determine free copper(II), natural organic ligands and copper complexing capacities in model copper(II) buffers incorporating ethylenediamine and EDTA, as well as copper binding with Suwanne River and Dismal Swamp fulvic acids, and the results showed that both techniques are able to measure comparable levels of labile copper(II) when the detection limits of the techniques overlapped. Ghode and Sarin [39] illustrated that a copper(II) ISE may be used in copper speciation in natural waters by showing that copper(II) speciation in the presence of polyamino carboxylic acids such as EDTA are internally consistent with the theoretically expected binding of copper(II) by these ligands, while Zolotov et al. [40] showed that a Chelex 100 preconcentration column used in combination with a copper(II) ISE FIA potentiometric technique is capable of providing accurate levels of total copper(II), as compared to GFAAS, in tap water, waste water and seawater. Most recently, Eriksen et al. [41] developed a CFA method for the analysis of free and total copper in Pacific Ocean seawater, and the total copper levels obtained by ISE potentiometry in acidified seawater compared favorably with those determined using GFAAS, while Rivera-Duarte and Zirino [42] used copper(II) titrations of San Diego Bay seawater to deduce the copper(II) complexing capacities and concomitant levels of free copper(II) yielding results that compared favorably with typical values in estuarine and coastal samples. To surmise, the aforementioned research has illustrated the validity of using a copper(II) ISE in determinations of free and total copper(II) in natural waters, and this avails the copper(II) ISE to studies of the bioavailability and toxicological response of copper in environmental science.

R. De Marco et al.

2.2. Free Copper and Toxicology Studies There have been numerous studies on the use of a copper(II) ISE in toxicological studies of aquatic organisms in natural waters [20 22, 43 49]. First, Sunda and co-workers [20 22] investigated the toxicological response of a marine

bacterium [20], an estuarine diatom [21] and unicellular alga [22] in seawater and river water, and found that ISE determined free copper(II) levels correlated very strongly with toxicological data for these aquatic organisms. In a series of similar papers [43 49], the influence of copper(II) complexation on the toxicity of various fresh and marine water species was monitored in parallel with copper toxicity studies. In all cases, the biological measure of toxicity (i.e., LC50, survival rates, embryo hatching rates, copper accumulation by the organism) yielded excellent correlations between toxicity and free copper(II) levels as measured using the copper(II) ISE. Notwithstanding, the ability of the copper(II) ISE for reliable analyses of free copper(II) in samples containing trace levels of copper(II) has been a controversial topic due to the outdated thesis that this ISE possesses a sensitivity of approximately 106 M total copper [74], and this view is sometimes perpetuated by environmental researchers [50] despite the overwhelming evidence that the copper(II) ISE is functional in seawater and other aquatic media. Ultimately, proof that the copper(II) ISE is indeed a reliable tool for the environmental analysis of free copper(II) in natural waters was provided in recent papers by Eriksen et al. [43] and Rivera-Duarte et al. [44] who conducted copper speciation and toxicity studies of a marine diatom and three larval species using the jalpaite copper(II) ISE. In each of these studies, seawater containing nanomolar concentrations of total copper (Macquarie Harbour seawater in Eriksen et als study [43], and San Diego Bay seawater in the study of Rivera-Duarte et al. [44]) was spiked with varying levels of copper, and total and free copper levels were assayed by GFAAS and copper(II) ISE potentiometry respectively, while the toxicological responses towards marine species (viz., growth inhibition of the marine diatom [43] and LC50 values for larval development of three marine invertebrates [44]) were recorded simultaneously. In both studies, there was a clear correlation between the toxicological response of marine species to copper in seawater, and the free copper(II) levels determined using the ISE, not total copper levels assayed using GFAAS. Furthermore, these studies also found that a toxicological response occurs at levels above 1011 M Cu2, and this data is consistent with literature data for toxicological studies of marine species. In conclusion, the body of research on the behavior of the copper(II) ISE in natural waters demonstrates unequivocally that this sensor is suitable for determinations of free copper(II) in natural waters; a master variable that is responsible for the bioavailability and toxicity of copper(II) in the environment. Despite the many alleged artifacts that are inherent in assays of ultratrace levels of free copper(II) in challenging and complex samples such as natural waters, the correlation of copper toxicity data for a variety of aquatic organisms with ISE determined free copper(II) levels, together with the expected threshold for toxicological response above 1011 M Cu2, as accepted widely in the scientific literature, shows that all of the arguments about the aberrant response of the copper ISE are largely

Electroanalysis 19, 2007, No. 19-20, 1987 2001 www.electroanalysis.wiley-vch.de  2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Ion-Selective Electrode Potentiometry in Environmental Analysis

1993 all cases, the data obtained using the ISE and alternate spectrometric methods (e.g., differential spectrometry, ionexchange/GFAAS, dialysis titration, fluorescence quenching, and UV-visible spectrometry) were comparable. A unique and interesting approach has employed a combination of potentiometric stripping analysis (PSA) and copper ISE potentiometry to elucidate the copper(II) complexing capacity, conditional stability constants of copper(II)-organic complexes and free copper(II) in seawater, thereby giving an integrated view of the copper(II) speciation in the presence of natural organic ligands [56, 72, 73]. Xue and Sunda [74] published a seminal paper on comparative measurements of free copper(II) levels, organic ligand concentrations, and the stability constants of copper(II)-organic complexes in lake waters using ligand competition/ASV at three detection windows as well as ISE potentiometry. It is evident that the copper(II) ISE is capable of providing reliable and comparable free copper(II) values in the lake samples investigated at total copper levels in excess of 108 M; however, the ISE yielded higher aCu2 values by about 0.75 log(aCu2) units when the total copper level in the sample dropped to beneath the threshold ISE detection window of 108 M. Significantly, this important research demonstrated unequivocally that, although the metal speciation in natural waters is kinetically controlled and the system is in a steady-state or quasiequilibrium, ISE and ASV are capable of providing quantitative analytical data about the steady-state copper(II) speciation of natural waters, and this information is indeed representative of the true solution chemistry of natural waters. In summary, the aforementioned copper complexation studies in natural waters have highlighted the problems associated with copper ISE in determinations of free copper(II) in environmental samples containing < 108 M total copper, and illustrate that a ligand competition/ASV methodology is ideally suited to these samples. Nevertheless, it is important to note that Zirino et al. [12] demonstrated that the high detection limit of 108 M total copper evident with a static copper(II) ISE through contamination of the electrodes Nernst diffusion layer at this level may be alleviated by using a hydrodynamic flow regime (e.g., RDEs, FIA or CFA) to lower the copper(II) ISEs detection limit to approximately 1010 M. Without doubt, the preferred analytical methodology for the analysis of free copper(II) in environmental samples would utilize a copper(II) ISE in either the RDE, FIA or CFA analysis modes.

irrelevant, and that this (and other) ISE(s) can provide highly useful data in environmental surveys of the fate and impact of copper (and other metals) in the environment.

2.3. Copper Titrations in Determinations of Copper Complexing Capacities The role of ubiquitous organic ligands in natural waters on the complexation of metals is very significant, and it is widely accepted that copper(II) in seawater is almost completely complexed by these chelating ligands [51]. In Mackeys and Zirinos [51] interesting and insightful commentary about the role of copper(II) complexation by the organic ligands in seawater, it was proposed that there is a lack of equilibrium between spiked metals and the ubiquitous organic ligands in seawater, and they cited 10 key experimental observations to support this claim. Accordingly, these authors questioned the merit of ligand competition/metal titration methods (e.g., ASV, CSV, ISE, etc.) that assume inherently the establishment of a thermodynamic equilibrium in the computation of the associated copper(II)-organic complex formation constants and copper(II) complexing capacities. Nevertheless, the following section will demonstrate that this approach is representative of the steady-state or quasi-equilibrium established with the speciation of trace metals in seawater, and this may indeed be indicative of the exact conditions that are controlling the binding of copper(II) and its concomitant bioavailability in seawater. An alternative approach to the characterization of metal speciation in environmental samples is via metal titrations of the ubiquitous organic ligands in natural samples. Of the many analytical techniques trialed, the copper(II) ISE has gained widespread use in studies of the stability constants and copper complexing capacities of natural organic ligands in swamps, lagoons, rivers, seawater, lakes, ponds, soil extracts, peats, synthetic seawater media, etc. [52 74]. Clearly, the copper(II) ISE has demonstrated its robustness and broad utility in copper speciation studies of natural organic ligands in a wide range of environmental samples, and this method offers an attractive alternative to competitive ligand / adsorptive stripping voltammetric techniques [23, 24, 104, 105]. Notably, there have been an array of studies into the copper(II) complexing capacities of ubiquitous organic ligands in natural waters using ISE potentiometric titration methods. For example, various authors [55, 58, 63, 66, 68] demonstrated that it is possible to determine the copper(II) speciation of natural waters using the data obtained in potentiometric titrations, and the stability constants for model ligands (e.g., carbohydrates, EDTA, glycine, alanine, tiron, etc.) together with natural organic ligands are commensurate with the literature data highlighting the validity of this approach. Most impressive are reports into comparable copper(II) ISE titration and spectrometric studies of the copper(II) complexing capacity of natural waters [52, 54, 57, 64, 71]. In

3. Other Trace Metals in the Environment


Section 2 demonstrated clearly that the vast majority of environmental applications with ISEs has concentrated on the analysis of free and total copper, along with copper complexing capacities in natural samples. Notwithstanding, these same analytical approaches are also applicable to other metal ISE systems, and there has been a limited

Electroanalysis 19, 2007, No. 19-20, 1987 2001 www.electroanalysis.wiley-vch.de  2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1994 amount of research on metal speciation in natural waters using crystalline membrane and glassy membrane iron(III) [75, 76, 106 111], cadmium(II) [77 80, 112 114] and mercury(II) [81 83, 115, 116] ISEs, as well as polymeric ISEs for chromium(VI) [88, 89] and lead(II) [6, 7].

R. De Marco et al.

3.1. Iron(III) in Seawater Iron is an important limiting trace metal nutrient in the environment [104, 105, 117], as it limits the growth of phytoplankton and biomass production in certain oceanic zones. As iron finds its way predominantly into oceans by atmospheric dust deposition and by upwelling of water, it is present at very low levels in isolated areas such as The Southern Ocean near Antarctica, and this was exemplified by an iron fertilization event that led to a phytoplankton bloom in The Southern Ocean, as observed using satellite imaging of the fertilization zone [118, 119]. Clearly, a robust iron(III) ISE that permits a monitoring of free iron(III) in seawater would be of immense interest to environmental scientists and analytical chemists. Using an extension of the approach for lowering the detection limit of the copper(II) ISE based on the positive influence of hydrodynamic flow on the ISE [12], De Marco et al. [75] developed a chalcogenide glass iron(III) ISE CFA method for the determination of free iron(III) in acidified and UV photooxidized Southern Ocean or open ocean seawater at pH 3.86 comprising a total iron(III) concentration of 5 1010 M, as determined by GFAAS. Notably, the CFA ISE method yielded a log(aFe3) 11.1 0.15 units for 10 repetitive injections of seawater against calibration data obtained using saline iron(III)-citrate buffer solutions, and this outcome correlated brilliantly with the iron(III) speciation calculated result of log(aFe3) 10.8. The results of this study demonstrated that the CFA Fe(III) ISE method is extremely useful in measurements of free iron(III) in UV photooxidized seawater. In two important papers by De Marco and co-workers [76, 106], the response of the iron(III) chalcogenide glass ISE was investigated in raw and UV photooxidized open ocean seawater together with a collection of saline calibration buffers comprising a variety of iron(III) binding ligands, and the concomitant response mechanism of the iron(III) ISE in seawater media was explored using EIS and XPS of the electrode kinetics and surface chemistry of the sensor in seawater media. This research demonstrated that the iron(III) ISE is capable of providing a near collinear response for the electrode in salicylate and citrate buffers over a concentration range of 1023 to 101 M aFe3; however, EDTA revealed an offset in ISE response, although the behavior was still Nernstian [76]. Furthermore, this ligand interference is also manifested by natural seawater ligands, and it was shown that the ISE can be used to provide meaningful response data in organic free, UV photooxidized seawater yielding a free iron(III) level commensurate with the well-known inorganic speciation of iron(III) in seawater [76]. Most significantly, the EIS and XPS mech-

anistic study [106] showed that the observed electrode slope of 30 mV/decade for the iron(III) ISE in saline media is consistent with a reversible dual response process involving both electron transfer and ion-exchange that is compliant with its concomitant Nernst equation, and that the iron(III) RDE is also compliant with the Levich equation for the diffusion controlled response in seawater showing that it is also possible to facilitate the dispersal of electrode released iron out of the Nernstian diffusion layer by using hydrodynamic flow in either the RDE, FIA or CFA modes of analysis. By analogy to the response of the copper(II) ISE in seawater, which also illustrated a compliance to the Levich equation for an RDE [12, 33], it is clearly possible to lower the electrode release of iron from the iron chalcogenide ISE to 1010 M or less permitting the analysis of open ocean samples comprising sub-nanomolar levels of analyte, as has been demonstrated in the CFA analysis of iron(III) in Southern Ocean seawater [75]. A further mechanistic study of the iron(III) chalcogenide glass ISE using secondary ion mass spectrometry (SIMS), EIS and XPS in various media [108, 109] suggested that chloride, hydroxide, nitrate, etc. in seawater together with ubiquitous organic ligands do not pose a serious problem for the electroanalysis of Fe3 in seawater, but it is necessary to condition the electrode in seawater prior to analysis, while regular recalibration, reconditioning (preferably overnight) and frequent polishing is required, if the ISEs response in seawater is to remain internally consistent with its response in iron(III) calibration standards. In recent research, De Marco and co-workers [110, 111] have developed a novel synchrotron radiation-grazing incidence X-ray diffraction (SR-GIXRD)/potentiometry/ EIS capability to monitor in-situ the so-called modified surface layer (MSL) of the iron(III) chalcogenide glass ISE in artificial and real seawater, and it was shown that the surface crystalline phases of this sensor (i.e., metal selenides) are attacked by chloride and hydroxide in artificial seawater, but are protected by the natural organic ligands in seawater. The EIS/SR-GIXRD data [110] illustrated that a destruction of the MSL in artificial seawater is due to a complete removal of all surface crystalline phases, and that raw seawater comprising natural organic ligands together with a mimetic seawater ligand system containing ethylenediamine, salicylic acid and EDTA is capable of protecting the MSL of the iron(III) ISE against this oxidative dissolution process. Unpublished data has shown that this mimetic seawater ligand calibration buffer provides meaningful aFe3 data in seawater, and this paper will be published soon. In conclusion, the previous research on the response of the iron(III) ISE in seawater has shown that the adoption of either an RDE, FIA or CFA approach in tandem with a mimetic seawater ligand calibration system can be used in the reliable electroanalysis of free iron(III) in seawater, and this may provide a valuable research tool for analytical chemists interested in studying this important trace metal nutrient in the environment.

Electroanalysis 19, 2007, No. 19-20, 1987 2001 www.electroanalysis.wiley-vch.de  2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Ion-Selective Electrode Potentiometry in Environmental Analysis

1995 such as artificial neural networks (ANN), principal component analysis (PCA) and partial least squares regression (PLS) to analyze waste incinerator exhausts in chemical plants. In summary, the crystalline membrane cadmium(II) ISE possesses a comparable utility to the copper(II) ISE, and may be used in long-term surveying of water quality in the aquatic environment.

3.2. Cadmium(II) in Natural Waters Cadmium toxicity is of great concern to environmental scientists especially in rapidly industrialized regions like East Asia, noting that monsoonal rains can wash cadmium onto flooded farming lands. Despite the environmental importance of cadmium, there have been very few studies of cadmium in the environment using a cadmium(II) ISE. Several authors used the cadmium(II) ISE in studies of cadmium(II) speciation in environmental samples. For example, Simoes et al. [112] studied the speciation of cadmium in artificial and real seawater using a crystalline membrane cadmium(II) ISE and ASV, and it was found that both methods gave comparable stability constants for the binding of cadmium by ubiquitous organic ligands. Similarly, Lee et al. [113] studied the cadmium binding affinity to soil and commercial humic acids using ultrafiltration (UF) coupled with GFAAS in combination with the cadmium(II) ISE, and identified two ligands, with the ISE and UF/ GFAAS techniques yielding comparable formation constants for cadmium(II)-organic complexes. Last but not least, Fish and Brassard [114] demonstrated that natural organic matter (NOM) in sediment samples induced fouling of the cadmium(II) ISE and a concomitant drift in the ISE potential as a function of soaking in samples. These authors demonstrated that a dialysis membrane is capable of eliminating the diffusion of NOM to the electrode surface, thereby obviating the usual long-term ageing of the sensor in real samples. Trojanowicz et al. [77] developed an FIA method for the analysis of free cadmium in natural waters, and used this method in the analysis of free cadmium in lake and river samples. Excellent agreement between GFAAS and ISE levels for free cadmium(II) was observed in these environmental samples, and the use of masking agents for lead(II), copper(II) and iron(III) can eliminate these chemical interferences in FIA assays of free cadmium(II). Although Trojanowicz et al. [77] only demonstrated the applicability of their ISE-FIA technique in contaminated estuarine samples, it is highly likely that the influence of hydrodynamic flow in FIA will permit a diminution in the level of electrode released cadmium to sub-nanomolar levels, as is the case with the copper(II) ISE [12], thereby enabling a use of the cadmium(II) ISE in the electroanalysis of open ocean seawater samples comprising subnanomolar levels of total cadmium. In an important cadmium(II) speciation study of seawater, Sunda [78] determined free cadmium(II) in raw seawater using a Chelex 100 ion-exchange method in conjunction with the cadmium(II) ISE technique, and both methods yielded a ratio of free-to-total cadmium of 101.5, while Sherman et al. [79] used the cadmium(II) ISE in measurements of free cadmium(II) in pond water and sediments, noting that this master variable correlated with the toxicity data (i.e., LC50) for Fathead Minnows. Last, Mortensen et al. [80] developed an array electrode employing a variety of chalcogenide glass ISEs, and used this sensor in conjunction with pattern recognition techniques

3.3. Mercury(II) in Seawater Mercury is a highly toxic mineralogical pollutant that poses a serious risk to the environment, as evidenced by the Minamata Bay incident in Japan where an entire community was poisoned as a result of contaminated seafood. Notwithstanding, there has been a limited amount of research into the development of an ISE sensor technique for the monitoring of mercury(II) in the environment. Shatkin et al. [81] developed a HgS/Ag2S composite membrane that gave a near-perfect Nernstian response to Hg2 in saline mercury(II) buffers in the range 1015 to 102 M aHg2. Similarly, Yin et al. [82] studied the interaction of Hg2 with soil derived humic acid, and found that calculated equilibrium and ISE determined free mercury(II) levels were within the experimental uncertainty, and De Marco and Shackleton [83] along with De Marco et al. [115] demonstrated that the mercury(II) chalcogenide glass ISE was able to provide a Nernstian response over a dynamic range of 20 orders of magnitude and that a seawater passivation interference effect may be circumvented by using a standard addition analysis technique and/or CFA analysis. Last but not least, De Marco et al. [116] used a variety of cutting-edge materials and surface characterization techniques to elucidate the mechanistic chemistry of the mercury(II) chalcogenide glass ISE in seawater media, and showed that the ubiquitous organic ligands of natural waters are able to prevent detrimental silver chloride fouling of the sensor through the peptization of precipitated silver chloride, noting that this desirable effect was also observed with the copper(II) electrode [28]. Regrettably, as open ocean seawater and other uncontaminated natural samples comprise picomolar amounts of dissolved mercury [10], it is unlikely that a mercury(II) ISE will possess sufficient analytical sensitivity to permit the reliable electroanalysis of mercury in uncontaminated environmental samples, but the mercury(II) ISE should suffice for contaminated coastal, lake, river, estuarine, seawater, etc. samples. Accordingly, the electroanalysis of mercury in uncontaminated samples using a mercury(II) ISE necessitates the adoption of a preconcentration technique such as the Chelex 100 method developed by Zolotov et al. [40] in the analysis of trace amounts of copper(II) in various uncontaminated environmental samples.

Electroanalysis 19, 2007, No. 19-20, 1987 2001 www.electroanalysis.wiley-vch.de  2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1996 3.4. Lead(II) and Chromium(VI) An optimized polymeric lead(II) ISE was employed in a comparative inductively coupled plasma/mass spectrometry (ICP/MS) and ISE analysis of total lead in natural water samples [6], and demonstrated that the ISE method is at least as accurate as ICP/MS down to lead(II) levels of approximately 0.5 mg L1. This excellent analytical sensitivity makes the sensor ideal for the analysis of lead(II) in environmental samples, and Slaveykova et al. [7] have used the ISE to monitor free lead(II) levels as function of the toxicological response of an aquatic algae in the presence of Suwannee River fulvic acid, noting a clear correlation between lead(II) uptake by aquatic algae and ISE determined levels of free lead(II). Last but not least, Choi and Moon [88, 89] have developed a liquid supported HCrO4 ISE for the analysis of chromium(VI) in waste waters. The authors also used this new chromium(VI) sensor in continuous flow monitoring, which not only provided scope for automation of the analyses, but also helped to control the electrode release of analyte into the ISEs Nernst diffusion layer, thereby enabling trace analyses in environmental samples. It is clearly evident that all of the metal ISEs surveyed in this section have very similar response attributes to the solid-state copper(II) ISE that has been used extensively and successfully in the analysis of natural waters, and these sensors are therefore expected to fulfill an important role in environmental studies of these trace metals in the environment.

R. De Marco et al.

4. Fluoride in the Environment


Fluoride is an environmentally important analyte, as it is utilized extensively in the minerals processing industry, as well as in drinking water, and finds its way into the environment. Its toxicological effects such as dental fluorosis, chronic toxicosis in animals, and its influence on nutrient cycling due to reaction with environmental nutrients such as organic carbon, aluminum and iron, impact dramatically on the ecosystem [86]. Accordingly, this is an analyte that has attracted attention in the analysis of environmental samples using ISEs. In an important study by Low and Bloom [86], the fluoride ISE was used to measure the Aeolian deposition of fluoridebased compounds in the Tamar Valley of Tasmania Australia surrounding its nearby aluminum smelter. Most significantly, it was found that, within a 3 km radius of the smelter where the observed fluoride fluxes are very high (viz., sometimes as high as 12.568 mg m2 day1), plant damage is clearly observable, and these high levels are at least two orders of magnitude higher than the normal background level. This study identified a need to lower the fluoride emission levels of operational aluminum smelters, and it was highlighted that contemporary practices in aluminum smelting would make a substantially lower emission possible.

In an important environmental survey of fluoride in Polish natural waters, drinking waters and the urine of kindergarten children living in the vicinity of the natural waters sampled [87], it was shown that the fluoride ISE is able to provide accurate data on the levels of fluoride in the urine of infants, thereby providing an indicator for the impact of fluoride on young children who are most sensitive to fluorosis. Similarly, Bhagavan and Raghu [120] used a fluoride ISE to study the fluoride content of bore waters, along with the blood and urine samples of villagers in the Anantapur District of India. Significantly, it was found that high fluoride bearing bore waters in upstream check dams led to expectedly high blood and urine fluoride levels of the villagers, especially in the 5 11 year-old males. Although the surveyed young males did not show any detectable signs of dental or skeletal fluorosis, the authors urged the need for further medical studies with these subjects. Furthermore, the authors recommended that the inhabitants of this area engage in a higher dietary intake of calcium and tamarind to combat this potentially difficult environmental problem. Pierre et al. [121] conducted a study into the urinary excretion of fluoride from aluminum industry workers exposed to common pot-room pollutants (i.e., HF, NaF, AlF3, Al2O3, etc.) using a fluoride ISE, and the data enabled the establishment of a kinetic model to account for the excretion of fluoride and aluminum compounds in humans. Furthermore, it was noted that a subject expressed their maximum in excretion rate several hours after a shift with a minimum rate achieved at the end of a shift. Consequently, an appropriate occupational hygiene sampling regime must take account of this important characteristic, if reliable exposure data is to be obtained. Rix and co-workers [84, 85, 122] developed a standard addition method for the analysis of fluoride in natural waters, and adapted this approach to a microprocessor controlled instrument [85, 122] capable of automated analyses. These researchers applied their technique to a variety of seawater samples (viz., Port Lonsdale Victoria, Corio Bay Geelong, Murray River Swan Hill, Albert Park Lake and Hepburn Springs Victoria), and found that the sensitivity of the ISE was sufficient to detect the ambient levels of fluoride and that the adjoining super-phosphate manufacturing industry in Geelong had contaminated the shallow regions of Corio Bay with high concentrations of fluoride. Clearly, an ISE approach to environmental analysis is also adaptable to anion analyses, if a robust and reliable sensor is available for the analytical task. Obviously, the single crystal lanthanum fluoride ISE possesses the requisite physical, chemical as well as analytical robustness and ruggedness to make it useful in environmental analysis.

5. Miscellaneous Analytes
Despite the suitability of ISEs to field analyses, and a widespread appeal for using these devices in early detection/

Electroanalysis 19, 2007, No. 19-20, 1987 2001 www.electroanalysis.wiley-vch.de  2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Ion-Selective Electrode Potentiometry in Environmental Analysis

1997 a combined nitrate/pH titrimetric system was developed for the analysis of anoxic activated sludge [124]; a commercial ammonia ISE method was developed for the analysis of seawater [125]; a solid-state polyprrole ammonium ISE was employed in FIA for the analysis of a variety of natural waters [126] ; ISEs have been adapted to continuous on-line monitoring of nitrate and ammonium in waste water, sewerage effluent and Danube River water [127]; ISEs have been utilized in the monitoring of chloride in natural waters [128 130]; calcium, magnesium and water hardness in natural waters have been assayed using ISEs [99, 100, 131, 132]; a tetraphenylborate ISE has been used in back titrations of sulfate in seawater in the presence of excess 2aminoperimindium ion [101]; ISEs have been developed to monitor chloride, nitrate and sulfate arising from aerial deposition onto aluminium and zinc metallic substrates as a means of studying the corrosion impact of polluted urban air exposure on electronic equipment [102]; a robust solid-state chloride ISE was employed in the monitoring of chloride in concrete as a tool for monitoring the corrosion tendency of steel in reinforced concrete structures [103]; nonionic surfactants that are emerging environmental pollutants have been studied in rivers using ISE FIA [133]; copper, ammonium and nitrate ISEs were used to analyze reservoir water [134]; a potassium ISE was used in the monitoring of irrigation water [135] and soils [136]. The aforementioned bulleted list of applications in environmental science leaves little to the imagination about the vast array of opportunities that exist for ISEs in environmental monitoring of pollutants, nutrients, harmful and corrosive reagents, emerging pollutants, etc. Notwithstanding, there is a series of recent papers that highlight the many innovative approaches that are possible with ISEs in environmental science. First, Kounaves [137] described the development of an ISE electrode array that may be used with a robotic chemical laboratory for the analysis of calcium, potassium, sodium and chloride as a means of detecting bacterial growth on Mars. Next, Dzyadevych et al. [138] described an enzyme sensitized ISFET pH sensor system for the detection of acetylcholinesterase as an early warning system for detecting organophosphorus pesticides in the environment, and the results showed that this new sensor is capable of providing accurate analytical data that follows the biotoxicity response of a luminescent bacterium. Last, Lei et al.s [139] excellent review on microbial biosensors described an ISE approach in which pH, NH4, Cl, etc. or gas sensors (pCO2 and pNH3) have been sensitized by a microbial biofilm so they may be used to target environmentally important analytes such as organophosphate fertilizers, penicillin, tryptophan, urea, trichloroethylene, etc.

monitoring devices, there are very few environmental studies outside of the metal and fluoride assays that have been described in depth in the previous sections. In this section, the authors will present a brief overview of miscellaneous ISE applications in environmental science. Given the extreme toxicity of cyanide and recent environmental disasters such as poisoning of fish in the Danube River, there have been several attempts to develop a crystalline membrane cyanide ISE for the electroanalysis of natural waters. Hefter and Longmore [90] used a cyanide ISE in continuous monitoring of cyanide in seawater, and demonstrated that the well documented chloride electrode interference may be corrected when the chloride selectivity coefficient and concentration are known. Alternatively, Neshkova and co-workers [91, 92] explored a new generation of electrodeposited thin film silver chalcogenide cyanide ISEs as FIA detectors for the analysis of natural waters, noting that these electrodes possess sufficient sensitivity for the analysis of environmental samples and that the electrodeposition method used in the preparation of the sensor lends itself to in-situ electrode regeneration; an important attribute when these electrodes are used in longterm monitoring in field instrumentation providing a possible source of electrode fouling during continuous electrode exposure to complex media such as seawater. In essence, the novel approach of Neshkova and co-workers [91, 92] is a promising method that is expected to gain widespread use in future environmental research. A very interesting recent paper by Metilda et al. [93] on the use of an ion imprinted polymer ISE documents a novel uranyl ISE incorporating a uranyl-organic complex trapped within a polymer framework that is utilized in a standard PVC membrane electrode. In this work, the authors showed that this new ISE has a detection limit of 2 108 M, which is suitable for the analysis of uranyl ions in seawater, and a comparative analysis of seawater gave uranyl ion concentrations that are as accurate as those determined using neutron activation analysis (NAA). Moreover, this is an excellent example of timely and important ISE research, as the world is striving for energy solutions to air pollution and greenhouse gas emission problems through an intensified utilization of nuclear energy, and the global mining activities for uranium-based compounds are poised to increase. There is a large list of papers describing the use of ISE devices in the analysis of cations and anions in a variety of environmental samples. Herewith is a list of ISE applications on miscellaneous analytes: ISEs have been used in the monitoring of cyanide, nitrite, fluoride and ammonia in the Houston Ship channel [94] and the Wadi El Raiyan Lakes in the Egyptian desert [95]; an ISFET device was developed for on-line monitoring of potassium and ammonium pollution in environmental samples [96, 123]; a cobalt-wire ISE has been used in the analysis of phosphate in waste waters [97];

Electroanalysis 19, 2007, No. 19-20, 1987 2001 www.electroanalysis.wiley-vch.de  2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1998

R. De Marco et al.

6. Future Directions and Recommendations


Like all electrochemical sensors, ISEs are ideally suited to use in electronic tongues in which an array of carefully selected ISEs with appropriate cross-sensitivities may be used to establish a sensor response surface that can be interrogated using standard pattern recognition techniques such as artificial neural networks, principal component analysis, partial least squares regression etc. Krantz-Rlcker et al. [140] presented an excellent review of the work undertaken using chalcogenide ISE arrays for the analysis of drinking waters, and this approach is clearly transferable to the analysis of natural waters in the environment. Portable battery-operated microprocessor controlled electrochemical instrumentation has been available for many years [141], and this approach is clearly amenable to environmental analysis in the field, either in in-situ using submersible CTD instrumentation (viz., Camusso et al.s ISE study of copper in Lake Orta [26]) or in on-site monitoring programs. Without doubt, a critical issue with the adoption of ISEs in environmental analysis is a matching of their analytical sensitivity with the immense challenges of trace analyte detection and the potential of electrode fouling in complex natural samples containing a plethora of potential interfering species. Previous research has illustrated the desirable effects of hydrodynamic flow in RDE, FIA and CFA ISE analytical devices on both the detection limit and adsorption phenomena in complex environmental samples [12, 29, 33, 142]. Accordingly, it is strongly recommended that environmental analyses with ISEs are undertaken in a flow analysis mode using techniques such as FIA or CFA. Nowadays, it is well-known that polymeric membrane ISEs have excellent analytical attributes, as compared to many analytical techniques, and their superior selectivity and low limits of detection make them ideally suited to environmental analysis [143]. Unfortunately, conventional polymeric ISEs do not possess the physical and chemical robustness of their crystalline membrane ISE counterparts, and this becomes problematic during environmental analysis, particularly in field studies. Nevertheless, it is presently possible to make physically and chemically robust solidcontact ISEs on metallic substrates using a toughened methyl methacrylate/decyl methacrylate copolymer membrane, and this approach has been used with a range of solidcontact polymeric ISEs (e.g., silver, lead, calcium, potassium and iodide) providing nanomolar response with good response times and excellent reproducibility [144]. The authors strongly recommend this class of polymeric ISE for use in the analysis of trace analytes in the environment. An exciting analytical approach with polymeric ISEs in environmental analysis is the concept of switchtrodes [145]. With this method, two polymer membrane ISEs are programmed to give kinetically controlled detection limits so as to produce a peak-shaped differential signal between the electrode pair when the activity of the sample is resonant with the intermediate activity needed for switchtrode response. As the ISEs are in a characteristic super-Nernstian

response domain under these conditions, a switchtrode expresses a chemical amplification, which is expected to be useful in the environmental analysis of trace analytes. Belli and Zirino established that the standard addition analysis of metals in natural waters comprising ubiquitous organic ligands is problematic, as a variation in the complexing ability of the sample changes at different levels of added analyte, and this gives rise to an anomalous super-Nernstian response in environmental monitoring [19]. As the majority of ISEs are prone to matrix interference effects (e.g., chloride adsorption, ligand adsorption, aging phenomena, etc.), an inability to conduct standard addition calibrations makes it extremely difficult to undertake reliable analyses in environmental samples. Notwithstanding, Malon et al. [146] reported an exciting approach with polymeric membrane ISEs in which the ISE response may be interrogated through a variation in the composition of the filling solution, and it is possible to use this approach in a so-called backside calibration of the ISE while it bathes in an unperturbed sample. Notably, these authors utilized the new backside calibration methodology in the analysis of lead(II) in Budapest, Warsaw and Zrich tap waters, as well as Zrich Sihl River water, and the analytical data demonstrated that the new method is at least as accurate as standard addition potentiometry. Clearly, this method is expected to provide the environmental chemistry community with a robust and rugged ISE method for the analysis of metals in natural waters. The holy grail for environmental analysis would entail the long-term deployment of ISEs and their associated instrumentation at field sites, and real-time and remote monitoring of their response characteristics using telemetry and computer technology (see Figure 2 and reference [147]). Unfortunately, this approach is fraught with danger as the ISE devices are likely to become fouled by biological films, and the sample matrix itself is likely to foul the sensor surface rendering its response unrepresentative of the calibration data set unless steps are taken to regularly clean and recalibrate the electrodes. Nevertheless, it is possible that ISEs for toxic trace metals such as copper, cadmium, mercury, lead, etc. will be resistant to biofouling events since the membrane dissolution-based release of the analyte will display an antifouling effect, and the desirable influence of hydrodynamic flow in RDE, FIA and CFA devices may eliminate this chemical-based fouling of the membrane. In any event, this approach has its merits, and it deserves the attention of environmental scientists interested in real-time remote monitoring of the water quality of natural samples. Most importantly, the long-term deployment of an ISE in a complex environmental sample such as seawater is expected to alter its response characteristics through an adulteration of the membrane surface, thereby changing the ISE response characteristics against calibration standards, as well as degrading its stability and reproducibility in the sample. As a worst-case scenario, the ISE surface may become poisoned and/or passivated during long-term exposure to the sample matrix (i.e., chloride, hydroxide, organic ligands, etc.), thereby yielding a totally non-functional ISE. Accordingly, these factors will conspire during

Electroanalysis 19, 2007, No. 19-20, 1987 2001 www.electroanalysis.wiley-vch.de  2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Ion-Selective Electrode Potentiometry in Environmental Analysis

1999
[3] R. Jasinski, I. Trachtenberg, D. Andrychuk, Anal. Chem. 1974, 46, 364. [4] A. Zirino, C. Clavell, P. F. Seligman, Mar. Chem. 1983, 12, 25. [5] T. Sokalski, A. Ceresa, T. Zwickl, E. Pretsch, J. Am. Chem. Soc. 1997, 119, 11347. [6] A. Ceresa, E. Bakker, B. Hattendorf, D. Gunther, E. Pretsch, Anal. Chem. 2001, 73, 343. [7] V. I. Slaveykova, K. J. Wilkinson, A. Ceresa, E. Pretsch, Environ. Sci. Technol. 2003, 37, 1114. [8] Z. Szigeti, I. Bitter, K. Toth, C. Latkoczy, D. J. Fliegel, D. Gunther, E. Pretsch, Anal. Chim. Acta 2005, 532, 129. [9] J. Buffle, R. S. Altman, M. Filella, A. Tessier, Geochim. Cosmochim. Acta 1990, 54, 1535. [10] R. J. Motekaitis, A. E. Martell, Mar. Chem. 1987, 21, 101. [11] R. De Marco, D. J. Mackey, A. Zirino, Electroanalysis 1997, 9, 330. [12] A. Zirino, R. De Marco, I. Rivera, B. Pejcic, Electroanalysis 2002, 14, 493. [13] E. Bakker, P. Buhlmann, E. Pretsch, Chem. Rev. 1997, 97, 3083. [14] P. Buhlmann, E. Pretsch, E. Bakker, Chem. Rev. 1998, 98, 1593. [15] E. Bakker, D. Diamond, A. Lewenstam, E. Pretsch, Anal. Chim. Acta 1999, 393, 11. [16] R. L. Solsky, Anal. Chem. 1990, 62, 21R. [17] J. Gulens, Ion-Sel. Electrode Rev. 1981, 2, 117. [18] J. Gulens, Ion-Sel. Electrode Rev. 1987, 9, 127. [19] S. L. Belli, A. Zirino, Anal. Chem. 1993, 65, 2583. [20] W. G. Sunda, P. A. Gillespie, J. Mar. Res. 1979, 37, 761. [21] W. Sunda, R. L. Guillard, J. Mar. Res. 1976, 34, 511. [22] W. G. Sunda, A. M. Lewis, Limnol. Oceanogr. 1978, 23, 870. [23] K. H. Coale, K. W. Bruland, Limnol. Oceanogr. 1988, 33, 1084. [24] K. W. Bruland, E. L. Rue, J. R. Donat, S. A. Skrabal, J. W. Moffett, Anal. Chim. Acta 2000, 405, 99. [25] J. Barica, J. Fish. Res. Board Can. 1978, 35, 141. [26] M. Camusso, G. Tartari, A. Zirino, Environ. Sci. Technol. 1991, 25, 678. [27] R. De Marco, Anal. Chem. 1994, 66, 3202. [28] R. De Marco, Mar. Chem. 1996, 55, 389. [29] R. De Marco, R. Eriksen, A. Zirino, Anal. Chem. 1998, 70, 4683. [30] M. T. Neshkova, Anal. Chim. Acta 1993, 273, 255. [31] D. J. Mackey, R. De Marco, R. Eriksen, Croat. Chem. Acta. 1997, 70, 207. [32] A. Zirino, S. L. Belli, D. A. van der Weele, Electroanalysis 1998, 10, 423. [33] A. Zirino, D. A. VanderWeele, S. L. Belli, R. De Marco, D. J. Mackey, Mar. Chem. 1998, 61, 173. [34] R. Stella, M. T. Ganzerli-Valentini, Anal. Chem. 1979, 51, 2148. [35] S. E. Cabaniss, M. S. Shuman, Anal. Chem. 1986, 58, 398. [36] T. R. Holm, Chem. Speciation Bioavailability 1990, 2, 63. [37] T. R. Holm, C. D. Curtiss, III., ACS Symp. Ser. 1990, 416, 508. [38] T. F. Rozan, G. Benoit, H. Marsh, Y. Chin, Environ. Sci. Technol. 1999, 33, 1766. [39] R. Ghode, R. Sarin, Orient. J. Chem. 1996, 12, 125. [40] Y. A. Zolotov, L. K. Shpigun, I. Y. Kolotyrkina, E. A. Novikov, O. V. Bazanova, Anal. Chim. Acta 1987, 200, 21. [41] R. S. Eriksen, D. J. Mackey, P. Alexander, R. De Marco, X. D. Wang, J. Environ. Monitoring 1999, 1, 483. [42] I. Rivera-Duarte, A. Zirino, Environ. Sci. Technol. 2004, 38, 3139. [43] R. S. Eriksen, D. J. Mackey, R. van Dam, B. Nowak, Mar. Chem. 2001, 74, 99.

Fig. 2. A schematic diagram showing a possible real-time remote sensing system for use with ISEs.

long-term deployment of ISEs in the field to provide highly unreliable ISE analytical data. Under these conditions, it is essential to regularly recalibrate and rejuvenate the electrodes, as required, and this high level of maintenance precludes the use of ISEs in remote sensing applications, especially involving the long-term deployment of ISE sensors. Notwithstanding, short-term in-situ field studies using properly conditioned and recently calibrated ISEs are realizable, as evidenced by the investigations of Camusso et al. [26] on the stratification of copper(II) in Lake Orta Italy using a copper(II) ISE fitted to a CTD device. It is the authors honest opinion that field monitoring with ISEs will be achievable if environmental scientists undertake on-site or shipboard determinations of analytes, as illustrated in the previous and seminal papers by Zirino and co-workers [4, 33], using an FIA or CFA method that is known to ameliorate the problems of electrode drift, electrode fouling, electrode passivation, electrode dissolution and electrode carry-over arising from the cleansing effect of hydrodynamic flow on the surface of the ISE [11]. Accordingly, it is strongly recommended that field research in environmental science is undertaken using this conservative and realizable strategy.

7. Acknowledgements
The Australian Research Council (ARC) and Australian Institute of Nuclear Science and Engineering (AINSE) are acknowledged for financial support.

8. References
[1] A. Zirino, P. F. Seligman, Mar. Chem. 1981, 10, 249. [2] M. J. Smith, S. E. Manahan, Anal. Chem. 1973, 45, 836.

Electroanalysis 19, 2007, No. 19-20, 1987 2001 www.electroanalysis.wiley-vch.de  2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

2000
[44] I. Rivera-Duarte, G. Rosen, D. Lapota, D. B. Chadwick, L. Kear-Padilla, A. Zirino, Environ. Sci. Technol. 2005, 39, 1542. [45] R. J. Erickson, D. A. Benoit, V. R. Mattson, H. P. Nelson, Jr., E. N. Leonard, Environ. Toxicol. Chem. 1996, 15, 181. [46] O. Berson, M. E. Lidstrom, Environ. Sci. Technol. 1996, 30, 802. [47] H. Ma, S. D. Kim, D. K. Cha, H. E. Allen, Environ. Toxicol. Chem. 1999, 18, 828. [48] H. R. Daly, M. J. Jones, B. T. Hart, I. C. Campbell, Environ. Toxicol. Chem. 1990, 9, 1013. [49] J. K. Fraser, C. A. Butler, M. H. Timperley, C. W. Evans, Environ. Toxicol. Chem. 2000, 19, 1397. [50] C. M. G. van den Berg, Mar. Chem. 2000, 71, 331. [51] D. J. Mackey, A. Zirino, Anal. Chim. Acta 1994, 284, 635. [52] J. R. Tuschall, Jr., P. L. Brezonik, Anal. Chim. Acta 1983, 149, 47. [53] A. S. Joksic, L. Dutra, S. A. Katz, J. L. McCulley, Am. Lab. 2002, 34, 26. [54] D. K. Banerjee, E. P. Jagadeesh, Environ. Sci. Res. 1991, 42, 341. [55] R. Stella, M. T. Ganzerli-Valentini, Pergamon Ser. Environ. Sci. 1980, 3, 581. [56] D. A. Roman, L. Rivera, Bol. Soc. Chil. Quim 1995, 40, 55. [57] J. A. Sweileh, D. Lucyk, B. Kratochvil, F. F. Cantwell, Anal. Chem. 1987, 59, 586. [58] J. W. Haas, Jr., Mar. Chem. 1986, 19, 299. [59] E. M. Logan, I. D. Pulford, G. T. Cook, A. B. Mackenzie, Eur. J. Soil Sci. 1997, 48, 685. [60] J. P. Giesy, J. J. Alberts, D. W. Evans, Environ. Toxicol. Chem. 1986, 5, 139. [61] G. K. Brown, S. E. Cabaniss, P. MacCarthy, J. A. Leenheer, Anal. Chim. Acta 1999, 402, 183. [62] T. Midorikawa, E. Tanoue, Y. Sugimura, Anal. Chem. 1990, 62, 1737. [63] N. Radic, Anal. Lett. 1979, 12, 115. [64] R. E. Truitt, J. H. Weber, Anal. Chem. 1981, 53, 337. [65] W. Fish, F. M. M. Morel, Can. J. Fish. Aquat. Sci. 1983, 40, 1270. [66] R. Ghode, R. Sarin, Orient. J. Chem. 1996, 12, 237. [67] D. R. Turner, M. S. Varney, M. Whitfield, R. F. C. Mantoura, J. P. Riley, Geochim. Cosmochim. Acta 1986, 50, 289. [68] N. Parthasarathy, J. Buffle, Anal. Chim. Acta 1994, 284, 649. [69] J. Buffle, F. Greter, W. Haerdi, Anal. Chem. 1977, 49, 216. [70] J. Buffle, P. Deladoey, F. L. Greter, W. Haerdi, Anal. Chim. Acta 1980, 116, 255. [71] W. Fish, F. M. M. Morel, Can. J. Chem. 1985, 63, 1185. [72] D. A. Roman, L. Rivera, Mar. Chem. 1992, 38, 165. [73] H. M. V. M. Soares, M. T. S. D. Vasconcelos, Anal. Chim. Acta 1994, 293, 261. [74] H. Xue, W. G. Sunda, Environ. Sci. Technol. 1997, 31, 1902. [75] R. De Marco, B. Pejcic, X. D. Wang, Lab. Robot. Autom. 1999, 11, 284. [76] R. De Marco, D. J. Mackey, Mar. Chem. 2000, 68, 283. [77] M. Trojanowicz, P. W. Alexander, D. Brynn Hibbert, Anal. Chim. Acta 1998, 370, 267. [78] W. G. Sunda, Mar. Chem. 1984, 14, 365. [79] R. E. Sherman, S. P. Gloss, L. W. Lion, Water Res. 1987, 21, 317. [80] J. Mortensen, A. Legin, A. Ipatov, A. Rudnitskaya, Y. Vlasov, K. Hjuler, Anal. Chim. Acta 2000, 403, 273. [81] J. A. Shatkin, H. S. Brown, S. Licht, Anal. Chem. 1995, 67, 1147. [82] Y. Yin, Allen, H. E., C. P. Huang, P. F. Sanders, Anal. Chim. Acta 1997, 341, 73. [83] R. De Marco, J. Shackleton, Talanta 1999, 49, 385.

R. De Marco et al. [84] C. J. Rix, A. M. Bond, J. D. Smith, Anal. Chem. 1976, 48, 1236. [85] K. A. Phillips, C. J. Rix, Anal. Chem. 1981, 53, 2141. [86] P. S. Low, H. Bloom, Atmos. Environ. 1988, 22, 2049. [87] P. Konieczka, B. Zygmunt, J. Namie snik, Bull. Environ. Contam. Toxicol. 2000, 64, 794. [88] Y. W. Choi, S. H. Moon, Environ. Monit. Assess. 2001, 70, 167. [89] Y. W. Choi, S. H. Moon, Environ. Monit. Assess. 2004, 92, 163. [90] G. T. Hefter, A. R. Longmore, Int. J. Environ. Anal. Chem. 1984, 16, 315. [91] M. T. Neshkova, E. M. Pancheva, V. Pashova, Sens. Actuators, B 2006, B119, 625. [92] A. R. Surleva, V. D. Nikolova, M. T. Neshkova, Anal. Chim. Acta 2007, 583, 174. [93] P. Metilda, K. Prasad, R. Kala, J. M. Gladis, T. P. Rao, G. R. K. Naidu, Anal. Chim. Acta 2007, 582, 147. [94] M. A. Saleh, E. Ewane, B. L. Wilson, Chemosphere 1999, 39, 2357. [95] M. A. Saleh, E. Ewane, J. Jones, B. L. Wilson, Ecotoxicol. Environ. Saf. 2000, 45, 310. [96] M. Daniel, M. Janicki, W. Wroblewski, A. Dybko, Z. Brzozka, A. Napieralski, Water Sci. Technol. 2004, 50, 115. [97] R. De Marco, B. Pejcic, Z. Chen, Analyst 1998, 123, 1635. [98] H. A. Zamani, M. R. Ganjali, M. Adib, J. Braz. Chem. Soc. 2007, 188, 215. [99] M. E. Thompson, Science 1966, 153, 866. [100] T. Imato, K. Ishii, N. Ishibashi, Anal. Sci. 1992, 8, 631. [101] T. Masadome, Y. Asano, Talanta 1999, 48, 669. [102] G. B. Munier, L. A. Psota, B. T. Reagor, B. Russiello, J. D. Sinclair, J. Electrochem. Soc. 1980, 127, 265. [103] M. F. Montemor, J. H. Alves, A. M. Simoes, J. C. S. Fernandes, Z. Lourenco, A. J. S. Costa, A. J. Appleton, M. G. S. Ferreira, Cem. Concr. Compos. 2006, 28, 233. [104] M. Gledhill, C. M. G. van den Berg, Mar. Chem. 1994, 47, 41. [105] E. L. Rue, K. W. Bruland, Mar. Chem. 1995, 50, 117. [106] R. De Marco, B. Pejcic, Anal. Chem. 2000, 72, 669. [107] R. De Marco, D. J. Mackay, Mar. Chem. 2000, 71, 333. [108] B. Pejcic, R. De Marco, K. Prince, Surf. Interface Anal. 2002, 33, 748. [109] B. Pejcic, R. De Marco, K. Prince, Surf. Interface Anal. 2002, 33, 759. [110] R. De Marco, Z. T. Jiang, J. Martizano, A. Lowe, B. Pejcic, A. van Riessen, Electrochim. Acta. 2006, 51, 5920. [111] R. De Marco, Z. T. Jiang, B. Pejcic, A. van Riessen, Electrochim. Acta 2006, 51, 4886. [112] S. M. de L. Simoes, M. C. T. A. Vaz, J. J. R. Frausto da Silva, Talanta 1981, 28, 237. [113] M. H. Lee, S. Y. Choi, K. H. Chung, H. Moon, Bull. Korean Chem. Soc. 1993, 14, 726. [114] S. J. Fish, P. Brassard, Talanta 1997, 44, 939. [115] R. De Marco, B. Pejcic, S. Cook, Lab. Robot. Autom. 2000, 12, 194. [116] R. De Marco, B. Pejcic, K. Prince, A. van Riessen, Analyst 2003, 128, 742. [117] F. J. Millero, W. Yao, J. Aicher, Mar. Chem. 1995, 50, 21. [118] E. R. Abraham, C. S. Law, P. W. Boyd, S. J. Lavender, M. T. Maldonado, A. R. Bowie, Nature 2000, 407, 727. [119] A. J. Watson, D. C. E. Bakker, A. J. Ridgwell, P. W. Boyd, C. S. Law, Nature 2000, 407, 730. [120] S. V. B. K. Bhagavan, V. Raghu, Environ. Geochem. Health. 2005, 27, 97. [121] F. Pierre, F. Baruthio, F. Diebold, P. Biette, Occup. Environ. Med. 1995, 52, 396. [122] K. A. Phillips, C. J. Rix, Anal. Chim. Acta 1985, 169, 263.

Electroanalysis 19, 2007, No. 19-20, 1987 2001 www.electroanalysis.wiley-vch.de  2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Ion-Selective Electrode Potentiometry in Environmental Analysis [123] J. J. Wang, P. L. Bishop, Environ. Technol. 2005, 26, 381. [124] B. Petersen, K. Gernaey, P. A. Vanrolleghem, Water Sci. Technol. 2002, 45, 181. [125] A. G. A. Merks, Neth. J. Sea Res. 1975, 9, 371. [126] D. P. Quan, C. X. Quang, L. T. Duan, P. H. Viet, Environ. Monit. Assess. 2001, 70, 153. [127] S. Winkler, L. Rieger, E. Saracevic, A. Pressl, G. Gruber, Water Sci. Technol. 2004, 50, 105. [128] E. Wang, S. Kamata, Anal. Chim. Acta 1992, 261, 399. [129] P. J. Watkins, Electroanalysis 2001, 13, 1215. [130] M. Trojanowicz, W. Matuszewski, Anal. Chim. Acta 1983, 151, 77. [131] S. Xie, H. Ji, A. Wu, G. Chen, Chin. J. Oceanol. Limnol. 1997, 15, 32. [132] E. D. Gravley, T. J. Manning, Florida Scientist 1995, 58, 320. [133] S. Martinez-Barrachine, M. del Valle, L. Matia, R. Prats, J. Alonso, Anal. Chim. Acta 2002, 454, 217. [134] B. S. Smolyakov, A. M. Nemirovskii, V. V. Kokovkin, L. A. Pavlyuk, D. F. Plekhanov, J. Anal. Chem. 1995, 50, 994. [135] T. Gieling, H. H. van den Vlekkert, Adv. Space Res. 1996, 18, 135.

2001
[136] S. M. Brouder, M. Thom, V. I. Adamchuck, M. T. Morgan, Commun. Soil Sci. Plant Anal. 2003, 34, 2699. [137] S. P. Kounaves, Chem. Phys. Chem. 2003, 4, 162. [138] S. V. Dzyadevych, A. P. Soldatkin, V. N. Arkhypova, A. V. Elskaya, J. Chovelon, C. A. Georgiou, C. Martelet, N. Jaffrezic-Renault, Sens. Actuators, B 2005, 105, 81. [139] Y. Lei, W. Chen, A. Mulchandani, Anal. Chim. Acta 2006, 568, 200. [140] C. Krantz-Rulcker, M. Stenberg, F. Winquist, I. Lundstrom, Anal. Chim. Acta 2001, 426, 217. [141] A. M. Bond, H. A. Hudson, S. N. Tan, F. L. Walter, Trends Anal. Chem. 1988, 7, 159. [142] Q. Ye, M. E. Meyerhoff, Anal. Chem. 2001, 73, 332. [143] E. Bakker, E. Pretsch, Trends Anal. Chem. 2001, 20, 11. [144] K. Y. Chumbimuni-Torres, N. Rubinova, A. Radu, L. T. Kubota, E. Bakker, Anal. Chem. 2006, 78, 1318. [145] T. Vigassy, W. E. Morf, M. Badertscher, A. Ceresa, N. F. del Rooij, E. Pretsch, Sens. Actuators, B 2001, B76, 477. [146] A. Malon, E. Bakker, E. Pretsch, Anal. Chem. 2007, 79, 632. [147] H. B. Glasgow, J. M. Burkholder, R. R. Reed, A. J. Lewitus, J. E. Kleinman, J. Exp. Mar. Biol. Ecol. 2004, 300, 409.

Electroanalysis 19, 2007, No. 19-20, 1987 2001 www.electroanalysis.wiley-vch.de  2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Anda mungkin juga menyukai