Open access peer-reviewed chapter

Recent Advances in Voltammetric Sensing

Written By

Harsha Devnani and Chetna Sharma

Submitted: 21 July 2022 Reviewed: 13 October 2022 Published: 02 December 2022

DOI: 10.5772/intechopen.108595

From the Edited Volume

Frontiers in Voltammetry

Edited by Shashanka Rajendrachari, Kiran Kenchappa Somashekharappa, Sharath Peramenahalli Chikkegouda and Shamanth Vasanth

Chapter metrics overview

455 Chapter Downloads

View Full Metrics

Abstract

The practical day to day life is largely affected by the products that we use, the air that we breathe, the soil that is used to grow crops, the water we drink and use for various household chores or industrial purposes. The purity analysis of these products or estimation of useful inorganic and organic analytes is of utmost importance for avoiding health and environment risk. Everyone wants to be aware that what they are eating or applying on their skin is safe for them. A diabetic patient needs to monitor their blood sugar levels constantly. The air, water and soil quality needs constant monitoring to avoid health hazards. Not just this, chemical analysis is crucial as a crime investigation technique to identify suspects. Fuel quality and storage needs to be tested for eliminating unwanted losses. The electrochemical techniques are inherently fast, selective and sensitive and some systems are portable as well which is a boon for on-site monitoring. Voltammetric techniques like cyclic voltammetry, stripping voltammetry, impedance spectroscopy, amperometry and other techniques provide specific information of the analytes to be tested. This chapter will highlight the voltammetric techniques used for different types of analyte sensing and the advances that have taken place recently related to voltammetric sensing.

Keywords

  • voltammetry
  • sensing
  • electrochemistry
  • cyclic voltammetry
  • stripping voltammetry

1. Introduction

The electrochemical sensor is a broad integrated area encompassing physical aspects, analytical science, material science, electronic fabrication, biochemistry along with statistical analysis. This chapter restricts to the electroanalytical methods involving analytical science and electrochemistry which lay the foundation for voltammetric sensing. Electroanalytical measurements are based on the measurement of potential (potentiometry), current (voltammetry) and the amount of electricity consumed or the matter transformed during electrolysis (coulometry). These techniques are inherently fast, sensitive, selective and offer low detection and quantification limits [1]. In contrast to other analytical techniques like chromatography, ICP-OES, MS and others, electrochemical analysis does not require heavy instrumentation and tedious sample preparation. In fact, miniaturized sensors are portable and are thus handy for on-site monitoring wherever required.

Each of the above mentioned electroanalytical measurement methods involve a specifically designed electrochemical cell. Although potentiometric sensors are more lucrative for on-site operations, voltammetric sensors are more sensitive and fast. Recently, there has been a surge in researching new underlying principles for electrochemical sensing. As a result, sensors are now being developed taking advantage of changes in ionic conductivity, resistivity and impedance [2]. An altogether new concept based on imaging of local electrochemical current utilizing the optical signal from the electrode surface (surface plasmon resonance) has been reported by Shan et al. in 2012 [3]. This derives foundation from the fact that the current density can be determined from the local surface plasmon resonance signal.

Electroanalytical techniques like linear sweep voltammetry (LSV), cyclic voltammetry (CV), pulse voltammetry (PV), stripping voltammetry (SV), chronoamperometry provide an in-depth qualitative and quantitative information for an electro-active species deeming them to be a potential alternative to more commonly used spectrometric or chromatographic techniques. The understanding of an electrode process helps to explore mechanisms for in vivo studies, which is beneficial in analyzing how a drug works when administered to a human body. It is being utilized to speciate and determine ultra trace analytes in complex clinical and industrial samples. Sensor development is a thrust area in the field of chemistry, biology and environmental sciences [4]. A chemical sensor is defined as, “a small device that as the result of a chemical interaction or process between the analyte and the sensor device, transforms chemical or biochemical information of a quantitative or qualitative type into an analytically useful signal” [5]. Chemical sensors find wide applications in industry, critical care, quality control, process controls, pollutant monitoring, safety alarms, clinical diagnostics, food adulteration testing, in agriculture and forensics.

With the advent of electrochemistry, initially the voltammetric methods posed a lot of difficulties as a result of which they could not be much benefited from at the time. Gradually, with significant advancements over a period of time, voltammetric sensing soon became popular as an attractive analytical method. All the voltammetric techniques fundamentally involve the application of an electrode potential (E) and measuring the corresponding current (i) in the respective electrochemical cell. In most voltammetric techniques, the applied potential (E) is varied and the resulting current (i) is measured over a period of time (t). The applied potential relates to the change in electroactive species which can undergo a redox reaction and thus there is change in its concentration, which can be related to the corresponding current of the cell [6].

The voltammetric sensors have lured the researchers as they offer high sensitivity for inorganic as well as organic species along with a wide linearity range using a variety of electrolytes over a wide range of temperatures. Moreover, it has a fast response time and offers simultaneous determination of analytes at times. Voltammetric studies extend a deep insight into the kinetics and mechanism of the electrochemical process under study, making them attractive for sensing applications. Stripping analysis by far is the most sensitive voltammetric technique which boasts the advantages that it does not require derivatization and is also less sensitive to matrix changes in comparison to other analytical techniques [7, 8]. It is being widely used for analysis of heavy metals present in trace amounts in environmental samples. Cathodic stripping voltammetry (CSV) and anodic stripping voltammetry (ASV) have also been used for achieving sensitive determination of various analytes like nucleotides, nucleosides and nucleobases [9].

Advertisement

2. Voltammetric sensors

The voltammetric sensors initially were developed from mercury, carbon materials and inert metals as working electrodes. The use of mercury has long been forbidden due to its toxic nature and the noble metals demand high cost. The mercury electrode is almost out of the picture due to its difficult handling and limited anodic potential range. The carbon electrodes are still in use in a number of forms viz. glassy carbon, diamond, fullerene, graphite, nanotubes, graphene and graphene oxide [2]. In the past few years, carbon paste electrodes have gained momentum as working electrodes as they can be screen printed for production at mass scale. Another set of electrodes is the screen printed electrodes which are used as a robust miniaturized version and are viable for commercial technology. Most of the electrochemical sensors in the market, like the glucose sensor, employ screen printed electrodes as the working electrode [10, 11]. There is immense potential for electrochemical sensors to be established in markets other than that of glucose. The recent COVID-19 pandemic has demanded an urgent need for fast diagnostic processes which can help prevent spread of infectious diseases and provide timely diagnosis of neurodegenerative disease which is a crucial factor [12, 13, 14]. The voltammetric sensors hold promise for such diagnostic methods as they are fast, simple in operation, offer real time analysis and can be mass-produced for portable use as they can be miniaturized compared to other existing diagnostic methods [15, 16, 17].

The research is now targeted in the direction of new electrode materials which could offer comparable sensitivity and cost economical. Carbon and gold based electrodes are more popular owing to their biocompatibility, stability and good electron transfer kinetics. The bare electrodes often lack sensitivity and selectivity needed for the analysis due to poor charge transfer at the electrode surface. This requires modification of electrode surfaces to enhance charge transfer kinetics and avoid interferences [18, 19]. In this light, the nanomaterials are paving the way for smart electrochemical sensors as they provide enhanced electrochemical surface area with large surface area to volume ratio allowing improved interfacial kinetics exhibiting electrocatalytic activity for sensitive and selective determination of analytes [18, 19, 20]. The synergistic effect of nanomaterials is also being actively pursued by combining two or more nanomaterials to form a composite which would derive advantages from all individual components based on the sensing need [7, 8, 21, 22, 23, 24]. Although nanomaterials have proved to be beneficial, there are certain challenges yet to overcome like handling sophisticated instruments and the stability of nanoparticles utilizing stabilizers or capping agents in certain cases to avoid agglomeration which will disrupt electrode-electrolyte interfacial kinetics [25, 26].

There are mainly three major hurdles encountered in the development of electrochemical sensors: the attainment of a low limit of detection (LOD); limiting the interaction of unwanted interfering species; maintaining sensor stability and achieving reproducibility in complex real matrices [27]. The LOD indicates the lowest concentration/quantity that could be detected for an analyte and is a major criterion for a sensor performance as many times we are dealing with analytes that are present at trace levels in real samples. With the advancement of nanomaterial modified surfaces, picomolar level of detection has been achieved [28, 29, 30]. These modified surfaces still pose the challenge of stability and reproducibility. Moreover, the sensor needs to be validated for real samples otherwise it does not hold importance in the market. The test of an electrochemical sensor as a diagnostic tool is only validated if it is stable and functional in a real matrix [31, 32, 33]. Real matrices involve interferences that hinder electrode performance. In the area of medical applications, this is being dealt with by exploring advanced materials to improve electrode surfaces. While the passive methods involve the use of polymers to create a hydrophilic and non-charged layer to limit protein absorption, active methods aim to develop stronger shear forces than the adhesion forces of the bound interferents on the surface [34, 35]. Recently, the use of sol–gel materials along with ceramics and nanomaterials for electrode surfaces have proven to enhance the stability of the sensor [36, 37]. The research has progressed over the years to overcome these challenges and establish effective diagnostic tools in the market for various applications. This chapter highlights the advancement that has taken place in recent years in the development of voltammetric sensors for a variety of applications.

2.1 Voltammetric food sensors

Food is prone to contamination of heavy metals, pesticides and adulterants. Metals come from a variety of natural and anthropogenic sources. However, man-made activities such as urbanization and agriculture can increase their levels, pollute water and soil, and damage the environment. Moreover, pesticides and food additives when added beyond their permissible limits are making way through the food chain, affecting human health invariably [38]. Ingestion of dangerous metal-rich vegetables, crustaceans and other foods can damage stable cells, alter metabolism, lead to carcinogenic mutations and toxicological effects on human organs [39]. The consequences of these elements on human health have inspired extensive analysis of major contaminants in food samples. The well-established spectrometric techniques viz. atomic emission or absorption spectroscopy [40], mass spectroscopy (MS) and inductively coupled plasma-MS (ICP-MS) have been widely used for this analysis in the past decade. Locatelli and Melucci (2013) detected levels of copper, mercury, lead, zinc and cadmium in several vegetables using a rectangular ASV approach and a standard addition method. The study was performed using two electrodes, gold electrode for detection of mercury and the other made of mercury to determine lead, copper, zinc and cadmium. For these studies, lettuce, spinach, and tomatoes were differentiated as washed and unwashed categories. The two groups were mineralized after treatment with an acid mixture in different amounts [41, 42]. The investigation showed that the values for lead and cadmium exceed the regulatory limits of the European Council. Table 1 lists the range of voltammetric techniques used recently to sense a variety of heavy metals with nanomolar detection limits being achieved at robust electrode systems.

SampleTargetTechniqueElectrodeLODReference
Tomato, potato, and mangoCu, Cd, and PbCV and DPSVPGE modified with MWCNTs1.03 μg l−1 for
Cd, 2.12 μg l−1
for Cu, and 1.62 μg l−1 for Pb
[43]
Fish liverCuSWVHanging mercury drop electrode0.6 nmol l−1 and 0.9 nmol l−1[44]
FishZn, Cd, Cu, Pb, and HgZn, Cd, Cu, Pb, and HgHanging mercury drop electrode1–10 ng l−1,[45]
RiceCdDPASVAu nanoparticle modified CPE1.94 nmol l−1[46]
Chicken, duck, and turkeyRoxarsone (ROX)DPVSPCE modified with lanthanum molybdates12.4 nmol l−1[47]
Apple, potato and lemonCdSWVPGE/bimetal oxide nanoparticles/graphene oxide1.85 ng l−1[48]
Beans and cornsPb, Cd, and CuLSASVGCE0.1 mol l−1[49]
MushroomZn, CuDPASVAmalgam electrode10.3 mg kg−1; 72.9 mg kg−1[50]
Canned tuna fishHgDPVGCE/Cu-MOF nanocubes0.0633 nmol l−1[51]

Table 1.

Literature survey of metal analysis in food samples.

LSASV: Linear Sweep Anodic Stripping Voltammetry, DPV: Differential Pulse Voltammetry, SWV: Square Wave Voltammetry, DPASV: Differential Pulse Anodic Stripping Voltammetry, GCE: Glassy Carbon Electrode, SPCE: Screen Printed Carbon Electrode, PGE: Pencil Graphite Electrode, MWCNT: Multiwalled Carbon Nanotubes, MOF: Metal Organic Framework.

Zabihpour et al. (2020) have reported a robust vanillin (flavoring agent) electrochemical sensor derived from carbon paste electrode (CPE) substrate enriched with NiFe2O4 nanoparticles and 1-hexyl-3-methylimidazolium chloride (1H3MCl). NiFe2O4 nanoparticles were formed using a co-precipitation approach and the characterization findings pointed to a spherical NiFe2O4 nanoparticle with a diameter of 22 nm [52]. The CV technique was used to measure the vanillin oxidation peaks at potentials of +690 and + 650 mV on the surface of CPE and NiFe2O4/1H3MCl/CPE. The DPV analysis at NiFe2O4/1H3MCl/CPE showed a strong electrocatalytic capacity towards the electrooxidation of vanillin and revealed two distinct oxidation signals at potentials of 640 and 1050 mV. NiFe2O4/1H3MCl/CPE was thus successfully used as an analytical sensor to determine vanillin and tryptophan levels in coffee, milk, chocolate and cookie samples [52, 53].

It is also required to estimate nutraceuticals or phytopotentials of certain plants for therapeutic purposes. Voltammetry provides a gateway to such analysis with high sensitivity and selectivity. Zheng et al. (2022) have reviewed the evaluation of antioxidant activity using electrochemical sensors. The development of research in this area has been outlined with its initial stages in 1999, thereafter gaining momentum in 2010 and has remained so ever since. A total of 758 articles were published during this period. Electrochemical methods were used for the first time mainly for quantitative analysis, as well as other analytical approaches. Subsequently, CV was used to directly measure the electrochemical properties of various antioxidants and evaluate their antioxidant capacity. There were several advantages in this scenario when compared to the conventional DPPH assay [54]. The most direct application scenario for the evaluation of antioxidant capacity is in the food industry as it can have a direct impact on their price and nutritional value. One of the most widespread and produced antioxidants present commonly in food is vitamin C, which is electrochemically active and thus can be detected by electrochemical sensors. The development of an electrochemical sensor for Vitamin C brought a major breakthrough in the field of analytical chemistry [55]. Bounegru and Apetrei (2020) reported a nanomaterial based voltammetric sensor for the qualitative and quantitative determination of caffeic acid using CV (Figure 1). Carbon nanofibre (CNF) and MWCNT modified carbon based SPEs (C-SPE) were utilized to study the electrochemical behavior of caffeic acid in aqueous solution (pH 3.6). The LOD and LOQ values were seen to be in the range of 10−7 – 10−9 M (Figure 1) which indicates good sensitivity and in fact the electrochemical results were also compared to the spectrophotometric data. Also, among the two naomaterials used CNF based sensor proved to be better in terms of sensitivity and performance. The optimized sensor (CNF/C-SPE) was also tested for real samples (Active Detox (Herbagetica), DVR-Stem Glycemo (DVR Pharm) and green tea (Alevia) as depicted in Figure 2 yielding satisfactory results [56].

Figure 1.

(a) Zoomed-in view of the anodic peak zone of the CV registered with CNF/C-SPE immersed in caffeic acid solutions with the concentrations in the 0.1–40 μM range. (b) Linear dependence between the anodic peak current and the concentration of the caffeic acid solution [56].

Figure 2.

CVs of CNF/C-SPE immersed in solutions of (a) active detox, (b) DVR-stem Glycemo and (c) green tea, recorded at scan rates between 0.1 and 1.0 V·s−1 [56].

Qin et al. (2020) reported a TiO2/electro reduced graphene oxide (TiO2/ErGO) nanoparticles based electrochemical sensor for the simultaneous analysis of ponceau 4R and tartrazine. The nanocomposite was prepared by ultrasonically dispersing TiO2 nanoparticles in the grapheme oxide solution followed by an electro-reduction step. The TEM analysis (Figure 3) indicated the uniform distribution of the TiO2 nanoparticles in ErGO nanoflakes confirming the nanocomposite formation which resulted in enhanced adsorptive stripping DPV current response as is expected because of the synergistic effect of the nanomaterials. The two colorants could be detected in a nanomolar range using the respective sensor with selectivity, sensitivity and stability. The sensor was also applied for real sample which was orange juice in this case [57].

Figure 3.

TEM images of TiO2 nanoparticles (A) and TiO2/GO nanocomposites (B); XRD patterns of nanomaterials (C) [57].

2.2 Voltammetric sensing of drug & pharmaceuticals

As pharmaceutical firms evolve to keep pace with expanding population of humans globally, the global pharmaceutical industry reached $1.25 trillion in 2019 and is expected to reach $1.5 trillion by 2023. The implications of this data suggest that the pharmaceutical sales grew by 207.9 percent between 2005 and 2019. As a result, two main problems emerged: many drugs are misused/abused thus pressing the need for researching economical, portable, and effective sensors for monitoring drug overdose or biomedical monitoring; and secondly, tracking or monitoring of pharmaceutical or other contaminants in water sources to maintain human and ecological health is also necessity of today’s world [58]. Based on the unique electrochemical interfaces between the nanomaterials and the analyte at hand, there are several approaches that may be more suitable for the detection of pharmaceuticals [59]. Voltammetry, potentiometry, amperometry, and electrochemical impedance spectroscopy (EIS) are examples of common electrochemical methods. Due to their higher sensitivity than CV, pulse methods such as normal pulse voltammetry (NPV), DPV and SWV are more commonly utilized for electrochemical detection. These pulse approaches significantly increase the signal of interest corresponding to the analytes compared to the capacitive current response [60].

Voltammetry and polarography are the most commonly used electroanalytical methods in pharmaceutical and biological analysis. In the 1930s and 1940s, the first examples of pharmaceutical analysis using polarographic techniques were documented. Most pharmaceutically active chemicals have been found to be electrochemically active [61]. Electrochemistry is a well-established and rapidly growing field with numerous potential applications in the pharmaceutical industry as is evident from Table 2 [56]. The use of solid electrodes in the voltammetric determination of drugs is gaining popularity due to its simple modification providing a large electrochemical active surface area for fast charge transfer kinetics.

SampleTargetTechniqueElectrodeDrugLinear RangeLODReference
Non toxic MaterialHClDPVCPEDexamethosone0.5–1 μg l−14,10 μg l−1[62]
Non toxic MaterialHClDPVCPEPrednisolone0.5–1 μg l−14,10 μg l−1[63]
Non toxic MaterialHClDPVCPEHydrocortisone0.5–1 μg l−14,10 μg l−1[64]
MWCNTHClCVCPETinadozol4.9 μg l−18.3 μg l−1[65]
River WaterBufferCVSPEIbufuran1.6 μg l−14.7 μg l−1[66]
River WaterBufferDPVSPEFlunitrazepam2 μg l−11 μg l−1[67]
Coca colaBufferDPVSPEesomeprazole3.5 μg l−110 μg l−1[68]
Alco-popBufferDPVSPEgemifloxin0.4 μg l−12 μg l−1[69]

Table 2.

Literature survey of electrochemical analysis in drug & pharmaceuticals.

ElectrodeModifierElectrochemical MethodAnalytesLinear RangeDetection LimitReferences
CPE1,4-BBFT/ILSWVIsoproterenol6.0 × 10−8–7.0 × 10−4 M12.0 nM[77]
LIGSalmonella typhimuriumSWVAptamer10 CFU/mL67–6.7 × 105 CFU/mL[78]
CPEFC/CNTDPVN-acetylcysteine1.0–400.0 μM0.6 μM[79]
SPES. pullorum & S. gallinarumCVAntibody1.61 × 101 CFU/mL101–109 CFU/mL[78]
DPV3 CFU/mL10–107 CFU/mL[80]
CPEFC/MWCNTDPVCysteamine0.7–200 μM0.3 μM[81]
folic acid5.0–700 μM2.0 μM[82]
GMES. typhimuriumCVAu2.4 × 102 to 2.4 × 1072.4 × 102 cfu/mL[83]
DPV3.1 × 10−5–3.3 × 10−3 M9.0 × 10−6 M[84]
CPEFC/CNTDPVNorepinephrine0.47–500.0 μM0.21 μM[85]
CPE2CBF/GOSWVHydrochlorothiazide5.0 × 10−8–2.0 × 10−4 M20.0 nM[86]
CPEFM/TiO2 nanoparticleDPVMethyldopa2.0 × 10−7–1.0 × 10−4 M8.0 × 10−8 M[87]
CPEFCD/CNTDPVNorepinephrine0.03–500.0 μΜ22.0 nM[88]
CPE2CBF/CNTSWVN-acetylcysteine5.0 × 10−8–4.0 × 10−4 M2.6 × 10−8 M[88]

Table 3.

Electrochemical biosensor developed for the detection of some analytes.

Various modifiers were used to improve the recorded current intensity and the sensitivity of the electrode for the detection of pharmaceutical traces. Additionally, since the solid electrode is non-toxic, it can be used with minimal safety precautions. The phrase disposable electrode makes electrochemical drug determination easier than it was in the 1940s. This word is associated with solid electrodes, the most common home-made conventional solid electrodes, including carbon paste electrodes, screen-printed electrodes, boron-doped diamond electrodes, and pencil graphite electrodes. Both screen-printed and pencil graphite electrodes are single-use electrodes [70].

Mehmandoust et al. (2021) reported gold/silver core–shell nanoparticles (Au@Ag CSNPs) with conducting polymer poly(3,4-ethylenedioxythiophene) polystyrenesulfonate (PEDOT: PSS) and functionalized multi-carbon nanotubes (F-MWCNTs) on glassy carbon electrodes to create a novel and sensitive voltammetric nanosensor for the first time to monitor trace levels of favipiravir [70]. Under optimized conditions and at a typical working potential of + 1.23 V (vs. Ag/AgCl), Au@Ag CSNPs/PEDOT: PSS/F-MWCNT/GCE revealed linear quantitative ranges from 0.005 to 0.009 and 0.009 to 1.95 μM with a detection limit of 0.046 nM (S/N = 3) with acceptable relative standard deviations (1.1–4.9 percent) for pharmaceuticals, urine samples, and human plasma without any sample pretreatment (1.12–4.93 percent) [71]. Amino acids, biologics, and antiviral drugs had little or no effect, and the sensing system exhibited exceptional reproducibility, repeatability, stability, and reusability. The results showed that this approach has potential applications in the diagnosis of FAV in clinical samples. The increased surface area and the synergistic interaction between the bimetallic core-shell nanoparticles and the entrapped carbon structures by the conducting polymer resulted in high sensing properties [72]. Ziyatdinova and Gimadutdinova (2021) developed a CeO2.Fe2O3 nanoparticles based sensor for the estimation of lipoic acid which is extensively used in medicine as drug therapy. The SEM analysis (Figure 4) indicated the uniform distribution of nanoparticles on the electrode surface resulting in enhanced catalytic oxidation currents, reduced overpotential and improved electron transfer kinetics. The designed optimized sensor was also successfully tested in pharmaceutical dosages for the lipoid acid [73]. Kavieva and Ziyatdinova (2022) subsequently reported a sensor based on SeO2nanopartciles and surfactant (cetylpyridinium bromide; CPB) for the determination of indigo carmine which is a widely used colorant in pharmaceutical industry. The immobilization of SeO2 nanoparticles on the electrode surface was confirmed with SEM analysis presented in Figure 5. The sensor was characterized using CV (Figure 6), chronoamperometry and electrochemical impedance spectroscopy (EIS) dictating reduction in charge transfer resistance at the fabricated sensor accompanied with roughly 4 times higher electroactive surface area [74].

Figure 4.

Surface morphology of bare GCE (a) and CeO2.Fe2O3 nanoparticles modified GCE [73].

Figure 5.

SEM images of bare GCE (a) and SeO2-CPB/GCE (b) [74].

Figure 6.

CVs of 50 μM indigo carmine at SeO2-CPB/GCE at various scan rates in pH 5 phosphate buffer [74].

2.3 Voltammetric biosensors

Since the first documented glucose biosensor more than half a century ago, biosensors have developed at an exponential rate. The use of biosensors in many fields such as diagnostics, environment, healthcare, and pharmaceuticals has resulted in the biosensor business growing into a multi-million dollar market that is expected to thrive in the coming years. In this massive biosensor sector, new advances in biosensors in terms of nanobiosensors should expand the range of such technologies in the market.

Biosensors are generally based on the interaction of target species with the surface of a biomolecular receptor coupled to a transducer to generate a signal. A biosensor system can therefore be categorized based on how the detection was performed, such as electrochemical, mass, or optical. The method of detection can also be used to describe a nanobiosensor; however, the transducer used in such circumstances is usually from the field of nanotechnology, i.e. the use of nanoparticle or nanostructured interfaces. Optical nanoprobes detect various analytes using highly active metal nanoparticles such as gold or silver [75]. Although biosensors include well-established bioanalytical techniques, nanobiosensors have seen significant improvements in their sensitivity, resulting in a significant shift in the research trend from biosensors to the development of nanobiosensors. Table 3 lists the range of electrochemical biosensors reported for analytes of importance in various sectors with their respective detection limit.

Kusior (2022) has reported a very interesting study for the determination of glucose without an enzyme receptor at copper oxide nanomaterial modified electrode wherein the facet exposition of Cu2O nanoparticles has been linked to the current-potential profiles. The nanoparticles were synthesized using a wet chemical approach and different facets were exposed to current analysis for glucose determination (Figures 7 and 8). It is claimed that the electrochemical sensing or biosensing of Cu2O nanoparticles depended on {100} and {111} facets, with former possessing more neutral states and the latter positive state. The study was performed by synthesizing particles in different sizes by addition of surfactants and the electrochemical performance was tested by CV and amperometry [76].

Figure 7.

SEM images of obtained polyhedral Cu2O particles depending on surfactant used [76].

Figure 8.

Schematic representation of Au particle adsorption at the polyhedral surface with SEM image of modified Cu2O grains [76].

2.4 Voltammetric sensors for agriculture purposes

Contamination of pesticides and herbicides in soil, groundwater, rivers, lakes, storm water and air is a major problem. Pesticides are widely used in agriculture around the world and are a key tool for controlling weeds, insects and infections. They were defined as chemicals or mixtures designed to resist, eliminate, prevent or limit the presence or effect of biological organisms capable of causing crop damage [89]. Since 1960, the increased use of pesticides has allowed farmers to significantly increase productivity while avoiding crop losses caused by pests [90]. Herbicides are pesticides that are used to kill or slow the growth of weeds. They were categorized based on their activity (contact or systemic), use (soil, pre-emergent or post-emergent) and mode of action on plant biochemical mechanisms [91, 92]. In addition, their target is classified as non-selective (kills all surrounding plants) or selective (attacks only weeds and leaves the crop alone) [93]. Glyphosate (N-(phosphonomethyl)glycine) was the most commonly used herbicide, followed by atrazine (2-chloro-4-ethylamino-6-isopropylamino-1,3,5-triazine) and 2,4D-dichlorophenoxyacetic acid) [94]. Priscila discussed the detection of herbicides (1990–2018) to obtain certain qualities and increase the detection limit, the matrix can be changed using metals, metal oxides, polymers, clay materials or micro or nanoparticles.

SampleTransducersAnalyteBiosensing ElementsReference
SoilAmperometric2,4-DAcetyl-Cholinesterase[95]
SoilBioluminescenseDiuron, paraquatCyanobacterial[96]
SoilPotentiometricSimazinePeroxidase(Biocatalytic)[97]
SoilImpedance2,4-DAntibody[99]
SoilPotentiometricHydrazideAcetyl-Cholinesterase[101]
Drinking WaterPotentiometricIsoproturonAntibody[100]
Waste WaterPotentiometricSimazinePeroxidase(Biocatalytic)[102]
Orange JuiceAmperometricAtrazineAntibody[98]

Table 4.

Sensors developed for the detection of some herbicides.

Electrochemical sensors are good candidates for environmental monitoring and have been successfully fabricated using various types of electrode materials. The applicability and advantages of voltammetric techniques over traditional methods have been well established. Surface modification of the electrode improves its properties, allowing small amounts of a specific analyte to be detected or quantified; for metals or metal oxide electrodes are now able to detect non-electroactive herbicides. Nanostructures have also been shown to have several advantages and are a promising field of study. There are a significant number of biosensor papers that provide exciting new potential to increase their performance for herbicide detection (Table 4). Most of the work presented used real samples such as tap and lake water, poisoned soil and food hazardous substances (fruits and vegetables), but there is still a need to incorporate these new devices into commercial analysis or industries.

2.5 Electrochemical energy sensors

Fuels are an essential part of today’s civilization with significant economic and ecological importance. The current search for new sustainable alternative fuels to fossil fuels was based on minimizing the emission of pollutants into the atmosphere such as CO2. Biofuels, as they were widely known, were commonly produced from vegetable oils and their composition can vary depending on their origin, resource types and their matrices. Metals and toxins in fuels, regardless of their origin, such as fossil or alternative, remain a problem that needs to be solved. A large number of metals can be used as catalysts in the oxidation processes of fuels when combined with gasoline, either by contamination or adulteration, and the results can damage automotive engine components [100]. Among many types of current fuels, it is customary to examine and assess the metal contamination in their composition, which exceeds the limitations of regulatory authorities to avoid economic and environmental losses associated with the concentration of specific metals in fuel matrices. There are several variables that affect the maximum amount of metal allowed in gasoline. It was determined by the regulatory body, the type of fuel and even the metal involved. Lead was not permitted in gasoline in Europe and manganese is limited to 2 mg/l [88]; however no metals are permitted in gasoline in the United States [101]. There was extensive literature on the evaluation of metals in fuels using a variety of approaches, most of which have a high material and resource analysis. Due to its mobility, high sensitivity and low cost compared to other approaches, electrochemical techniques offer a promising alternative route for these kinds of studies.

Dos Santos and Ferreira discussed in a 2020 publication the analysis of Pb2+ and Cu2+ in microemulsified biodiesel using a boron-doped diamond electrode [90]. This process has been used to produce biodiesel from corn oil, according to the researchers [102]. In comparison to other traditional detection methods, the results obtained were good. The analysis was performed despite the fact that these metals are not controlled in biodiesel. Ferreira and co-workers in 2021 used B-doped diamond electrodes in ethanol solution with ASV to measure Fe3+, Cu2+, Zn2+, Pb2+ and Cd2+. Combining interactions and interferences between species, was investigated [103]. A sample of fuel ethanol was tested in a hydroethanolic environment using this approach after the experimental parameters were tuned, with concentrations of investigated ions below the detection limits for the proposed method (Table 5).

SampleTargetTechniqueElectrode typeLODReference
BiodieselCuDPASVHg film electrode4.69 nmol-1[104]
BiodieselPbDPASVHg film electrode2.91 nmol-1[105]
Bio-Ethanol fuelFeLSSVNafion2.4 μmol-1[106]
Bio-Ethanol fuelCuDPSVCNT170 nmol l-1[107]
BiodieselCaSWVGCE1.6 nmol l−1[108]
BiodieselTinSWASVBi Film electrode0.14 μmol l−1[109]
Bio-Ethanol fuelZnSWASVGDE5 μg l−1[110]
BiofuelCuSWASVVulcan functionalized1.2 nmol l−1[111]

Table 5.

Electrochemical energy sensors for metal analysis.

2.6 Voltammetric environmental sensors

Some benefits are likely to emerge in the existing water monitoring environment when the mode of analysis moves based on sampling techniques to exclusively Conductivity-Temperature-Depth (CTD) probes have been used underwater since the 1980s, there has been a clear shift towards decentralized techniques approaches in a portfolio of known analytical techniques show clear promise for use in An evaluation of submersible devices in terms of their analytical properties, autonomy, miniaturization, and portability [111]. Plant modified carbon paste electrodes have been reported in literature recently for the estimation of heavy metals. One study was done in 2016 wherein bagasse (waste from sugarcane) was used as a biomaterial in a sensor to detect the hazardous metals Pb and Cd. An electrochemical approach was used to investigate the performance of the bagasse-based carbon paste sensor [112]. SEM, FTIR, CV and BET analysis were used to characterize the modifier’s surface which indicated mesoporous pore distribution of bagasse and the specific surface area was calculated to be 89.3 m2/g [100]. The accumulation was carried out in pH 6 acetate buffer and stripping analysis was done in HCl at a scan rate of 50 mV/s. The linearity was observed for lead and cadmium in the concentration ranges of 100–600 gL−1 and 500–1200 gL−1 with detection limits of 10.1 and 170.64 gL−1, respectively for 10 mins of accumulated time.

Electrochemical sensors have proven ideal for this function, particularly in ion detection. Maria Cuartero reported (2021) an investigation of electrochemical sensors that showed promise for measuring ions in seawater such trace elements, nutrition, and carbon species The preceding five years were chosen as the major era for investigation, but older contributions to the area or goods introduced are included when important technical advancements are highlighted. There was a distinct absence of electrochemical sensors used in in-situ applications, which exacerbated when saltwater was considered: only a few examples have been demonstrated under such difficult conditions.

A glassy carbon electrode (GCE) modified by simultaneous electropolymerization of L-lysine (L-lys) and -cyclodextrin (-CD) film (P-CD-L-lys/GCE) was described as a novel electrochemical platform for the detection of pymetrozine. Using CV in 0.1 M H2SO4, the electrochemical activity of P-CD-L-lys/GCE towards pymetrozine was investigated [113]. The sensor’s potential value was shown using LSV to estimate pymetrozine concentration analytically. The linear range is 4.0108 mol/L to 1.0106 mol/L. Pymetrozine’s detection limit was determined to be 1.3108 mol/L (S/N = 3). We were able to acquire satisfactory pymetrozine results in real sample analysis using this approach. The usage of amino acid materials and cyclodextrin in environmental analysis was supported in this study. Chaiyo et al. (2020) reported a simple, low-cost, and highly sensitive voltammetric sensor based on a Nafion, ionic liquid, and graphene composite, modifying a screen-printed carbon electrode for the simultaneous determination of Zn2+, Cd2+, and Pb2+ in drinking water at the nanomolar level [114]. Del Valle and co-workers reported the immobilization of 4-carboxybenzyl-18-crown-6 and 4-carboxybenzo-15-crown-5 on monolayers of aryldiazonium salts anchored on the surface of graphite-epoxy composite electrodes for the simultaneous determination of Cd2+, Pb2+ and Cu2+ in synthetic water samples at to the ppb level (ca. nanomolar) using differential pulse anodic stripping voltammetry (DPASV) [115]. Then, using a comparable electrode modification, Perez-Rafols et al. developed an electronic tongue for the detection of Cd2+, Pb2+, Tl+ and Bi3+ in synthetic water samples [101]. For the detection of trace metals, molecularly imprinted polymers [102, 103, 104] and metal-shaped nanoparticles [116, 117] have also been proposed. Voltammetric analysis of Pb2+ in enriched water has been reported using MoS2/rGO flower composite with ultrathin nanosheets [118], Ni/NiO/MoO3/chitosan 3D foam at the p-n junction interfacial barrier for micromolar level Cu2+ detection [119] and Mn-mediated MoS2nanosheets have been reported recently as a new approach to Pb2+ sensing [120]. A VIP (Voltammetric In-Situ Profiler) for the detection of Cu2+, Pb2+, Cd2+ and Zn2+ using a Hg-based electrode [121] and a kayak equipped with Zn2+voltammetric sensors based on liquid crystal polymer bismuth film [122] are the only two cases reported in the literature that demonstrated in-situ operation in seawater. Single and multiple analyte detection methods have been investigated in the last five years. Maria’s (2021) report includes a number of electrochemical techniques for the detection of trace metals, although demonstrated applications in unspiked saline samples are quite limited (Table 6). For nitrogenous nutrients, two in-situ solutions were effectively used to produce NO3 and NO2 profiles in saltwater.

SensorAnalyteElectrodeLODReference
SWSVTrace Metals Cu2+Hg10 −11 M[123]
SWSVTrace metal Pb2+Hg10−11 M[124]
SWSVTrace metal Cd2+Hg10−11 M[125]
SWSVTrace metal Zn2+Hg10−11 M[126]
SWSVTrace metal Zn2+Bi1 nM[127]
ISENutrients NO3−Iniline acid1 μM[128]
ISMNutrients NO2−Disalination1 μM[129]
ISMTANGC/POT with nonactine1 μM[130]
SWVPhosphateAu,Mo-P1 μM[131]
ISMCarbon speciesGC/CNT/ISM with ionophoresμM, pH = 9[132]
Acid base titrationAlkanilitySolid electrodepH < 4[133]
Amphoteric biosensorNOxBacteria chamber1 μM[134]

Table 6.

Electrochemical environmental sensors for metal analysis.

After seawater treatment, potentiometric detection of NO3 and NO2 at relatively low doses is possible utilizing multiple electrodes combined in a flow cell. Amperometric NOx and NO2 biosensing is possible using various bacterial chambers in an electrode configuration that can be deployed in any water system using an analyzer or submerged device allowing direct contact of the sensors with the water column at micromolar concentration with separate electrodes. There are no demonstrated in-situ uses of electrochemical devices for in-situ detection of ions in saltwater, according to studies. At the laboratory scale, new advancements are ongoing, with in-situ installation compatibility promised but rarely fulfilled. The few sensors that have been integrated in submersibles appear to be in the early stages of commercialization. As a result, meaningful data is only gathered through the inventors’ measurements [135]. Electrochemical sensors have the potential to transform seawater analysis programs, but bringing this vision to fruition would need well-planned phases and targeted research.

2.7 Electrochemical sensing in forensics

Homicide cases usually involve weapons that release many compounds when fired. These compounds are identified in gunshot residue that may be found at a crime scene, on a suspects’ clothing or other items, which then become useful in the investigation of a crime by providing evidence. This gunshot residue (GSR) consists of both inorganic and organic parts that need analysis for a successful investigation. In addition, elements such as antimony, lead, bismuth and copper present in GSR are hazardous to human health and the environment and therefore their determination is necessary [136]. Increasing security needs necessitate the deployment of field-deployable detectors capable of detecting GSR and nitro-aromatic explosive compounds in real time.

Erden et al. (2011) presented the measurement of antimony and lead in gunshot residues using differential pulsed cathodic adsorptive stripping (DPCAS) and square-wave cathodic adsorption stripping voltammetry (SWCASV) [137]. A hanging drop of mercury was used as a working electrode for GSR samples obtained during test firings at the Police Criminal Laboratory from the shooters’ hands. The results demonstrated that both the DPCAdSV and SWCAdSV approaches could be used to successfully determine metals in GSR, both qualitatively and quantitatively. Vuki et al. (2012) then reported the simultaneous detection of Pb and Sb. along with other organic components of GSR at propellant on a glassy carbon electrode using CV and cyclic-SWV [138]. The thin film Hg GCE was used for Ba analysis because it requires detection of the presence of mercury [139].

Ceto et al. (2012) reported Zn, Pb, Cu, and nickel amalgams on bare and modified SPE by SWV in samples obtained straight from shooters at a nearby shooting range. The fascinating element is that they were able to validate several handling situations including secondary contact with the GSR, loading the handgun, and finally discharging. The subject that fired the firearm was identified other than the ability to distinguish the time of contact with the GSR with respect to different subjects. The use of microelectrodes for lead analysis in GSR was demonstrated by Salles et al. (2012) using the SWV technique [140]. Microelectrodes have substantial benefits for analysis, such as the capacity to deal with small-volume samples, downsizing of equipment, and the removal of mixing in the pre-concentration process. Their study suggested the possibility of linking lead originating from the GSR to discrimination between different types of weapons and ammunition. Wearable, completely stiff finger sensors for quick on-site voltammetric assessment of GSR and explosive surface residues were described by Bandodkar et al. (2013). To build new forensic fingers, they screen-printed the three-electrode arrangement onto a nitrile finger pad and overlay another finger pad with an ionogel electrolyte layer [141]. The novel integrated sampling/detection system uses “microparticle voltammetry” (VMP) to transmit minute quantities of surface-confined analytes directly onto a fingertip-based electrode array. Voltammetric measurements of sample residues are performed by bringing the working electrode into direct contact with a second finger bed coated in ionogel electrolyte (worn on the thumb), therefore completing the solid-state electrochemical cell [142]. The sampling and screening process took less than four minutes and results in GSR and explosives having distinct voltammetric signatures. The use of a solid, flexible ionogel electrolyte eliminates any liquid handling, reducing leakage, portability, and contamination concerns. The fingertip detection gadget demonstrated great specificity for detecting GSR and nitroaromatic chemical residues. It can tolerate constant mechanical stress without losing its attractiveness, a low-cost way for conducting on-site crime scene investigations in a range of forensic settings [143]. Hashim et al. 2016 reported gold modified screen printed electrode for Cu (II) analysis in GSR using CV (Figure 9). It was also compared with ICP-OES analysis with 94% accuracy [144].

Figure 9.

CV at gold modified SPE for various Cu(II) concentration with inset showing calibration plot [144].

The research area was long forgotten, but is now re-emerging as a result of the sensitivity and adaptability that this approach provides. Ott et al. published a paper in 2020 on the analysis of Pb, Sb, and Cu by SWASV employing bare SPE. Two samples were taken in this investigation, one from the hands of volunteers with no direct touch and the other from the hands of those who had recently discharged a weapon. They examined data from 395 genuine shooter samples and 350 background samples, making it the biggest GSR research yet done [145]. Using a simple, quick, and sensitive voltammetric assay, they were able to identify all metals as well as organic bullet residues at the same time. It was detailed how to make a biochar-modified carbon paste electrode. Oliveira et al. used DPAdSV in 2021 to assess lead ions in gunshot residue and hair coloring samples. GSR samples were acquired using an IMBEL 9 mm pistol and 100% cotton fabric, with no hair color specified. To eliminate interference from other metals contained in gunpowder components such as Sb (III), Cu (II), Cr (III), and Fe (II), this work employed the conventional addition technique to measure lead in the acquired samples [146].

Bessa et al. (2021) presented a different approach to lead detection in Lucilia cuprina, a necrophagous insect that functioned as a GSR biomarker due to GSR consumption, As a result, an intriguing lead in forensic investigations is provided. Their survival rates were investigated, as well as the influence of lead on their embryonic growth [147]. As a working electrode, a bismuth film was deposited in an SPCE using the SAWASV process. The results showed that lead could be detected on larval samples even in the presence of interference such as Cu, Ba, and Cd. The effect of lead on mortality and larval development time was also successfully investigated [148, 149]. The revival of this area has opened up a whole new avenue for research leading to breakthrough in criminal investigations.

Advertisement

3. Conclusion

Voltammetry holds the key to future diagnostic tools for fast, real-time and sensitive analysis. It has found use in almost every spectrum of environment, medical care, agriculture, forensic, investigation, food and cosmetic industries for a range of analytes to be determined selectively or simultaneously depending on the requirement. The resurgence of the interest in electrochemical methods for the lucrative metal analysis required for various applications is derived from the advantages it caters to, like sensitivity and versatile application. Nanomaterials and polymers are the advanced materials being explored for their synergistic effect to modify electrode surfaces to overcome the challenges that stand in the way. The research is targeted to develop fast, stable, reproducible, miniaturized real-time sensors for use in various fields.

Advertisement

Acknowledgments

The authors are thankful to their respective organizations for the facilities provided to the authors to carry out their work. The authors would also like to acknowledge MDPI publisher for their open access journals which allowed us to reproduce content in form of figures.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Majeed S, Naqvi STR, Najamul Haq M, Ashiq MN. Electroanalytical techniques in biosciences: Conductometry, coulometry, voltammetry, and electrochemical sensors. In: Egbuna C, Patrick-Iwanyanwu KC, Shah MA, Ifemeje JC, Rasul A, editors. Analytical Techniques in Biosciences. Elsevier, Academic Press; 2021. pp. 157-178. DOI: 10.1016/B978-0-12-822654-4.00004-X
  2. 2. Power AC, Morrin A. Electroanalytical sensor technology. In: Khalid MAA, editor. Electrochemistry. London: Intech Open; 2013. DOI: 10.5772/51480
  3. 3. Chen W, Cai S, Ren Q-Q, Wen W, Zhao Y-D. Recent advances in electrochemical sensing for hydrogen peroxide: A review. The Analyst. 2012;137:49-58. DOI: 10.1039/C1AN15738H
  4. 4. Zhang X, Ju H, Wang J. Electrochemical Sensors, Biosensors and their Biomedical Applications. Elsevier, Academic press; 2008. pp. 307-581. DOI: 10.1016/B978-0-12-373738-0.X5001-6
  5. 5. Stetter JR, Penrose WR, Yao S. Sensors, chemical sensors, electrochemical sensors, and ECS. Journal of The Electrochemical Society. 2003;150:S11. DOI: 10.1149/1.1539051
  6. 6. Laina AL. Voltammetric Sensors for the Determination of Pharmaceuticals. Kochi: Cochin Institute of Science and Technology; 2013
  7. 7. Devnani H, Ansari S, Satsangee SP, Jain R. ZrO2-graphene-chitosan modified carbon paste for sensitive and selective determination of dopamine. Materials Today Chemistry. 2017;4:17-25. DOI: 10.1016/j.mtchem.2017.02.004
  8. 8. Ansari S, Ansari MS, Devnani H, Satsangee SP, Jain R. CeO2/g-C3N4nanocomposite: A perspective for electrochemical sensing of antidepressant drugs. Sensors and Actuators B: Chemical. 2018;273:1226-1236. DOI: 10.1016/j.snb.2018.06.036
  9. 9. Fojta M, Jelen F, Hevrek L, Palecek E. Electrochemical stripping techniques in analysis of nucleic acids and their constituents. Current Analytical Chemistry. 2008;13:250-262. DOI: 10.2174/157341108784911415
  10. 10. Morrin A. Inkjet printed electrochemical sensors. In: Korvink JG, Smith PJ, Shin D-Y, editors. Inkjet Based Micromanufacturing. 1st ed. Weinheim: Wiley-VCH; 2012. DOI: 10.1002/9783527647101.index
  11. 11. Newman JD, Turner APF. Home blood glucose biosensors: A commercial perspective. Biosensors Bioelectronics. 2005;20:2435-2453. DOI: 10.1016/j.bios.2004.11.012
  12. 12. Jiang Z, Feng B, Xu J, Qing T, Zhang P, Qing Z. Graphene biosensors for bacterial and viral pathogens. Biosensor & Bioelectronics. 2020;166:112471. DOI: 10.1016/j.bios.2020.112471
  13. 13. Vermisoglou E, Panacek D, Jayaramulu K, Pykal M, Frebort I, Kolar M, et al. Human virus detection with graphene-based materials. Biosensors & Bioelectronics. 2020;166:112436. DOI: 10.1016/j.bios.2020.112436
  14. 14. Kim K, Lee CH, Park CB. Chemical sensing platforms for detecting trace-level Alzheimer's core biomarkers. Chemical Society Reviews. 2020;49:5446-5472. DOI: 10.1039/D0CS00107D
  15. 15. Ligler FS, Gooding JJ. Lighting up biosensors: Now and the decade to come. Analytical Chemistry. 2019;91:8732-8738. DOI: 10.1021/acs.analchem.9b00793
  16. 16. Idilia A-CN, Ploense KL, Csordas AT, Kuwahara M, Kippin TE, et al. Seconds-resolved pharmacokinetic measurements of the chemotherapeutic irinotecan in situ in the living body. Chemical Science. 2019;10:8164-8170. DOI: 10.1039/C9SC01495K
  17. 17. Li H, Dauphin-Ducharme P, Ortega G, Plaxco KW. Calibration-free electrochemical biosensors supporting accurate molecular measurements directly in undiluted whole blood. Journal of American Chemical Society. 2017;139:11207-11213. DOI: 10.1021/jacs.7b05412
  18. 18. Quesada-González D, Merkoçi A. Nanomaterial-based devices for point-of-care diagnostic applications. Chemical Society Reviews. 2018;47:4697-4709. DOI: 10.1039/C7CS00837F
  19. 19. Muniandy S, The SJ, Thong KL, Thiha A, Dinshaw IJ, Lai CW. Carbon nanomaterial-based electrochemical biosensors for foodborne bacterial detection. Critical Reviews in Analytical Chemistry. 2019;49:510-533. DOI: 10.1080/10408347.2018.1561243
  20. 20. Curulli A. Nanomaterials in electrochemical sensing area: Application and challenges in food analysis. Molecules. 2020;25:5759. DOI: 10.3390/molecules25235759
  21. 21. Patel BR, Nouroozifar M, Kerman K. Review-nanocomposite-based sensors for voltammetric detection of hazardous phenolic pollutants in water. Journal of The Electrochemical Society. 2020;167:037568. DOI: 10.1149/1945-7111/ab71fa
  22. 22. Devnani H, Satsangee SP, Jain R. A novel graphene-chitosan-Bi2O3 nanocomposite modified sensor for sensitive and selective determination of a monoamine neurotransmitter epinephrine. Ionics. 2016;22:943-956. DOI: 10.1007/s11581-015-1620-y
  23. 23. Silva LP, Silva TA, Moraes FC, Fatibello-Filho O. A voltammetric sensor based on a carbon black and chitosan-stabilised gold nanoparticle nanocomposite for ketoconazole determination. Analytical Methods. 2021;13:4495-4502. DOI: 10.1039/D1AY01321A
  24. 24. Hassasi S, Hassaninejad-Darzi SK, Vahid A. Production of copper-graphenenanocomposite as a voltammetric sensor for anti-diabetic metformin using response surface methodology. Microchemical Journal. 2022;172:106877. DOI: 10.1016/j.microc.2021.106877
  25. 25. Ismail NS, Hoa LQ, Huong VT, Inoue Y, Yoshikawa H, Saito M, et al. Electrochemiluminescence based enzymatic urea sensor using nanohybrid of isoluminol-gold nanoparticle-graphene oxide nanoribbons. Electroanalysis. 2017;29:938-943. DOI: 10.1002/elan.201600477
  26. 26. Ponnaiah SK, Periakaruppan P, Vellaichamy B. New electrochemical sensor based on a silver-doped iron oxide nanocomposite coupled with polyaniline and its sensing application for picomolar-level detection of uric acid in human blood and urine samples. Journal of Physical Chemistry B. 2018;122:3037-3046. DOI: 10.1021/acs.jpcb.7b11504
  27. 27. Ferrag C, Kerman K. Grand challenges in nanomaterial-based electrochemical sensors. Frontiers in Sensors. 2020;29:583822. DOI: 10.3389/fsens.2020.583822
  28. 28. Alizadeh T, Atashi F, Ganjali MR. Molecularly imprinted polymer nano-sphere/multi-walled carbon nanotube coated glassy carbon electrode as an ultra-sensitive voltammetric sensor for picomolar level determination of RDX. Talanta. 2019;194:415-421. DOI: 10.1016/j.talanta.2018.10.040
  29. 29. Wu Y, Tilley RD, Gooding JJ. Challenges and solutions in developing ultrasensitive biosensors. Journal of American Chemical Society. 2019;141:1162-1170. DOI: 10.1021/jacs.8b09397
  30. 30. Gupta P, Tsai K, Ruhunage CK, Gupta VK, Rahm CE, Jiang D, et al. True picomolar neurotransmitter sensor based on open-ended carbon nanotubes. Analytical Chemistry. 2020;92:8536-8545. DOI: 10.1021/acs.analchem.0c01363
  31. 31. Tseng RC, Chen CC, Hsu SM, Chuang HS. Contact-lens biosensors. Sensors. 2018;18:2651. DOI: 10.3390/s18082651
  32. 32. de Castro LF, de Freitas SV, Duarte LC, de Souza ACJ, Paixão TR, TomazelliColtro WK. Salivary diagnostics on paper microfluidic devices and their use as wearable sensors for glucose monitoring. Analytical Bioanalytical Chemistry. 2019;411:4919-4928. DOI: 10.1007/s00216-019-01788-0
  33. 33. Baghelani M, Abbasi Z, Daneshmand M, Light PE. Non-invasive continuous glucose monitoring system using a chipless printable sensor based on split ring microwave resonators. Science Reports. 2020;10:12980. DOI: 10.1038/s41598-020-69547-1
  34. 34. Lichtenberg JY, Ling Y, Kim S. Non-specific adsorption reduction methods in biosensing. Sensors. 2019;19:2488. DOI: 10.3390/s19112488
  35. 35. Li Y, Xu Y, Fleischer CC, Huang J, Lin R, Yang L, et al. Impact of anti-biofouling surface coatings on the properties of nanomaterials and their biomedical applications. Journal of Material Chemistry B. 2018;6:9-24. DOI: 10.1039/C7TB01695F
  36. 36. Li T, Li Y, Zhang T. Materials, structures, and functions for flexible and stretchable biomimetic sensors. Accounts of Chemical Research. 2019;52:288-296. DOI: 10.1021/acs.accounts.8b00497
  37. 37. Meng LA, Turner PF, Mak WC. Soft and flexible material-based affinity sensors. Biotechnology Advances. 2020;39:107398. DOI: 10.1016/j.biotechadv.2019.05.004
  38. 38. Malitesta C, Di Masi S, Mazzotta E. Electrochemical biosensors to biomimetic sensors based on molecularly imprinted polymers in environmental determination of heavy metals. Analytical chemistry. 2017;10:3389. DOI: 10.3389/fchem.2017.00047
  39. 39. Mahale RS, Shashanka R, Vasanth RSV. Voltammetric determination of various food azo dyes using different modified carbon paste electrodes. Biointerface research in applied. Chemistry. 2022;12:4. DOI: DOI. 10.33263/BRIAC124.45574566
  40. 40. Saidur MR, Aziz ARA, Basirun WJ. Recent advances in DNA based electrochemical biosensors for heavy metal ion detection. Review Biosensor and Bioelectronics. 2017;90:125-139. DOI: 10.1016/j.bios.2016.11.039
  41. 41. Uloma A, Ihugba CO, Nwoko FR, Tony-Njoku AA, Ojiaku LI. Heavy metal determination and health risk assessment of oyster mushroom pleurotus tuber regium (Fr.) singer, collected from selected markets in Imo State. NIGERIAN American. Journal of Environmental Protection. 2018;6:22-27. DOI: 10.12691/env-6-1-4
  42. 42. Covaci E, Darvasi E, Ponta M. Simultaneous determination of Zn, Cd, Pb and Cu in mushroom by differential pulse anodic stripping voltammetry using thesis. Study University Babes-Bolyai University Of Cluj-Napoca. 2017;3:133-144
  43. 43. Kupchyk O. Determining heavy metals in mushroom samples by stripping voltammetry. Food Science and Technology. 2018;12:2. DOI: 10.15673/fst.v12i2.939
  44. 44. Raheem HAE, Hassan RYA, Khaled R, Farghali A, Sherbiny MIEL. New sensing platform of poly (ester-urethane) urea doped with gold nanoparticles for rapid detection of mercury ions in fish tissue. Royal Society of Chemistry. 2021;11:31845-31854. DOI: 10.1039/D1RA03693A
  45. 45. Zhang L, Liao Q, Shao S, Zhang N, Shen Q, Liu C. Heavy metal pollution, fractionation, and potential ecological risks in sediments from Lake Chaohu (eastern China) and the surrounding Rivers. International journal Environmental Research Public Health. 2015;12:14115-14131. DOI: 10.3390/ijerph121114115O
  46. 46. Veerakumar P, Sangili A, Manavalan S, Lin KC. Metal oxide-carbon nanocomposite-modified electrochemical sensors for toxic chemicals. ACS Applied Material & Interfaces. 2021;2021:173-212. DOI: 10.1016/B978-0-12-820727-7.00010-0
  47. 47. Hashemi SE, Payehgadar M, Isaghi Z, Kargar H. Ultra trace level square wave anodic stripping voltammetric sensing of mercury(II) ions in environmental samples using a Schiff base-modified carbon paste electrode. International Journal of Environmental Analytical Chemistry. 2019;99:1148-1163. DOI: 10.1080/03067319.2019.1618455
  48. 48. Singh S, Numan A, Zhan Y, Singh V, Van HT, Nam ND. A novel highly efficient and ultrasensitive electrochemical detection of toxic mercury (II) ions in canned tuna fish and tap water based on a copper metal-organic framework. Journal of Hazardous Materials. 2020;399:123042. DOI: 10.1016/j.jhazmat.2020.123042
  49. 49. Gomes NO, Mendonca CD, Machado SAS, Olivera ONJR, Pereira PAR. Flexible and integrated dual carbon sensor for multiplexed detection of nonylphenol and paroxetine in tap water samples. Analytica Chimica Acta. 2021;359:188. DOI: 10.1007/s00604-021-05024-4
  50. 50. Butmee P, Mala J, Damphathik C, Kunpatee K, Tumcharern G, Mehmeti KM, et al. A portable selective electrochemical sensor amplified with Fe3O4@Au-cysteamine-thymine acetic acid as a conductive mediator for determination of mercuric ion. Talanta. 2021;221:121669. DOI: 10.1016/2020.121669
  51. 51. Kokulnathan T, Vishnuraj R, Wang TJ, Kumar EA, Pulithadahtil B. Heterostructured bismuth oxide/hexagonal-boron nitride nanocomposite: A disposable electrochemical sensor for detection of flutamide. Ecotoxicology & Environment Safety. 2021;207:111276. DOI: 10.1016/j.ecoenv.2020.111276
  52. 52. Zabiha Pour T, Shahidi S-A, Karimi-Maleh H, Saraei G. A Voltammetric food analytical sensor for determining vanillin based on amplified NiFe2O4 nanoparticle/ionic liquid sensor. Journal of Food Measurement and Characterization. 2020;14:1039-1045. DOI: 10.1007/s11694-019-00353-8
  53. 53. Amini K et al. Review—Recent advances in electrochemical sensor technologies for THC detection. Journal of Cannabis Research. 2022;12:4. DOI: 10.1186/s42238-022-00122-3
  54. 54. Cantalapiedra A, Gismera MJ, Sevilla MT, Procopio JR. Sensitive and selective determination of phenolic compounds from aromatic plants using an electrochemical detection coupled with HPLC method. Phytochemistry Analytical. 2014;25:247-254. DOI: 10.1002/pca.2500
  55. 55. Zhang P, Chun Z, Shao Q, Fu L, Luo Y, Gu D, et al. Evaluation of the phytochemicals and antioxidant activity of Lophatherum gracile Brongn based on chemical fingerprinting by HPLC with electrochemical detection. Journal of Seperation Science. 2021;44:3777-3788. DOI: 10.1002/jssc.202100318
  56. 56. Bounegru AV, Apetrei C. Voltammetric sensors based on nanomaterials for detection of Caffeic acid in food supplements. Chem. 2020;8:41. DOI: DOI.10.3390/chemosensors8020041
  57. 57. Qin Z, Zang J, Liu Y, Wu J, Li G, Liu J, et al. A simple but efficient Voltammetric sensor for simultaneous detection of Tartrazine and Ponceau 4R based on TiO2/electro-reduced graphene oxide nanocomposite. Chem. 2020;8:70. DOI: DOI.10.3390/chemosensors8030070
  58. 58. Yang H, Zhou M, Yang W, Ren G, Ma L. Rolling-made gas diffusion electrode with carbon nanotube for electro-Fenton degradation of acetylsalicylic acid. Chemosphere. 2018;206:439-446. DOI: 10.1016/j.chemosphere.2018.05.027
  59. 59. Rajendrachari S, BE K. Biosynthesis of silver nanoparticles using leaves of acacia Melanoxylon and their application as dopamine and hydrogen peroxide sensors. Physical chemistry. Research. 2020;8:25. DOI: 10.22036/PCR.2019.205211.1688
  60. 60. Liu F, Zhao J, Wang S, Du P, Xing B. Effects of solution chemistry on adsorption of selected pharmaceuticals and personal care products (PPCPs) by Graphenes and carbon nanotubes. Environmental Science and Technology. 2014;48:13197-13206. DOI: 10.1021/es5034684
  61. 61. Shashanka R, Swamy B. E K, simultaneous electro-generation and electro-deposition of copper oxide nanoparticles on glassy carbon electrode and its sensor application. SN Applied Sciences. 2020;2:956. DOI: 10.1007/s42452-020-2785-1
  62. 62. Boxall ABA. The environmental side effects of medication: How are human and veterinary medicines in soils and water bodies affecting human and environmental health? European Molecular Biology Organisation. 2004;5:1110-1116. DOI: 10.1038/sj.embor.7400307
  63. 63. Brody JG, Aschengrau A, McKelvey W, Swartz CH, Kennedy T, Ruthann AR. Breast cancer risk and drinking water contaminated by wastewater: A case control study. Environmental Health. 2006;5:28. DOI: 10.1186/1476-069X-5-28
  64. 64. Sengar A, Vijayanandan A. Human health and ecological risk assessment of 98 pharmaceuticals and personal care products (ppcps) detected in Indian surface and wastewaters. Science Total Environmental. 2022;807:150677. DOI: 10.1016/j.scitotenv.2021.150677
  65. 65. Kruc R, Dragon K, Gorski J. Migration of pharmaceuticals from the Warta River to the aquifer at a riverbank filtration site in Krajkowo (Poland). Water. 2019;11:1238. DOI: 10.3390/w11112238
  66. 66. Hena S, Gutierrez L, Croué JP. Removal of pharmaceutical and personal care products (PPCPs) from wastewater using microalgae: A review. Journal of Hazardous Materials. 2021;403:124041. DOI: 10.1016/j.jhazmat.2020.124041
  67. 67. Kuczynska J, Nieradko-Iwanicka B. The effect of ketoprofen lysine salt on mucosa of rat stomach after ethyl alcohol intoxication. Biomedicine & Pharmacotherapy. 2021;141:111938. DOI: 10.1016/j.biopha.2021.111938
  68. 68. Cao F, Dong Q, Li C, Chen J, Ma X, Huang Y, et al. Electrochemical sensor for detecting pain reliever/fever reducer drug acetaminophen based on electrospun CeBiOx nanofibers modified screen-printed electrode. Sensors & Actuators B: Chemical. 2018;256:143-150. DOI: /10.1016/2017.09.204
  69. 69. Della Pelle F, Angelini C, Sergi M, Del Carlo M, Pepe A, Compagnone D. Nano carbon black-based screen-printed sensor for carbofuran, isoprocarb, carbaryl and fenobucarb detection: Application to grain samples. Talanta. 2018;186:389-396. DOI: 10.1016/j.talanta.2018.04.082
  70. 70. Medsen KG, Skonberg C, Jurva U, Cornett C, Hansen SH, Johansen TN, et al. Bioactivation of diclofenac in vitro and In vivo: Correlation to electrochemical studies. Chemical Research in Toxicology. 2008;21:1107-1119. DOI: 10.1021/tx700419d
  71. 71. Asif AH, Wang S, Sun H. Hematite-based nanomaterials for photocatalytic degradation of pharmaceuticals and personal care products (PPCPs): A short review. Current Opinion in Green and Sustainable Chemistry. 2021;28:10044. DOI: 10.1016/j.cogsc.2021.100447
  72. 72. Bhadra BN, Jhung SH. Adsorptive removal of pharmaceuticals and personal care products from water with functionalized metal-organic frameworks: Remarkable adsorbents with hydrogen-bonding abilities. Scientific Reports. 2016;6:34462. DOI: 10.1038/srep34462
  73. 73. Kavieva L, Giyatdenova G. Voltammetric sensor based on SeO2 nanoparticles and surfactants for indigo carmine determination. Sensors. 2022;22:3224. DOI: 10.3390/s22093224
  74. 74. Giyatdenova G, Gimadutdenova L. Cerium(IV) and iron(III) oxides nanoparticles based Voltammetric sensor for the sensitive and selective determination of Lipoic acid. Sensors. 2021;21:7622-7639. DOI: 10.3390/s21227639
  75. 75. Fan X, Gao J, Li W, Huang J, Yu G. Determination of 27 pharmaceuticals and personal care products (PPCPs) in water: The benefit of isotope dilution. Frontiers in Environmental Science Engineering. 2020;14:8. DOI: 10.1007/s11783-019-1187-3
  76. 76. Kusior A. Voltammetric detection of glucose—The electrochemical behavior of the copper oxide materials with well-defined facets. Sensors. 2022;22:4783. DOI: 10.3390/s22134783
  77. 77. Corcoran J, Tyler CR, Winter MJ. Pharmaceuticals in the aquatic environment: A critical review of the evidence for health effects in fish. Critical Reviews in Toxicology. 2010;40:287-304. DOI: 10.3109/10408440903373590
  78. 78. Moreno FA, Baeza JAM, Sanchez MG, Jimenez JG, Nguyen VAIC. Experimental validation of depth cameras for the parameterization of function of patient in clinical tests balance. Sensors. 2017;17:424. DOI: 10.3390/17020424
  79. 79. Barbosa PFP, Viiera EG, Paim LL, Nakamura PR, Andrade RDA, Carmo DRD. Voltammetric techniques for pesticides and herbicides detection. International Journal of Electrochemical Science. 2019;14:3418-3433. DOI: 10.20964/2019.04.60
  80. 80. Calixto CMF, Cervini P, Cavalheiro ETG. Determination of atenolol in environmental water samples and pharmaceutical formulations at a graphite-epoxy composite electrode. Materials Research. 2009;10:561-570. DOI: 10.1080/03067310903582358
  81. 81. Fockmeniok SM etal. Sensitive carbon fibre microelectrode for the quantification of diuron in quality control of a commercialized formulation. International Journal of Analytical Chemistry. 2022;22:9994639. DOI: 10.1155/2022/9994639
  82. 82. Prasertying P, Jantawong N, Sonsa-Ard T, Wongpakdee T, Khoonrueng N, Buking S, et al. Gold leaf electrochemical sensors: Applications and nanostructure modification. The Analyst. 2021;146:1579-1589. DOI: 10.1039/D0AN02455D
  83. 83. Ghaderpoori M, Jafari A, Nazari E, et al. Preparation and characterization of loaded paraquat- polymeric chitosan/xantan/tripolyphosphatenano capsules and evaluation for controlled release. Journal of Environmental Health Science and Engineering. 2020;18:1057-1066. DOI: 10.1007/s40201-020-00527-3
  84. 84. Simões FR, Mattoso LHC, VazCMP. Modified carbon paste-polyaniline electrodes for the electrochemical determination of the herbicide 2,4-D. Sensor Letters. 2004;2:221-225. DOI: 10.1166/sl.2004.049
  85. 85. Mousty C. Sensors and biosensors based on clay-modified electrodes—New trends. Applied Clay Science. 2004;27:159. DOI: 10.1016/j.clay.2004.06.005
  86. 86. Qiu J, Villemure G. Anionic clay modified electrodes: Electrochemical activity of nickel(II) sites in layered double hydroxide films. Journal of Electroanalytical Chemistry. 1995;395:159. DOI: /10.1016/0022-0728(95)04070-5
  87. 87. Paul PK, Hussain SA, Bhattacharjee D, Pal M. Preparation of polystyrene–clay nanocomposite by solution intercalation technique. Materials Research. 2013;36:361-366. DOI: 10.1007/s12034-013-0498-4
  88. 88. Bouabi YE, Loudiki AK, Matrouf M, Akbour RA, Laghrib F, Farahi A, et al. Clay-based graphite sensor for electrochemical determination of p nitrophenol in water samples. Case Studies in Chemical and Environmental Engineering. 2022;6:100225. DOI: 10.1016/j.cscee.2022.100225
  89. 89. Manisankar P, Selvanathan G, Vedhi C. Determination of pesticides using heteropolyacidmontmorillonite clay-modified electrodes with surfactant. Talanta. 2006;68:686-692. DOI: 10.1016/j.talanta.2005.05.021
  90. 90. Pifferi P, Soliveri G, et al. Electrochemical sensors cleaned by light: A proof of concept for on-site applications towards integrated monitoring systems. RSC Advances. 2015;87:1285. DOI: 10.1039/C5RA12219H
  91. 91. Villulas HM, Costa FIM, Nascente PAP, Bulhoes LOS. Sol -gel prepared Pt- modified oxide layers: Synthesis, characterization and electrolytic activity. Chemistry of Materials. 2006;18:5563-5570. DOI: 10.1021/cm0601178
  92. 92. Mello RLS, Costa FIM, Villulas HM, Bulhoes LOS. Preparation and electrochemical characterization of Pt nanoparticles dispersed on niobium oxide. Analytical Chemistry. 2003;28:2. DOI: 10.1590/S0100-46702003000200009
  93. 93. O'mahony AM, Samek IA, Sattayasamitsathit S, Wang J. Orthogonal identification of gunshot residue with complementary detection principles of voltammetry, scanning electron microscopy, and energy-dispersive X-ray spectroscopy: Sample, screen, and confirm. Analytical Chemistry. 2014;86:8031. DOI: 10.1021/ac5016112
  94. 94. Bandodkar AJ, Mahony AMO, Ramírez J, Samek IA, Anderson SM, Windmiller JR, et al. Solid-state forensic finger sensor for integrated sampling and detection of gunshot residue and explosives: Towards ‘lab-on-a-finger. The Analyst. 2013;18:138. DOI: 10.1039/c3an01179h
  95. 95. Sassolas A, Prieto-Simón B, Marty JL. Biosensors for pesticides detection. American Journal of Analytical Chemistry. 2012;3:210. DOI: 10.4236/ajac.2012.33030
  96. 96. Verma N, Bhardwaj A. Biosensor Technology for Pesticides. Applied Biochemistry and Biotechnology. 2003;60:29. DOI: 10.1007/s12010-015-1489-2
  97. 97. Hayat A, Marty JL. Aptamer based electrochemical sensors for emerging environmental pollutants. Frontiers in Chemistry. 2014;2:41. DOI: 10.3389/fchem.2014.0004
  98. 98. Rodriguez-Mozaz S, Maria J, Alda LDE, Barcelo D. Biosensors as useful tools for environmental analysis and monitoring. Analytical and Bioanalytical Chemistry. 2006;386:1025. DOI: 10.1007/s00216-006-0574-3
  99. 99. Salgado AM, Silva LM, Melo AF. Biosensor for environmental applications. Environmental Biosensor. 2011;386:1025-1041. DOI: 10.5772/20154
  100. 100. Webb S, Ternes T, Gibert M, Olejniczak K. Indirect human exposure to pharmaceuticals via drinking water. Toxicology Letters. 2003;142:157-167. DOI: 10.1016/s0378-4274(03)00071-7
  101. 101. Navrátilová I, Skládal P. Theimmunosensors for measurement of 2,4-dichlorophenoxyacetic acid based on electrochemical impedance spectroscopy. Bioelectrochemistry. 2004;42:11-18. DOI: 10.1016/j.bioelechem.2003.10.004
  102. 102. Maggioni S, Balaguer P, Chiozzotto C, Benfenati E. Screening of endocrine-disrupting phenols, herbicides, steroid estrogens, andestrogenicity in drinking water from the waterworks of 35 Italian cities and from PET-bottled mineral water. Environmental Science Pollution Research. 2013;20:1649-1660. DOI: 10.1007/s11356-012-1075-x
  103. 103. Zwiener C. Occurrence and analysis of pharmaceuticals and their transformation products in drinking water treatment. Analyst Bioanalytical Chemistry. 2007;387:1159-1162. DOI: 10.1007/s00216-006-0818-2
  104. 104. Tran-Minh C, Pandey PC, Kumaran. Studies on acetylcholine sensor and its analytical application based on the inhibition of cholinesterase. Biosensors and Bioelectronics. 1990;5:461. DOI: 10.1016/0956-5663(90)80035-C
  105. 105. Martiniano LC et al. Direct simultaneous determination of Pb(II) and Cu(II) in biodiesel by anodic stripping voltammetry at a mercury-film electrode using microemulsions. Fuel. 2013;103:1164-1167. DOI: 10.1016/j.fuel.2012.07.002
  106. 106. Almeida JMS et al. A simple electroanalytical procedure for the determination of calcium in biodiesel. Fuel. 2014;26:1794. DOI: 10.1016/j.fuel.2013.07.088
  107. 107. Zhang Y, Zhuang HS. Amperometric immunosensor based on layer-by-layer assembly of thiourea and nano-gold particles on gold electrode for determination of naphthalene. Chinese Journal of Analytical Chemistry. 2010;38:153-157. DOI: 10.1016/S1872-2040(09)60021-9
  108. 108. Ma R, Wang B, Lu S, Zhang Y, Yin L, Huang J, et al. Characterization of pharmaceutically active compounds in Dongting Lake, China: Occurrence, chiral profiling and environmental risk. Sci. Total Environ. 2016;268:557-558. DOI: 10.1016/j.scitotenv.2016.03.053
  109. 109. Frena M, Campestrini I, OCDE B, Spinelli A. In situ bismuth-film electrode for square-wave anodic stripping voltammetric determination of tin in biodiesel. Electrochimica Acta. 2011;56:4678. DOI: 10.1016/j.electacta.2011.02.111
  110. 110. Oleveria TDC, Freiteies JM, Richter EM, Munoz RAA. A batch injection analysis system with square-wave voltammetric detection for fast and simultaneous determination of naphazoline and zinc. Talanta. 2016;152:308-313. DOI: 10.1016/j.talanta.2016.02.031
  111. 111. Niu X, Lan M, Zhao H, Chen C, Li Y, Zhu X. Review: Electrochemical stripping analysis of trace heavy metals using screen-printed electrodes. Analytical Letters. 2013;46:2479-2502. DOI: 10.1080/00032719.2013.805416
  112. 112. Shashanka R, Jayprakash GK, Kumar M, Swamy BEK. Electrocatalytic determination of ascorbic acid using a green synthesized magnetite nano-flake modified carbon paste electrode by cyclic voltammetric method. Material Research Innovations. 2022;26:4. DOI: 10.1080/14328917.2021.1945795
  113. 113. Erickson J et al. A simple and inexpensive electrochemical assay for the identification of nitrogen containing explosives in the field. Sensors. 2017;17:1769. DOI: 10.3390/s17081769
  114. 114. Tercier ML, Buffle J, Zirino A, Vitre RR. In situ voltammetric measurement of trace elements in lakes and oceans. Analytica Chimica Acta. 1990;237:429-437. DOI: 10.1016/s0003-2670(00)83947-1
  115. 115. Tercier-Waeber ML, Buffle J, Graziottin F. A novel voltammetric in-situ profiling system for continuous real-time monitoring of trace elements in natural waters. Electroanalysis. 1998;10:355-363. DOI: 10.1002/(SICI)1521-4109(199805)10:6<355::AID-ELAN355>3.0.CO;2-F
  116. 116. Hughes G, Westmacott K, Honeychurch KC, Crew A, Pemberton RM, Hart JP. Recent advances in the fabrication and application of screen-printed electrochemical (bio) sensors based on carbon materials for biomedical, Agri-food and environmental analyses. Biosensors. 2016;6:50. DOI: 10.3390/bios6040050
  117. 117. Tyszczuk-Rotko K, Szwagierek A. Green electrochemical sensor for caffeine determination in environmental water samples: The bismuth film screen-printed carbon electrode. Journal of The Electrochemical Society. 2017;164:B342. DOI: 10.1149/2.0571707
  118. 118. Wang N, Kanhere E, Kottapalli AGP, Miao JM, Triantafyllou MS. Flexible liquid crystal polymer-based electrochemical sensor for in-situ detection of zinc (II) in seawater. Microchimica Acta. 2017;184:3007-3015. DOI: 10.1007/s00604-017-2280-6
  119. 119. Farré MI, Pérez S, Kantiani L, Barceló D. Fate and toxicity of emerging pollutants, their metabolites and transformation products in the aquatic environment. Trends in Analytical Chemistry. 2008;27:991-1007. DOI: 10.1016/j.trac.2008.09.010
  120. 120. Brillas E. A critical review on ibuprofen removal from synthetic waters, natural waters, and real wastewaters by advanced oxidation processes. Chemosphere. 2022;286:131849. DOI: 10.1016/j.chemosphere.2021.131849
  121. 121. Burns EE, Carter LJ, Snape J, Thomas-Oates J, Boxall ABA. Application of prioritisation approaches to optimize environmental monitoring and testing of pharmaceuticals. Journal of Toxicology Environment and Health B. 2018;21:115-141. DOI: 10.1080/10937404.2018.1465873
  122. 122. Asif AH, Wang S, Sun H. Hematite-based nanomaterials for photocatalytic degradation of pharmaceuticals and personalcare products (PPCPs): A short review. Current Opinion in Green Sustainable Chemistry. 2021;28:100447. DOI: 10.1016/j.cogsc.2021.100447
  123. 123. Nielsen M, Larsen LH, Jetten MSM, Revsbech NP. Bacterium-based NO2 biosensor for environmental applications. Applied Environmental Microbiology. 2004;70:6551-6558. DOI: 10.1128/AEM.70.11.6551-6558.2004
  124. 124. Revsbech NP, Glud RN. Biosensor for laboratory and lander-based analysis of benthic nitrate plus nitrite distribution in marine environments. Limnology and Oceanographic Methods. 2009;7:761-770. DOI: 10.4319/lom.2009.7.761
  125. 125. Bahnemann D. Photocatalytic water treatment: Solar energy applications. Journal Solar Energy. 2004;77:445-459. DOI: 10.1016/j.solener.2004.03.031
  126. 126. Rodrigues PG, Gonçalves LM, Magalhães PJ, Pacheco JG, Rodrigues JA, Barros AA. Voltammetric analysis of metallothioneins and copper (II) in fish for water biomonitoring studies. Environmental Chemistry Letters. 2011;9:405. DOI: 10.1007/s10311-010-0293-z
  127. 127. Castro SVF, Lima AP, Rocha RG, Cardoso RM, Montes RHO, Santana MHP, et al. Simultaneous determination of lead and antimony in gunshot residue using a 3D-printed platform working as sampler and sensor. Analytica Chimica Acta. 2020;1130:126. DOI: 10.1016/j.aca.2020.07.033
  128. 128. Garcia-Segura S, Lanzarini-Lopes M, Hristovski K, Westerhoff P. Electrocatalytic reduction of nitrate: Fundamentals to full-scale water treatment applications. Applied Catalysis B: Environmental. 2018;236:546-568. DOI: 10.1016/j.apcatb.2018.05.041
  129. 129. Legrand DC, Barus C, Garcon V. Square wave voltammetry measurements of low concentrations of nitrate using Au/AgNPs electrode in chloride solutions. Electroanalysis. 2017;29:2882-2288. DOI: 10.1002/elan.201700447
  130. 130. Koyun O, Sahin Y. Voltammetric determination of nitrite with gold nanoparticles/poly(methylene blue)-modified pencil graphite electrode: Application in food and water samples. Ionics. 2018;24:3187-3197. DOI: 10.1007/s11581-017-2429-7
  131. 131. Sun C et al. Electrochemical sensor for nitrite using a glassy carbon electrode modified with gold-copper nanochainnetworks. Analytical Method. 2019;19:39. DOI: 10.1039/C9AY01544B
  132. 132. Han YJ, Zhang R, Dong C, Cheng FQ, Guo YJ. Sensitive electrochemical sensor for nitrite ions based on rose-like AuNPs/MoS2/graphene composite. Biosensors and Bioelectronics. 2019;142:111529. DOI: 10.1016/j.bios.2019.111529
  133. 133. Hrastnik NI, Jovanovski V, Hocevar SB. In-situ prepared copper film electrode for adsorptive stripping voltammetric detection of trace Ni(II). Sensors & Actuators B: Chemical. 2020;30:127637. DOI: 10.1016/j.snb.2019.127637
  134. 134. Wang TT, Yue W. Carbon nanotubes heavy metal detection with stripping voltammetry: A review paper. Electroanalysis. 2017;29:2178-2189. DOI: 10.1002/elan.201700276
  135. 135. Ustundag I, Erka AI, Koralay T, Kadioglu YK, Jeon S. Gold nanoparticles included graphene oxide modified electrode: Picomole detection of metal ions in seawater by stripping voltammetry. Journal of Analytical Chemistry. 2016;71:685-695. DOI: 10.1134/S1061934816070108
  136. 136. Rajendrachari S, Ramakrishna D. 1 - functionalized nanomaterial-based electrochemical sensors: A sensitive sensor platform. Woodhead Publishing Series in Electronic and Optical Materials. 2022;3:25. DOI: 10.1016/B978-0-12-823788-5.00010-7
  137. 137. Large RR, Halpin JA, Lounejeva E, Danyushevsky LV, Maslennikov VV, Gregory D, et al. Cycles of nutrient trace elements in the Phanerozoic Ocean. Gondwana Research. 2015;28:1282-1293. DOI: DOI.10.1016/j.gr.2015.06.004
  138. 138. Daniel A, Laes-Huon A, Barus C, Beaton AD, Blandfort D, Guigues N, et al. Toward a harmonisation for using in situ nutrient sensors in the marine environment. Frontiers in Marine Science. 2020;6:773. DOI: 10.3389/fmars.2019.00773
  139. 139. Han HT, Tao WY, Hu XP, Ding XY, Pan DW, Wang CC, et al. Needle shaped electrode for speciation analysis of copper in seawater. Electrochimica Acta. 2018;289:474-482. DOI: 10.1016/j.electacta.2018.08.097
  140. 140. Zhang Y, Nie JT, Wei HY, Xu HT, Wang Q, Cong YQ, et al. Electrochemical detection of nitrite ions using Ag/Cu/MWNT nanoclusters electrodeposited on a glassy carbon electrode. Sensors & Actuators B: Chemical. 2018;258:1107-1116. DOI: 10.1016/j.snb.2017.12.001
  141. 141. Trejos T, Vander C, Pyl M-HK, Arroyo AAL, LE. Fast identification of inorganic and organic gunshot residues by LIBS and electrochemical methods. Forensic. Chemistry. 2018;8:146. DOI: 10.1016/j.forc.2018.02.006
  142. 142. Salles MO, Naozuka J, Bertotti M. A forensic study: Lead determination in gunshot residues. Microchemical Journal. 2012;101:49-53. DOI: 10.1016/j.microc.2011.10.004
  143. 143. Somashekharappa KK, Rajendrachari S. Sustainable development information management of carbon nanomaterial-based sensors. Emerging Research Trends in Devices and Applications. 2022;3:12. DOI: 10.1016/B978-0-323-91174-0.00001-9
  144. 144. Hashim NA, Hashim M, Zain ZM, Jaafar MZ. Copper determination in gunshot residue by cyclic Voltammetric and inductive coupled plasma-optical emission spectroscopy. MATEC Web of Conferences. 2016;59:04005. DOI: 10.1051matec 59 conf/2016 04005
  145. 145. Salles MO, Bertotti M, Paixão TRLC. Use of a gold microelectrode for discrimination of gunshot residues. Sensors and Actuators B: Chemical. 2012;166-167:848-852. DOI: 10.1016/j.snb.2012.02.097
  146. 146. Spanu D, Binda G, Dossi C, Monticelli D. Biochar as an alternative sustainable platform for sensing applications: A review. Microchemical Journal. 2020;159:105506. DOI: 10.1016/j.microc.2020.105506
  147. 147. Erden S, Durmus Z, Kiliç E. Simultaneous determination of antimony and Lead in gunshot residue by Cathodic adsorptive stripping Voltammetric methods. Electroanalysis. 2011;23:1967-1974. DOI: 10.1002/elan.201000612
  148. 148. Vuki M, Shiu KK, Galik M, Mahony AM, Wang J. Simultaneous electrochemical measurement of metal and organic propellant constituents of gunshot residues. The Analyst. 2012;137:3265-3270. DOI: 10.1039/c2an35379b
  149. 149. Bessa BGDO, Silva-Neto HDEA, Coltro WKT, Rocha TL, Lopes WR. Lead toxicity in Lucilia cuprina and electrochemical analysis: a simple and low-cost alternative for forensic investigation. Analytical & Bioanalytical Chemistry. 2021;413:3201-3208. DOI: 10.1007/s00216-021-03257-z

Written By

Harsha Devnani and Chetna Sharma

Submitted: 21 July 2022 Reviewed: 13 October 2022 Published: 02 December 2022