You are on page 1of 22

International Journal of Environmental Analytical

Chemistry

ISSN: (Print) (Online) Journal homepage: https://www.tandfonline.com/loi/geac20

Sensing of Alphacypermethrin Pesticide Using


Modified Electrode of Chitosan-Silver Nanowire
Nanocomposite Langmuir Blodgett Film

Sanjeev Bhandari, Abhijit Nath, L Robindro Singh, Manashjit Gogoi, Basudev


Pradhan & Mrityunjoy Mahato

To cite this article: Sanjeev Bhandari, Abhijit Nath, L Robindro Singh, Manashjit Gogoi, Basudev
Pradhan & Mrityunjoy Mahato (2022): Sensing of Alphacypermethrin Pesticide Using Modified
Electrode of Chitosan-Silver Nanowire Nanocomposite Langmuir Blodgett Film, International
Journal of Environmental Analytical Chemistry, DOI: 10.1080/03067319.2022.2052865

To link to this article: https://doi.org/10.1080/03067319.2022.2052865

View supplementary material

Published online: 05 Apr 2022.

Submit your article to this journal

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=geac20
INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY
https://doi.org/10.1080/03067319.2022.2052865

Sensing of Alphacypermethrin Pesticide Using Modified


Electrode of Chitosan-Silver Nanowire Nanocomposite
Langmuir Blodgett Film
Sanjeev Bhandaria, Abhijit Natha, L Robindro Singhb, Manashjit Gogoi c
,
Basudev Pradhand and Mrityunjoy Mahatoa
a
Physics Division, Department of Basic Sciences and Social Sciences, School of Technology, North Eastern Hill
University, Shillong, India; bDepartment of Nanotechnology, School of Technology, North Eastern Hill
University, Shillong, India; cDepartment of Biomedical Engineering, School of Technology, North Eastern Hill
University, Shillong, India; dNew Generation Photovoltaics Research Laboratory, Department of Energy
Engineering, Central University of Jharkhand, Ranchi, India

ABSTRACT ARTICLE HISTORY


We report on the sensing of a hazardous pesticide alphacyper­ Received 19 August 2020
methrin (ACM) using electrochemical technique through Accepted 2 March 2022
a nanocomposite film modified electrode. The nanocomposite KEYWORDS
film comprising of octadecylamine, chitosan, polyvinyl alcohol- Sensor; pesticide;
silver nanowires and haemoglobin, shortly (OCPAH), was pre­ alphacypermethrin;
pared by Langmuir–Blodgett (LB) film deposition technique on electrochemical; Langmuir–
various substrates and electrodes. The composite LB film was Blodgett film; polymer
characterised by UV-visible spectroscopy and scanning electron nanocomposite
microscopy, which confirms the stable and multilayer film. The LB
film modified electrode was used for sensing ACM pesticide by
cyclic voltammetry (CV), differential pulse voltammetry (DPV), and
square wave voltammetry (SWV) techniques. The achieved sen­
sing parameters such as limit of detection as 14 nM, 5 nM and
10 nM; linear range as 10–100 nM, 10–40 nM and 50–100 nM/10–
100 nM; and sensitivity as 0.418 µA/nM/cm2, 0.259 µA/nM/cm2
and 0.271 µA/nM/cm2 for CV, DPV and SWV techniques, respec­
tively. The reported sensor is found to have stability of 74% upto
20 cycles, the relative standard deviation (RSD) value for metal
ion/organic interference species as 2% and for real samples are
within 1.4% using CV technique. The reported nanocomposite-
based ACM pesticide sensor will open up new options for
research on LB film nanocomposite-based sensing of different
organophosphorus groups of pesticides.

1. Introduction
Pesticides are generally used in agriculture to kill or repel the plant attacking pests. Recent
times, use of pesticide has increased exponentially to meet the food demand of the growing
population [1]. Farm use of pesticide is highest in tropical regions, and it is estimated that
one-third of the global crop production currently depends on pesticide [2,3]. There was an
increase in agro production from 51 million tonnes (MT) (1950–1951) to 252 MT (2016–2017)

CONTACT Mrityunjoy Mahato mrityunjoyphy@gmail.com


Supplemental data for this article can be accessed here
© 2022 North-Eastern Hill University
2 M. MAHATO ET AL.

during the Indian green revolution, where pesticide played a major role [4,5]. However, 0.1%
of the applied pesticide reaches the target pest [6] and, remaining is contaminating water,
environment and even food chain, causing neurological, immunological, respiratory and
cancer [7]. To counter pesticide problems, some strategies are evolving such as to minimise
the use of synthetic pesticides, to maximise the use of biopesticide and to promote organic
farming [8]. Hence, it is essential to analyse pesticide quantitatively with a portable, low cost,
high sensitive pesticide sensor.
There are conventional large size analytical instruments such as gas chromatography-
mass spectrometry (GC-MS) and high-performance liquid chromatography (HPLC) [6,9].
However, these suffer from some disadvantages like time consuming, expensive, large
size, high maintenance cost and require trained technicians, which limit their field
application [10]. The approximate cost of GC-MS and HPLC are around 39.5k USD and
14.9k USD, respectively [11]. Hence, there is enough scope to develop a portable, sensitive
pesticide sensor with the help of techniques such as electrochemical, optical, etc [12].
Recent studies on enzyme-based and nanocomposite-based electrochemical sensors
have shown higher sensitivity for detection of pesticide; however, it requires careful
and innovative preparation of the composite matrices [13,14].
In this regard, the Langmuir–Blodgett (LB) method is a suitable tool for preparing
uniform multilayer composite film on the electrode [15,16]. Cyclic voltammetry (CV)-
based electrochemical tools can provide better stability of sensor [17], electron transfer
kinetics [18], reversibility of a reaction [19] and a better calibration curve [20]. The
conducting nanomaterials in the sensing matrix such as carbon nanotube (CNT), gold
nanoparticles (AuNPs), zirconium dioxide (ZrO2), tin oxide (SnO2), etc., have been used to
improve the sensor performance due to their high surface area, catalytic and conducting
properties [21–24]. Chitosan have been selected as robust, biocompatible matrix for
developing sensing platform due to its film-forming ability through its cationic amino
group and chelating property to metal ions [25]. Chitosan allows protein/enzymes to
retain their conformation through its water bound molecules [26]. Protein/enzymes have
been used as primary recognition elements for pesticide sensors due to their redox and
catalytic properties such as acetylcholinesterase (AChE), glucose oxidase, haemoglobin,
etc [27–29]. We have selected haemoglobin as redox-active protein in the composite film.
Haemoglobin acts as redox protein in the composite and silver nanowires (AgNWs) help
to improve current output of the sensor due to its higher surface area and catalytic
property [30].
Alphacypermethrin (ACM) (MW 416.3 g/mol) has been selected as the model and
target pesticide due to its versatile uses in agriculture and gardening [31]. It is
a pesticide belonging to the organophosphate group (OPs), which is structurally
related to neurotoxic and consists of a phosphorus-oxygen (P¼O) or phosphorus-
sulfer (P¼S) double bond [32]. ACM is a mixture of two cypermethrin stereoisomers
[33], which are used for a wide range of insect pests living in fruits, vegetables and
tobacco such as Lepidoptera, Coleopteran, etc [34]. ACM belongs to type II pyre­
throid and is toxic to mammals, humans and aquatic environments [35].
We report on the sensing of ACM pesticide, which is largely ignored in the literature
unlike its detection study by big size, costly instruments. A novel chitosan-silver nanowire-
based nanocomposite LB film was prepared to modify a platinum (Pt) electrode. We
achieved sensing parameters such as; limit of detection (LOD) as 14 nM, 10 nM and 5 nM;
INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY 3

linear range (LR) as (10–100 nM, 10–100 nM and 10–40 nM/50–100 nM); and sensitivity as
0.418 µA/nM/cm2, 0.271 µA/nM/cm2 and 0.259 µA/nM/cm2 using CV, square wave vol­
tammetry (SWV) and differential pulse voltammetry (DPV) techniques, respectively. The
reported sensor is found to have stability of 74% upto 20 cycles, relative standard
deviation (RSD) value for metal ion/organic interference as 2% and for real samples within
1.4% using CV technique. The results have been compared with other different existing
detection methods on ACM pesticide. Our study explores research options for sensing
study using LB nanocomposite for different kinds of pesticides.

2. Material and methods


2.1. Materials
Haemoglobin (Hb) stored below 4°C and octadecylamine (ODA) was purchased from
Sigma Aldrich. The chloroform, chitosan (CS) (extra pure with viscosity 100 mPas), silver
nitrate (AgNO3) (extra pure AR), polyvinylpyrrolidone (PVP), ethylene glycol, sodium
chloride (NaCl) and acetic acid were purchased from SRL, India. ACM was purchased
from Indofil Industries Limited, India. Milli-Q deionised water with pH = 6.5 and resistiv­
ity = 18.5 MΩ × cm was used for LB experiment. The chemical structures of the constituent
chemicals of the nanocomposite and pesticide analyte are presented in Figure 1.
The CS solution with a concentration of 0.1 M was prepared using acetic acid (0.1 M)
solvent with continuous stirring for 48 hours. The Hb solution (0.05 mg/ml) was prepared
by dissolving Hb in Milli-Q water. ODA solution (1 mM) was prepared by mixing ODA in
chloroform solvent with stirring.

2.2. Preparation of silver nanowire


The AgNWs were prepared following the polyol process with a little bit of mod­
ification [30]. PVP solution (50 mM) was prepared in 20 ml ethylene glycol solvent
under stirring. AgNO3 and NaCl were added to get the final concentration of AgNO3

Figure 1. Molecular structure of Chitosan (CS), Octadecylamine (ODA), Polyvinylpyrrolidone (PVP),


Haemoglobin (Hb) and Alphacypermethrin (ACM) used in the present work.
4 M. MAHATO ET AL.

(3 mM) in 20 ml. The mixture was then heated using a kitchen grade LG microwave
oven at 360 Watt for 3.5 minutes to obtain AgNWs. The final product was washed
thoroughly with acetone to remove ethylene glycol and PVP with repeated filtering
using filter paper (Whatman grade 602 H with pore size <2 µm) followed by
washing with deionised water for multiple times. The sample was then centrifuged
to remove unreacted PVP.

2.3. Cleaning of LB trough, substrate and electrode


The LB trough was filled with Milli-Q water and kept for 2 hours to remove any impurity
and dust in the trough through the process of leaching. The trough and the barrier were
cleaned by acetone, chloroform and tissue paper. The impurities present on the subphase
were sucked by the aspirator pump. This process was performed several times to ensure
that the water subphase was cleaned. All the substrates (glass, indium tin oxide (ITO),
quartz, silicon wafer) were cleaned by ultrasonication along with soap solution, acetone
and water. A uniform layer of water onto the slide confirmed the hydrophilicity of the
slide. The electrodes (working, counter and reference) were sonicated for 5 minutes each
with acetone, water and ethanol. It was kept in a desiccator for drying and for further
experiment.

2.4. LB film preparation


A computerised LB film deposition instrument (model APEX LB-2007DC, Apex
Instruments Co. India) was used for monolayer study and LB film preparation. The LB
trough is made up of teflon and enclosed in a plexiglass box to reduce film contamina­
tion. A Wilhelmy-type balance (with accuracy of ±0.01 mN/m) was used to measure
surface pressure on the water subphase of LB trough. The trough dimensions are width
(130 mm), length (330 mm) and depth (20 mm). The deionised Milli-Q water was used
to prepare the subphase. The pH and the resistivity of freshly prepared water were 6.8
and 18.2 MΩ × cm, respectively. All experiments were performed at a temperature of
20 ± 0.5°C. At least three independent runs were performed to check the
reproducibility.
The nanocomposite LB film has been prepared using LB technique. The composite
monolayer of OCPAH was prepared by spreading different solutions on a subphase
using a micro syringe (Hamilton, Switzerland) within the surface pressure limit of 0.5
mN/m. The ODA was used as a floating lipid matrix, chitosan as a biocompatible
polymer matrix, haemoglobin as electrochemically redox-active protein and AgNW
nanocomponent as electron transfer enhancer [36–38]. Initially, CS, PVP-AgNW and
Hb solutions were spread, followed by the spreading of ODA to form a floating
composite Langmuir monolayer. A reaction time of 30 minutes was given for stabilis­
ing the composite monolayer, and then the barrier was compressed at the rate of
5 mm/min. The pressure-area isotherm was recorded for each pure monolayer and
composite monolayer.
The LB films were prepared by dipping the substrate or electrode into the water
subphase before monolayer preparation. The film length and perimeter of the slide
were kept as 19 mm and 60 mm, respectively. The monolayer was transferred to the
INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY 5

substrate with lifting speed and dipping speed as 3 mm/min and 4 mm/min,
respectively. The drying time for above and below the water subphase was kept as
5 minutes and 1 minute, respectively. The drying time after the 1st and 2nd
layers was kept as 20 minutes so that the deposition of the first layer is strong on
the substrate.
Surface pressure-time (π-t) kinetics of the monolayer was carried out by spreading the
prepared solution on water subphase. The pressure-time measurements were recorded
by the LB film deposition system, which gives the kinetics of the monolayer as a function
of time.

2.5. Composite film characterisation


UV-visible spectroscopic characterisation was carried out on Hitachi Spectrometer
(Model U-3900) in the absorption mode with wavelength range 200–800 nm, scan
speed 300 nm/min, slit width 2 nm and optical path length 10 mm. The UV-visible
spectra of nanocomposite LB film were recorded on a rectangular shape Quartz
substrate. The fourier transform spectroscopic (FTIR) characterisation was done with
the Bruker spectrometer (Model ALPHA) with OPUS 7.5 software on a silicon wafer
substrate for powder and films.
Scanning electron microscopy (SEM) imaging of the composite films were carried out
using SEM (JEOL, Model JSM-6360) for micrometre scale and FE-SEM (ZIESS, Model
SIGMA 300) for nanometre scale. The nanocomposite LB film was deposited on a glass
slide for SEM study. The operating voltage used was 20 KV. The film samples were
mounted on brass stubs (30 mm diameter, 10 mm height) with the help of double side
adhesive tape and gold conducting coating was made by Gold sputter (JFC 1100, JEOL)
to increase conductivity and to protect nanocomposite film from the burning effect of
electron beam.
Transmission electron microscopy (TEM) characterisation was done by JEOL TEM
instrument (Model JEM-100 CX II) using Cu-coated carbon grid with a resolution of
1.4 A°, operating voltage 20–100 kV with 20 kV steps and magnification of 100,000–
450,000×. The X-ray diffraction (XRD) characterisation was done using the Rigaku XRD
instrument (Model TTRAX-III) using Cu-Kα X-ray with λ = 1.5406 A° and scanning rate
of 2°/minute. The XRD was used for AgNW characterisation.

2.6. Pesticide sensor characterisation


The platinum (Pt) working electrode was modified by LB film deposition technique.
Briefly, the prepared solutions of ODA, Hb, PVP-AgNW and CS were spread on the LB
subphase and a waiting time of 30 minutes was given to form a stable composite
monolayer. The LB film was lifted using the mentioned LB parameter on the elec­
trode and dried in a vacuum desiccator.
The CV, SWV and DPV measurements of LB film modified Pt electrode were carried on CH
Instrument (Model CHI660D) with three electrode systems such as working electrode (green
probe), counter electrode (red probe) and reference electrode (white probe). We used a glassy
carbon electrode (GCE) as a counter electrode, silver/silver chloride (Ag/AgCl) as a reference
and Pt as a working electrode. Initially, the CH instrument was given 15 minutes of warming
6 M. MAHATO ET AL.

Figure 2. (a) Pressure-area isotherm of pure ODA, pure CS, ODA+CS and OCPAH composite monolayer.
(b) Pressure-time kinetics data and their fitting data by double exponential decay equation (c) UV-
visible spectra of composite LB film with different layer numbers (1–17 layer) (d) Peak absorbance (at
350 nm) with layer no and its linear fitting.

time followed by a hardware test and selection of CV technique. The initial and final
voltages were kept at −1 V and +1 V, respectively. The scan rate and sensitivity were optimised
as 0.1 V/s and 5 × 10−5 A/V, respectively. A KCl solution (0.1 M) was chosen as electrolyte and
ACM pesticide was mixed with the electrolyte for sensing study in the concentration range of
10–100 nM.

3. Results and discussion


3.1. Study of LB isotherm, pressure-area kinetics and UV-visible spectra of
composite LB film
The composite monolayer was studied in the LB trough and subsequently transferred onto the
substrate and electrode surface. Figure 2a shows the surface pressure-molecular area (π–A)
isotherm of pure ODA, pure CS and the composite Langmuir monolayer. The lifting area/
molecule (nm2/molecule) of pure ODA and pure CS monolayer were found to be 0.12 and
0.2 nm2, respectively, which are in line with the earlier literature of ODA and CS monolayer
[39,40]. There is an appreciable shifting of the isotherm of composite monolayer compared to
INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY 7

the characteristic lifting area of pure ODA and CS isotherms, which indicate the formation of
composite [41,42]. The reported OCPAH nanocomposite LB film is the first ever to be reported
in the literature.
Figure 2b shows the pressure-time (P-t) kinetics study of the different composite
monolayer after normalising the initial pressure value. It indicates the exponential
decrease of surface pressure and final stabilisation within 30 min to 1 hour. The changes
in the kinetics of composite monolayer than the pure monolayer also indicate the
formation of composite monolayer [37,38]. The kinetics have been fitted with double
exponential decay equation (Equation 1) [43,44]. The fitting parameters as well as the
initial pressure of the decay kinetics are summarised in Table 1.
P ðt Þ t=τ1 t=τ2
¼ a1 e þ a2 e þc (1)
P ð0 Þ
Where a1, a2 and c are the parameters for reorganisation processes of the composite
monolayer and τ1 and τ1 are the corresponding time constants. The values of a1 and a2
signify two components of the reorganisation process of monolayer; initially diffusion of
the molecule and subsequent adsorption by air/water interface [44]. This phenomenon
was observed in the case of other protein monolayers such as haemoglobin, ovalbumin,
etc [36,37]. From Table 1, it is evident that the value of a2 is higher and the value of τ2 is
lower, which indicates that the adsorption process is getting stronger with higher
magnitude and faster with less time. It is interesting to note that there is negligible
variation of the value of ‘c’, which is merely a mathematical constant and does not affect
the reorganisation process of the monolayer [37].
Figure 2c shows the UV-visible spectra of OCPAH composite multilayer LB film with
different numbers of layers (1, 5, 9, 12 and 17 layers), and Figure 2d shows the linear fitting
of absorption intensity with layer numbers. The spectra contains a weak peak around
350 nm due to thin monolayer film, and it conforms to the presence of AgNW in the
composite film [45]. The correlation coefficient of linear fitting was achieved as R2 = 0.93,
which indicates that the layer by layer film was transferred linearly on the substrate
without any loss of material. The peak at 350 nm corresponds to AgNW in nanocomposite
film sample, while the absorption peak at 399 nm corresponds to AgNW in solution, and
this shifting suggests the formation of composite of AgNW with other components in the
film [46].

3.2. Characterisation of AgNWs


AgNWs have been used due to its high conductivity, high surface area due to wire shape
as well as electron transfer catalytic property, which is expected to increase the current
value in electrochemical study [47]. The large surface area of AgNW provides greater
chance to target analyte for reacting with the surface of nanowire, thereby increasing the
sensitivity of the sensor. Also, silver is a highly active electrocatalyst in alkaline solutions
that influence the kinetics of the reaction [48,49].
Figure 3a shows XRD peaks of AgNW corresponding to the planes (111), (200), (220),
(311) and (222) which confirms the face-centred cubic (FCC) silver nanostructure as per
the Joint Committee on Powder Diffraction Standards (JCPDS File No 04–0783) [45]. Plane
(111) has the maximum intensity and the calculated lattice constant for plane (111) to be
8 M. MAHATO ET AL.

0.4084 nm, which is in agreement with the literature value of 0.4086 nm [45], and it
indicates the high purity of FCC silver nanowire [50]. Figure 3b shows the UV-visible
spectra of AgNW in solution having peaks at 399 nm due to the plasmon excitation of
electrons of silver nanowire [45]. The shoulder peak at 350 nm is due to plasmon response
of bulk silver, which is also observed generally for silver nanowire due to wire shape [46].
Figure 3c shows the TEM images of AgNW having noncircular tips, which have
important consequences for sensing [51]. The PVP used for AgNW preparation is acting
as a reducing as well as capping agent. It may selectively bind to (100) facets and allow
crystal growth along (111) facets as evidenced by its corresponding stronger XRD peak
[52]. The d-value was found to be 0.235 nm from TEM fringe pattern (Figure 3d) and
0.226 nm from TEM SAED pattern, which is in agreement with the literature value of
0.235 nm [53].

3.3. Characterisation of composite LB Film


Figure 4 shows the FTIR spectra of the composite film and its components chemicals
in powder (Panel-I) and in film (Panel-II). It shows the characteristic IR peaks for
chitosan, PVP and ODA as agreed from literature [54–56]. The Table 2 shows the peak
assignment of the FTIR spectra of the nanocomposite. The peak at 1526 cm−1 in the
chitosan FTIR spectra is due to the N-H bending vibration bond and shifted to
1523 cm−1 due to Ag adsorption on nitrogen atoms after composite formation
[54]. The peak at 2884 cm−1 due to stretching vibration of C = O and shifted to
the lower wavenumber 2850 cm−1 due to the interaction of Ag with the oxygen
atom of PVP [57].
Figure 5 shows the SEM images of the nanocomposite LB film to search AgNW, Hb
protein as well as to study the film morphology. Figure 5(a,b) shows SEM image in
the micrometre scale, and Figure 5(c,d) shows FE-SEM in the nanometre scale show­
ing the uniform film containing few aggregated Hb protein molecules along with the
AgNW. The SEM image confirms the formation of uniform composite monolayer
containing Hb protein on glass substrate (Figure 5d). This kind of protein LB mono­
layer SEM image was found in our earlier study also [38, 41,42]. The SEM and FE-SEM
images were taken in two different SEM instruments just to come down to the nm
range to get the protein aggregated image, and it became a complementary evi­
dence also. The average diameter and length of the AgNWs were found to be 50 nm
and 5 µm, respectively, as evident in the SEM image (Figure 5a) and TEM image
(Figure 3a).

Table 1. Fitting Parameters of Pressures-Time (P-t) Data of Different Composite Monolayers by Double
Exponential Decay Equation.
Monolayer
Sl. No Constituent Initial Pressure (mNm−1) a1 and a2 τ1 and τ2 (s) C R2 value
1 ODA 23.35 0.418, 145.897, 892.352 0.419 0.99
0.262
2 CS-ODA 31.00 0.405, 149.701, 830.995 0.410 0.99
0.344
3 CS-ODA-AgNW 29.65 0.305, 852.532, 75.887 0.333 0.99
0.816
INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY 9

Figure 3. (A) XRD data of AgNW, (B) UV-visible absorption spectra of AgNW and Hb solution, (C) TEM
images of Ag-NW, (D) TEM fringe pattern of AgNW.

Figure 4. FTIR spectra of the nanocomposite and its different components in powder form (Panel-I)
and in LB film (Panel-II).
10 M. MAHATO ET AL.

Figure 5. (A-D) SEM images of the composite LB film at different resolution and magnification.

Table 2. Table for peak assignment of FTIR spectra of the nanocomposite.


Sl. No Peak Position Peak assignment Reference
1 1000–1300 C-OH stretching/vibration [58]
2 1523 N-H stretching chitosan [59]
3 1656 N-H Bending [59]
4 2850 CH3 and CH2 bond [60]
5 2917 C-H ant symmetric [60]

3.4. Determination of sensing parameters


ACM pesticide has been chosen as a model pesticide to carry out sensing study due to its
toxicity and versatile uses for killing a wider range of pest in agriculture [33]. In general,
there are three major parameters of a sensor such as sensitivity, LOD and LR. In the
electrochemical method of sensing, the calibration graph is derived from the changes of
major peak current with concentration of the analyte i.e. pesticide. Pt electrodes have
been chosen for electrode modification due to its higher current achieved in both bare
and modified electrode (Figures S1, S2A). Figure 6a, b shows the CV response of the
modified electrode with scan rate variation and its linear fitting. It is found that higher
scan rate results in higher current, which may affect mass transfer of pesticide, and lower
scan rate results in low current leading to lesser sensitivity with the same electrode.
INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY 11

Figure 6. (a) CV response of LB film modified Pt electrode with scan rate variation. (b) Linear plot of
cathodic peak currents vs scan rates. (c) Linear plot of cathodic peak currents vs square root of scan
rates. (d) Linear plot of logarithmic cathodic peak currents vs logarithmic scan rates. (e) Linear plot of
cathodic peak position vs scan rate. (f) Linear plot of anodic-cathodic peak separation vs ln (scan rate).

Hence, it needs to optimise to a moderate scan rate to accommodate both the issues and
is decided optimised scan rate as 100 mV/s as per experimentation (Figure 6b) and as per
reported scan rate for different pesticide sensors (20 mV/s, 50 mV/s, 100 mV/s) [61]. The
optimised scan rate (100 mV/s) has been fixed for all the necessary CV data such as, ACM
sensing (10–100 µM), interference study, real sample analysis and repeatability.
The CV data of modified electrodes (Figure 6a) comprises anodic peak (Epa = 0.102 V),
cathodic peak (Epc = −0.040 V) at scan rate of 0.1 V/s with a peak-to-peak separation (ΔEp)
of 0.155 V. The cathodic peak in case of Hb-ZrO2 composite modified electrode was
12 M. MAHATO ET AL.

reported at −0.37 V, and the observed difference in the peak position may be due to the
difference in composite material [62]. The composite formation of Ag and Hb protein was
also observed in our earlier work [38]. The formal potential (E0) is 0.077 V, which is the
average of the cathodic and anodic peak [63]. The linearity of peak current with scan rate
or (scan rate)1/2 indicates the electrochemical process to be diffusion controlled
(Figure 6b, c) [64]; however, the shifting of peak current position indicates the adsorption
controlled or double layer formation (Figure 6e) [65]. In general, the slope value of the plot
of logarithm of peak current vs scan rate can clarify the process to be purely diffusion
controlled (slope = 0.5), purely adsorption controlled (slope = 1) or a mixture of diffusion
and adsorption (slope in between 0.5 and 1) [66]. The slope value of 0.66 (Figure 6d)
together with cathodic peak shifting (Figure 6e), indicates the process to be adsorption
and diffusion control [65,66]. The anodic peak current to cathodic peak current ratio (ipa
/ipc) is found to be 1.26 i.e. greater than unity (1), which suggests the electrochemical
process to be quasi-reversible process (Figure S3) [67]. The electron transfer rate (Ks) is
found to be 0.2082 ± 0.0032 s−1 using Laviron’s method [68] (Equation 2) and intercept
from Figure 6f. The surface coverage or surface concentration (Γ) of the active agent is
found to be 2.212 × 10−7 mol.cm−2, using slope of Figure 6b and Equation 3 [64]. Here, n is
the number of electrons transferred, υ is the scan rate, ip is the peak current, F is the
Faraday constant, R is the gas constant, T is the temperature, α is the transfer coefficient,
and A is the area of the working electrode.
� �
RT nFΔEp
ln Ks ¼ αlnð1 αÞ þ ð1 αÞ ln α ln αð1 αÞ (2)
nFν RT

ip ¼n2 F2 AΓυ=4RT (3)

The sensor response with the amount of modifier needs to be standardised. In this
direction, we have lifted three different layers of LB nanocomposite film on the Pt
electrode and their CV response has been recorded. It shows the linearity in the cathodic
peak current with the layer number (Figure S2b). It is also to mention that the amount of
material (LB film layer) modified on the Pt electrode is difficult to measure. However, linear
changes with the layer number certainly indicate the equal changes with the equal
amount of electrode modifier. Also, the intrinsic technological advantage of LB trough
allows us to precisely control the number of layers and subsequently amount of electrode
modifier. Hence, LB film modified electrode sensors can be made standard and reliable
with repeatable peak current response.
Figure 7 (a,c,e) shows the CV, DPV and SWV data of the modified electrode with varying
concentration of ACM pesticide. Figure 7(b,d,f) shows the calibration curve of the
reported sensor using CV, SWV and DPV techniques. The CV signal variation with pesticide
concentration is less linear in anodic peak current (R2 = 0.81) (Figure S4) in comparison to
the cathodic peak current (R2 = 0.98) (Figure 7b), which is evidenced in the R2 value of the
linearity fitting. Figure S5a shows the changes of cathodic peak (−0.040 V) with ACM
concentration. Figure 7b shows the CV calibration curve for the reported sensor, which is
derived from cathodic peak current variation with ACM concentration. The sensitivity of
the sensor has been calculated using slope of the calibration graph and the working area
of the electrode (Equation 4) [69]. The sensitivity of the reported sensor using CV, DPV and
INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY 13

Figure 7. (a) CV data of composite LB film modified Pt electrode with concentration of pesticide in the
nM range. (b) CV peak current vs. pesticide concentration and its linear fitting. (c) SWV data of
composite LB film modified Pt electrode with concentration of pesticide in the nM range. (d) SWV peak
current vs. pesticide concentration and its linear fitting. (e) DPV data of composite LB film modified Pt
electrode with concentration of pesticide in the nM range. (f) Peak current vs. pesticide concentration
and its linear fitting.

SWV was found to be 0.418 µA/nM/cm2, 0.259 µA/nM/cm2 and 0.271 µA/nM/cm2, respec­
tively, which is a single reported data till date on ACM pesticide using electrochemical
techniques.

Slope of the calibration curve


Sensitivity ¼ (4)
Working area of electrode
14 M. MAHATO ET AL.

The area of the Pt working electrode was calculated using slide callipers and was found to
be 0.2826 cm2. It is to be noted that the area including PTFE coating is 0.3455 cm2 and this
entire area is being modified by LB method. However, PTFE is insulating in nature and the
working area (0.2826 cm2) will be functional only and is considered as the area of modified
Pt electrode. The LOD has been calculated using slope and residual standard deviation
(RSD) of the calibration curve (Equation 5) [70]. The LOD using CV, DPV and SWV is found
to be 14 nM, 5 nM and 10 nM, respectively, which is near the comparable detection limit
of ACM pesticide by heavy instruments like GC-MS, TLC, HPLC and UV-visible spectro­
scopy [71–74].

3:3 � SD
LOD ¼ (5)
Slope

The LR is the range of sensing parameters (such as concentration of pesticide), where


the response signal differs by maximum 5% or the response signal is linearly propor­
tional with tolerable R2 value of ≥0.95 [70,75]. The LR value of the reported sensor using
CV, DPV and SWV techniques has been calculated by the linear fitting of the signal
response with concentration of ACM pesticide (Figure 7b,d,f). The LR value is found as
10–100 nM (CV), 10–40 nM and 50–100 nM (DPV) and 10–100 nM (SWV) with R2 value of
the linear fitting as 0.98, 0.96 and 0.97 respectively. In DPV technique, the LR needed to
be split into two ranges (R2 = 0.96) to increase the R2 value from the single range
(R2 = 0.85).
Figure S5b shows the CV stability data of the reported sensor upto 20 number of cycles.
The peak current gradually changes in every cycle and attains a saturation. The increase in
peak indicates the population of active sites increases on the Pt electrode and the
decrease in peak current indicates the passivation of the surface of the electrode [76].
The stability of the ACM sensor has been calculated using Equation 6 [76], and the stability
value of 74% indicates that the signal response of the modified electrode to be stable.

Peak current of last cycle


Stability ¼ � 100 (6)
Peak current of first cycle

The reproducibility and repeatability of the reported sensor have been carried out
(Figure S6 a, b) using three numbers of LB film modified electrodes. It is found that RSD
from three electrode responses is 1.5%, which represents the reported sensor to be
reproducible. Figure S6 c, d shows a repeatability graph by the CV response for 20 cycles
and their RSD is found to be 8%. This indicates the sensor developed by LB film modified
electrode can give reproducible data as per the experimental data and also from the fact
that the LB film thickness can be controlled by the layer number.
In literature, pesticide sensing applications have been carried out using different
electrochemical methods such as CV, SWV and DPV [77,78]. CV is widely used as it
provides essential information, such as the process reversibility and types of redox
processes present in the analysis (matrix, analyte and electrodes) [79]. Whereas, the
pulse voltammetry techniques such as DPV and SWV are very sensitive, often allowing
direct analyses of analyte at the ppb (parts per billion) level and even the low ppt (parts
per trillion) [80]. DPV technique is comparably slow compared to CV [81]. Kalinke et al.
prepared a dopamine sensor using CV, DPV and SWV techniques and found the R2 value
INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY 15

Figure 8. (A) Cathodic peak current of LB film modified Pt electrode in presence of different interfering
agents and in presence of 15 nM ACM pesticide. (B) Comparison of cathodic peak current of LB film
modified Pt electrode in ideal distilled water and in real samples (chilli, tap water) and in presence of
15 nM ACM pesticide.

of their calibration curve to be 0.98, 0.99 and 0.99, respectively [82,83]. The reported
sensor achieve higher sensitivity in CV (0.418 µA/nM/cm2) than DPV (0.259 µA/nM/cm2)
and SWV (0.271 µA/nM/cm2) and broader LR (10–100 nM).

3.5. Interference studies and real sample analysis


The interference studies of the reported ACM pesticide sensor have been carried out
in presence of 2 µM metal ions (Zn2+, Cu2+, Ni2+, Pb2+, Al3+, Mg2+) as well as 2 µM
organic interfering species (ascorbic acid, aspartic acid, glutamic acid, citric acid,
glycine) along with 15 nM of ACM pesticide (Figure 8a). It is found that the inter­
fering species could change the CV response of LB film modified Pt electrode only
upto a little extent. The overall RSD of the sensor peak current response in the
presence of metal ions and organic interfering species are within 2%, which reveals
the good consistency of sensor response in the presence of interfering agents
(Figure 8a). Also, the RSD value in peak potential response in the presence of
interfering species is found to be within 2.4%, which is a complementary evidence
in the interference study (Figure S7).
The reported ACM pesticide sensors have been tested for real sample analysis in the
case of chilli and tap water in the presence of 15 nM ACM pesticide (Figure 8b). The
ACM pesticides have been widely used in chilli farming, and hence, it is used in real
sample analysis. The real sample analysis data have been compared with standard
distilled water results and found the RSD in case of chilli and tap water as 1.4% and
0.2%, respectively.
To the best of our knowledge, there are no literature reports on the ACM pesticide
sensing by electrochemical method exclusively by using LB nanocomposite film unlike
the large size instruments like GC-MS, HPLC. Table 3 shows the results of the reported
ACM pesticide sensor as well as detection results from other heavy instruments.
16 M. MAHATO ET AL.

Table 3. Sensing Data of the Reported ACM Sensor and Literature Data using Different Techniques.
Repeatability
Sl. No Technique Adopted LR with R2 Value LOD Sensitivity (RSD) Reference
1 Electrochemical (CV) 10–100 nM (0.98) 14 nM 0.418 µA/nM/cm2 8% Present work
2 DPV 10–40 nM (0.96) and 5 nM 0.259 µA/nM/cm2 2.7%
50–100 nM (0.96)
3 SWV 10–100 nM (0.97) 10 nM 0.271 µA/nM/cm2 2.5%
4 GC-MS 0.37–100 µg L−1 0.12 µg L−1 NR NR [71]
5 TLC NR 0.1 µgL−1 NR NR [72]
−1
6 UV-Visible Spectrometry NR 0.06 µgL NR NR [73]
7 HPLC 1–200 µg L−1 0.21 µgL−1 NR NR [74]
*CV = Cyclic Voltammetry, DPV = Differential Pulse voltammetry, SWV = Square wave voltammetry, GC-MS = Gas
chromatography-Mass Spectroscopy, TLC = Thin Layer Chromatography, HPLC = High-Performance Liquid
Chromatography, NR = Not Recorded.

3.6. ACM Pesticide interaction with nanocomposite


The nanocomposite used for electrode modification comprises haemoglobin, PVP-AgNW,
octadecylamine and chitosan. Haemoglobin has been used due to its direct electron
transfer capacity between Hb and electrode [84], which is wrapped by the suitable
biocompatible chitosan microenvironments [85]. The ACM-like organophosphate pesti­
cides contain sulphur, phosphorus or nitrogen groups, which are easy to adsorb on the
surface of AgNPs via covalent interaction [86]. Also, the reducing group of −OH or −NH in
ACM pesticides readily reacts with the oxygen-containing haemoglobin [87].
The interaction between nanocomposite and ACM pesticide has been explored experi­
mentally using CV (Figure 9a, b) and UV-visible spectroscopy (Figure 9c, d). It is found that
the cathodic peak position is shifting, and the peak current is decreasing with ACM
concentration. Finally, the cathodic peak becomes flat in the µM range of ACM, which is
the reflection of denaturation and destruction of haemoglobin due to µM concentration
of ACM pesticide. The denaturation of haemoglobin is also evident in the UV-visible
spectroscopic data by the abrupt shifting of the Soret band position in the µM range of
ACM (Figure 9d) [37]. In conclusion, the interaction of ACM existed with haemoglobin and
may be with AgNW through adsorption [37].

4. Conclusion
An electrochemical pesticide sensor has been developed to detect a hazardous ACM
pesticide using CV, DPV and SWV technique with sensitivity 0.418 µA/nM/cm2, 0.259 µA/
nM/cm2 and 0.271 µA/nM/cm2 using a novel chitosan-silver nanowire-based nanocom­
posite LB film modified electrode. To the best of our knowledge, the sensing study of the
ACM pesticide has not yet been explored in literature using nanocomposite matrix unlike
the big size costly instrument. The formation of the composite (OCPAH) was studied by
the surface pressure isotherm using LB technique as well as SEM and FTIR technique. The
achieved sensing parameters such as LOD as 14 nM, 10 nM and 5 nM; LR as (10–100 nM,
10–40 nM and 50–100 nM and 10–100 nM), respectively, for CV, DPV and SWV techniques.
The CV showed the broad LR (10–100 nM) with R2 = 0.98, high sensitivity (0.418 µA/nM/
cm2) compared to SWV and DPV techniques. The results were found to be promising as
INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY 17

Figure 9. (a) CV (cathodic peak current) response of LB film modified Pt electrode with ACM
concentration (nM to µM). (b) Changes in cathodic peak current and peak position with concentration
of ACM. (c) Normalised UV-visible spectra of haemoglobin with concentration of ACM, (d) Change in
haemoglobin Soret band position with concentration of ACM.

compared with other existing methods. The present study opens up new research options
on high sensitive sensor development for other different kinds of hazardous organopho­
sphorus pesticide using LB film-based nanocomposite modified electrodes.

Acknowledgements
Authors acknowledge DST-SERB, Government of India, for financial support through SERB projects
(No. EMR/2016/002634) sanctioned to Dr. Mrityunjoy Mahato. Sanjeev Bhandari thanks SERB for
providing the Non-Net JRF fellowship. We also thank SAIF (NEHU) for providing TEM Facilities. Also,
we thank the Department of Nanotechnology, NEHU and Department of Chemistry, Guwahati
University for cyclic voltammetry, XRD and FE-SEM characterisation.

Disclosure statement
No potential conflict of interest was reported by the author(s).

Funding
This work was supported by the Science and Engineering Research Board [EMR/2016/002634].
18 M. MAHATO ET AL.

ORCID
Manashjit Gogoi http://orcid.org/0000-0002-1753-192X

References
[1] P. Kumar, K.H. Kim and A. Deep, Biosens. Bioelectron. 70, 469 (2015). doi:10.1016/j.
bios.2015.03.06.
[2] E.-C. Oerke and H.-W. Dehne, Crop Prot. 23 (4), 275 (2004). doi:10.1016/j.cropro.2003.10.001.
[3] M. Eddleston, N.A. Buckley, P. Eyer and A.H. Dawson, The Lancet 371 (9612), 597 (2008).
doi:10.1016/S0140-6736(07)61202-1.
[4] T. Bhardwaj and J.P. Sharma, Int. J. Agric. Food Sci. Technol. 4, 817 (2013).
[5] Ministry of Agriculture & Farmers Welfare, Government of India, Annual Report 2016-17,
(2017).
[6] D. Pimentel and M. Burgess, Environ. Dev. Sustain. 14 (2) (2012). doi:10.1007/s10668-011-
9325-5.
[7] S. Yadav and G. Gaba, J. Appl. Nat. Sci. 10 (2), 540 (2018). doi:10.31018/jans.v10i2.1732.
[8] J. Paull, Nanomaterials in Food and Agriculture: The Big Issue of Small Matter for Organic
Food and Farming, Proceedings of the 3rd Scientific Conference International Society of
Organic Agriculture Research, Nanyangju, Korea (2011).
[9] B. Tang, J.E. Zhang, L.G. Zhang, Y.Z. Zang, X.Y. Li and L. Zhou, Journal of Chromatographic
Science 43 (7), 337 (2005). doi:10.1093/chromsci/43.7.337.
[10] C.-W. Lee, Y.-Y. Chao, J. Shiea, J.-H. Shen, H.-H. Lee and B.-H. Chen, Am. J. Emerg. Med. 36,
(2018) DOI:10.1016/j.ajem.2017.12.042.
[11] Price Report of Pesticide from Conquer Scientific Website, www.conquerscientific.com/lab-
equipment/mass-spectrometers-gcms-systems/, (Accessed on 6th Nov, 2019).
[12] X. Yan, H. Li and X. Su, Trends Anal. Chem. 103 (1), 1 (2018). doi:10.1016/j.trac.2018.03.004.
[13] W. Wang, X. Wang, N. Cheng, Y. Luo, Y. Lin, W. Xu and D. Du, Trends Anal. Chem. 132, 116041
(2020). doi:10.1016/j.trac.2020.116041.
[14] I.S. Kucherenko, O.O. Soldatkin, D.Y. Kucherenko, O.V. Soldatkina and S.V. Dzyadevych,
Nanoscale Adv. 1 (12), 4560 (2019). doi:10.1039/C9NA00491B.
[15] A. Masotti and G. Ortaggi, Mini-Rev. Med. Chem. 9 (4), 463 (2009). doi:10.2174/
138955709787847976.
[16] K. Chen, J. Li, L. Zhang, R. Xing, T. Jiao, F. Gao and Q. Peng, Nanotechnology 29 (44), 445603
(2018). doi:10.1088/1361-6528/aadbf7.
[17] S. Su, H. Sun, F. Xu, L. Yuwen, C. Fan and L. Wang, Microchim. Acta 181 (13–14), 1497 (2014).
doi:10.1007/s00604-014-1178-9.
[18] S.J. Percival and B. Zhang, J. Phys. Chem. C 120 (37), 20536 (2016). doi:10.1021/acs.
jpcc.5b11330.
[19] N. Elgrishi, K.J. Rountree, B.D. McCarthy, E.S. Rountree, T.T. Eisenhart and J.L. Dempsey,
J. Chem. Educ. 95 (197), 197 (2018). doi:10.1021/acs.jchemed.7b00361.
[20] E. Sinkala, J.E. McCutcheon, M.J. Schuck, E. Schmidt, M.F. Roitman and D.T. Eddington, Lab
Chip 12 (13), 2403 (2012). doi:10.1039/C2LC40168A.
[21] H. Parham, N. Rahbar and J. Hazard, Mater 177, 1077 (2010). doi:10.1016/j.
jhazmat.2010.01.031.
[22] J. Sun, L. Guo, Y. Bao and J. Xie, Biosens. Bioelectron. 28 (1), 152 (2011). doi:10.1016/j.
bios.2011.07.012.
[23] Q. Zhou, L. Yang, G. Wang and Y. Yang, Biosens. Bioelectron. 49 (25) (2013). doi:10.1016/j.
bios.2013.04.037.
[24] N. Jha and S. Ramaprabhu, Nanoscale 2 (5), 806 (2010). doi:10.1039/B9NR00336C.
[25] J.T. Martins, M.A. Cerqueira and A.A. Vicente, Food Hydrocoll. 27 (1), 220 (2012). doi:10.1016/j.
foodhyd.2011.06.011.
INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY 19

[26] F.J. Pavinatto, L. Caseli and O.N. Oliveira, Biomacromolecules 11 (8), 1897 (2010). doi:10.1021/
bm1004838.
[27] D. Dua, X. Huanga, J. Cai and A. Zhang, Sens. Actuators B Chem. 127 (2), 531 (2007).
doi:10.1016/j.snb.2007.05.006.
[28] J. Yu, H. Guan and D. Chi, J. Solid State Electrochem. 21 (4), 1175 (2017). doi:10.1007/s10008-
016-3468-0.
[29] A.K.M. Kafi, Q. Wali, R. Jose, T.K. Biswas and M.M. Yusoff, Microchim. Acta 184 (11), 4443 (2017).
doi:10.1007/s00604-017-2479-6.
[30] W. Wang, T. He, X. Liu, W. He, H. Cong, Y. Shen, L. Yan, X. Zhang, J. Zhang and X. Zhou, ACS
Appl. Mater. Interfaces 8 (32), 20839 (2016). doi:10.1021/acsami.6b08091.
[31] T.C. Kwong, Ther. Drug Monit. 24 (1), 144 (2002). doi:10.1097/00007691-200202000-00022.
[32] S.O. Obare, C. De, W. Guo, T.L. Haywood, T.A. Samuels, C.P. Adams, N.O. Masika, D.H. Murray,
G.A. Anderson, K. Campbell and K. Fletcher, Sensors 10 (7), 60 (2010). doi:10.3390/
s100707018.
[33] J.B. Knaak, C.C. Darry, X. Zhang, R.W. Gerlach, R.T. Velez, D.T. Chang, R. Goldsmith and J.
N. Blancato, Rev. Environ. Contam. Toxicol. 219 (1) (2012). doi:10.1007/978-1-4614-3281-4_1.
[34] G. Jyot, K. Mandal, R.S. Battu and B. Singh, Environ. Monit. Assess. 185 (7), 5703 (2013).
doi:10.1007/s10661-012-2977-2.
[35] B.B. Majumdar, G. Guha, A.N. Ray and B. Bala, Ann. Trop. Med. Public Health 5 (6), 615 (2012).
doi:10.4103/1755-6783.109330.
[36] M. Mahato, P. Pal, T. Kamilya, R. Sarkar and G.B. Talapatra, J. Phys. Chem. B 114 (1), 495 (2010).
doi:10.1021/jp908081r.
[37] M. Mahato, P. Pal, T. Kamilya, R. Sarkar, A. Chaudhuri and G.B. Talapatra, J. Phys. Chem. B
114 (20), 7062 (2010). doi:10.1021/jp100188s.
[38] M. Mahato, P. Pal, B. Tah, M. Ghosh and G.B. Talapatra, Colloids Surf. B 88 (1), 141 (2011).
doi:10.1016/j.colsurfb.2011.06.024.
[39] H. Li, G. Mao and K.Y.S. Ng, Thin Solid Films 358 (1–2), 62 (2000). doi:10.1016/S0040-6090(99)
00661-6.
[40] I. Ahmed, L. Dildar, A. Haque, P. Patra, M. Mukhopadhyay, S. Hazra, M. Kulkarn, S. Thomas, J.
R. Plaisier, S.B. Dutta and J.K. Bal, J. Colloid Interface Sci. 514, 433 (2018). doi:10.1016/j.
jcis.2017.12.037.
[41] M. Mahato, P. Pal, B. Tah and G.B. Talapatra, Colloid Surf. A 414, 375 (2012). doi:10.1016/j.
colsurfa.2012.08.064.
[42] M. Mahato, R. Sarkar, P. Pal and G.B. Talapatra, Indian J. Phys. 89 (10), 997 (2015). doi:10.1007/
s12648-015-0674-z.
[43] M. Mahato, P. Pal, T. Kamilya, R. Sarkar, A. Chaudhuri and G.B. Talapatra, Phys. Chem. Chem.
Phys. 12 (40), 12997 (2010). doi:10.1039/c0cp00344a.
[44] S. Joy, P. Pal, M. Mahato, G.B. Talapatra and S. Goswami, Dalton Trans. 39 (11), 2775 (2010).
doi:10.1039/b919871g.
[45] D. Kumar, H. S. Bhatti, H. S. Bhatti and H.S. Bhatti, Appl. Nanosci. 5 (7), 881 (2015). doi:10.1007/
s13204-014-0386-2.
[46] K.E. Korte, S.E. Skrabalak and Y. Xia, J. Mater. Chem. 18 (4), 437 (2008). doi:10.1039/b714072j.
[47] M.J. Song, S.W. Hwang and D. Whang, J. Appl. Electrochem. 40 (12), 2099 (2010). doi:10.1007/
s10800-010-0191-x.
[48] S. Ma, W. Lu, J. Mu, F. Liu and L. Jiang, Colloids Surf. A: Physicochem. Eng. Aspects 324 (1–3), 9
(2008). doi:10.1016/j.colsurfa.2008.03.029.
[49] A.J. Marenco, D.B. Pedersen, S. Wang, M.W.P. Petryk and H.B. Kraatz, Analyst 134 (10), 2021
(2009). doi:10.1039/b909748a.
[50] M. Maillard, S. Giorgio and M.P. Pileni, Adv. Mater 14 (15), 1084 (2002). doi:10.1002/1521-
4095(20020805)14:15<1084::AID-ADMA1084>3.0.CO;2-L.
[51] M. Chen, C. Wang, X. Wei and G. Diao, Phys. Chem. C 117 (26), 13593 (2013). doi:10.1021/
jp404563h.
[52] Y. Sun, B. Mayers, T. Herricks and Y. Xia, Nano Lett. 3 (7), 955 (2003). doi:10.1021/nl034312m.
20 M. MAHATO ET AL.

[53] L. Cui, Z. Du, W. Zou, H. Li and C. Zhang, RSC Adv. 4 (52), 27591 (2014). doi:10.1039/
C4RA02691H.
[54] L. Jin and R. Bai, Langmuir 18 (25), 9765 (2002). doi:10.1021/la025917l.
[55] Y. Wang, L. Shi, L. Gao, Q. Wei, L. Cui, L. Hu, L. Yan and B. Du, J. Colloid Interface Sci. 451 (7)
(2015). doi:10.1016/j.jcis.2015.03.048.
[56] A.T. Paulino, J.I. Simionato, J.C. Garcia and J. Nozaki, Carbohydr. Polym. 64 (98), 98 (2006).
doi:10.1016/j.carbpol.2005.10.032.
[57] G.W. Huang, N. Li, Y. Liu, C.B. Qu, Q.P. Feng and H.M. Xiao, Appl. Mater. Interfaces 11 (16),
15028 (2019). doi:10.1021/acsami.8b22053.
[58] J. Gong, W. Zhang, T. Liu and L. Zhang, Nanoscale 3 (8), 3123 (2011). doi:10.1039/c1nr10286a.
[59] D. Du, J. Ding, J. Cai and A. Zhang, Colloids Surf. B Biointerfaces 58 (2), 145 (2007).
doi:10.1016/j.colsurfb.2007.03.006.
[60] T. Jeyapragasam and R. Saraswathi, Sens. Actuators B Chem 191, 681 (2014). doi:10.1016/j.
snb.2013.10.054.
[61] L. Wu, W. Lei, Z. Han, Y. Zhang, M. Xia and Q. Hao, Sens. Actuators B Chem. 206, 495 (2015).
doi:10.1016/j.snb.2014.09.098.
[62] S. Liu, Z. Dai, H. Chen and H. Ju, Biosens. Bioelectron. 19 (9), 963 (2004). doi:10.1016/j.
bios.2003.08.025.
[63] H.Y. Zhao, X.X. Xu, J.X. Zhang, W. Zheng and Y.F. Zhen, Bioelectrochemistry 78 (2), 124 (2010).
doi:10.1016/j.bioelechem.2009.08.009.
[64] L. Fotouhi, A.B. Hashkavayi, M.M. Heravi and J. Exp, Nanoscience 8, 947 (2013). doi:10.1080/
17458080.2011.624554.
[65] H.T. Purushothama, Y. Arthoba Nayaka, M.M. Vinay, P. Manjunatha, R.O. Yathisha and K.
V. Basavarajappa, J. Sci-Adv. Mater. Dev. 3, 161 (2018). doi:10.1016/j.jsamd.2018.03.007.
[66] M.T. Cryan and A.E. Ross, Analyst 144 (1), 249 (2019). doi:10.1039/C8AN01547C.
[67] F. Haque, M. Rahman, E. Ahmed, P. Bakshi and A. Shaikh, Dhaka Univ. J. Sci 61 (2), 161 (2013).
doi:10.3329/dujs.v61i2.17064.
[68] E. Topoglidis, Y. Astuti, F. Duriaux, M. Grätzel and J.R. Durrant, Langmuir 19 (17), 6894 (2003).
doi:10.1021/la034466h.
[69] S.R. Balakrishnan, U. Hashim, G.R. Letchumanan, M. Kashif, A.R. Ruslinda, W.W. Liu,
P. Veeradasan, R.H. Prasad, K.L. Foo and P. Poopalan, Sens. Actuators A Phys. 220, 101
(2014). doi:10.1016/j.sna.2014.09.027.
[70] B. Mutharani, P. Ranganathan, S.M. Chen and C. Karuppiah, Microchim. Acta 186 (3) (2019).
doi:10.1007/s00604-018-3206-7.
[71] M. Tankiewicz, C. Morrison and M. Biziuk, Talanta 107 (4), 1 (2013). doi:10.1016/j.
talanta.2012.12.052.
[72] S. Babić, M. Petrović and M.K. Macan, J. Chromatogr. A 823 (1–2), 3 (1998). doi:10.1016/S0021-
9673(98)00301-X.
[73] S. Rahim, S. Khalid, M.I. Bhanger, M.R. Shah and M.I. Malik, Sens. Actuat. B: Chem. 259, 879
(2018). doi:10.1016/j.snb.2017.12.138.
[74] J. Zhang, H. Gao, B. Peng, S. Li and Z. Zhou, J. Chromatogr. A 1218, 6623 (2011). doi:10.1016/j.
chroma.2011.07.102.
[75] L.C. Rodríguez, A.M.G. Campaña and J.M.B. Sendra, Anal. Lett. 29 (7), 1231 (1996). doi:10.1080/
00032719608001471.
[76] J.G. Manjunatha, M. Deraman, N.H. Basri and I.A. Talib, Arabian J. Chem. 11 (2), 149 (2018).
doi:10.1016/j.arabjc.2014.10.009.
[77] D.J.E. Costa, J.C.S. Santos, F.A.C. Sanches-Brandão, W.F. Ribeiro, G.R. Salazar-Banda and M.C.
U. Araujo, J. Electroanal. Chem. 789, 100 (2017). doi:10.1016/j.jelechem.2017.02.036.
[78] J.B. Thakkar, S. Gupta and C.R. Prabha, Int. J. Biol. Macromol. 137, 895 (2019). doi:10.1016/j.
ijbiomac.2019.06.162.
[79] F.R. Simões and M. G. Xavier, in Chapter 6: Electrochemical Sensors in Nanoscience and Its
Applications edited by, A. L. D. Róz, M. Ferreira, F. L. De Leite and O. N (Oxford, United
Kingdom: Oliveira, William Andrew, 2017, 155).
INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY 21

[80] S. Li, C. Zhang, S. Wang, Q. Liu, H. Feng, X. Ma and J. Guo, Analyst 143 (18), 4230 (2018).
doi:10.1039/C8AN01067F.
[81] P.K. Brahman, N. Pandey, J.V.S. Kumar, P. Somarouthu, S. Tiwari and K.S. Pitre, Arabian J. Chem.
9, S1897 (2016). doi:10.1016/j.arabjc.2014.02.003.
[82] C. Kalinke, N.V. Neumsteir, G. de O. Aparecido, T.V. De. Barros Ferraz, P.L. De. Santos, B.
C. Janegitz and J.A. Bonacin, Analyst 145 (4), 1207 (2020). doi:10.1039/C9AN01926J.
[83] S. Pysarevska, L. Dubenska, S. Plotycya and Ľ. Švorc, Sens. Actuators B Chem. 270 (9), 9 (2018).
doi:10.1016/j.snb.2018.05.012.
[84] R. Kaur, S. Rana, K. Lalit, P. Singh and K. Kaur, Biosens. Bioelectron. 167 (1) (2020). doi:10.1016/
j.bios.2020.112486.
[85] R. Ahmad, O.S. Wolfbeis, Y.-B. Hahn, H.N. Alshareef, L. Torsi and K.N. Salama, Mater. Today
Commun. 17, 289 (2018). doi:10.1016/j.mtcomm.2018.09.024.
[86] Y. He, B. Xu, W. Li and Y. Haili, J. Agr. Food Chem. 63, 2932 (2015). doi:10.1021/acs.
jafc.5b00671.
[87] Z.F. Zhang, H. Cui, C.Z. Lai and L. Liu, J. Anal. Chem. 77 (10), 3324 (2005). doi:10.1021/
ac050036f.

You might also like