Uploaded by nilurkut

d1ay00074h

advertisement
Analytical
Methods
Volume 13
Number 11
21 March 2021
Pages 1311–1434
rsc.li/methods
Duc Anh Thai and Nae Yoon Lee
A paper-based colorimetric chemosensor
for rapid and highly sensitive detection
of sulfide for environmental monitoring
I
PAPER
M nde
ed xe
lin d in
e!
ISSN 1759-9679
Analytical
Methods
View Article Online
Published on 09 February 2021. Downloaded by Istanbul Technical University on 5/27/2023 12:46:04 AM.
PAPER
Cite this: Anal. Methods, 2021, 13, 1332
View Journal | View Issue
A paper-based colorimetric chemosensor for rapid
and highly sensitive detection of sulfide for
environmental monitoring†
Duc Anh Thai
and Nae Yoon Lee
*
In this study, we report on paper-based colorimetric detection of sulfide using a newly synthesized
chemical acting as a chemosensor, based on the deprotonation mechanism. Paper strips were also
fabricated and incorporated with the chemosensor for on-site monitoring. The presence of sulfide
induced deprotonation of a hydroxyl group of the chemosensor, which eventually resulted in a distinct
spectral change in the tube as well as a visible color change on a paper strip. The chemosensor showed
a highly selective colorimetric response to sulfide by changing its color from colorless to yellow without
any interference from a mixture containing other anions. Moreover, the chemosensor effectively
differentiated sulfide from other thiols, including cysteine and glutathione. The chemosensor
colorimetrically detected sulfide with a fast response time of 10 s under physiological conditions.
Received 13th January 2021
Accepted 9th February 2021
Practically, the paper test strip enabled colorimetric visualization of as low as 30 mM sulfide and a good
recovery in quantitative analysis in water samples. The introduced paper-based chemosensor is
DOI: 10.1039/d1ay00074h
rsc.li/methods
1
a promising colorimetric strategy with rapid, selective, and sensitive sensing abilities for sulfide
monitoring in environmental water samples.
Introduction
Sulde (S2) is known as a highly toxic substance, recognized by
its rotten egg malodorous smell, and is widely distributed in
natural water and wastewater.1 Sulde appears naturally in
petroleum and hot springs, while some amounts of it are
generated from the decay of organic matter (e.g. human and
animal wastes) by bacterial processes.2 Furthermore, it is also
produced from the industrial production of paper and pulp
mills, chemical fertilizers, manufacturing of sulfuric acid, and
dyes.3 In biological systems, sulde is endogenously produced
from L-cysteine catalyzed by cystathionine-b-synthase, cystathionine-g-lyase, and 3-mercaptopyruvate sulfurtransferase.4–6 Sulde is involved in several physiological functions,
including vascular smooth muscle cell proliferation, apoptosis,
insulin signaling, and oxygen sensing.7–9 However, abnormal
concentrations of sulde in bio-systems are associated with
serious health problems such as Alzheimer's disease, Down's
syndrome, liver cirrhosis, heart disease, and diabetes.10–12 Due
to the considerable interest in sulde, there is an urgent need
for the development of a novel approach for fast and sensitive
detection of sulde.
Department of BioNano Technology, Gachon University, 1342 Seongnam-daero,
Sujeong-gu, Seongnam-si, Gyeonggi-do, 13120, Korea. E-mail: nylee@gachon.ac.kr
† Electronic supplementary information (ESI) available: Experimental details,
characterization and spectral studies. See DOI: 10.1039/d1ay00074h
1332 | Anal. Methods, 2021, 13, 1332–1339
Gas chromatography and high-performance liquid chromatography techniques have been known as gold standard
methods for the detection and determination of sulde in
various samples, including environmental water and soil
samples.13 While chromatographic strategies possess the
advantages of separation capability and reliability, these techniques require expensive equipment, sophisticated training for
operation, high operating cost, and time-consuming analysis.14
In addition, although the methylene blue test is the most
common approach for the spectroscopic analysis of sulde,15
this method is necessarily catalyzed by concentrated HCl and is
poorly selective towards other thiol compounds.16 It is thus
necessary to develop a simple and quick system that enables the
selective detection of sulde. Recently, due to such desirable
properties such as sensitivity, ease of operation, and real-time
detection, chemosensors have emerged as an attractive method.
Indeed, various sulde sensors and their applications have been
explored following some important sensing mechanisms,
including the reduction of azides to amines,17,18 nucleophilic
reactions,19,20 and displacement of metal ions by sulde.21,22
However, most of the reported sensors still have drawbacks of
complicated synthesis and interference by other thiols, such as
cysteine and glutathione.23 In addition, most previous chemosensors are based on uorescence signaling, which still requires
sophisticated excitation and read-out instruments, which
makes these methods unsuitable for on-site monitoring of
sulde.
This journal is © The Royal Society of Chemistry 2021
View Article Online
Published on 09 February 2021. Downloaded by Istanbul Technical University on 5/27/2023 12:46:04 AM.
Paper
Considering these limitations, a deprotonation-based
colorimetric chemosensor has been consequently developed for
a fast response and non-interference detection of sulde.24,25
Moreover, chemosensors containing Schiff's base have been
particularly attractive due to their chromophoric characteristics
and high stability. A very limited number of Schiff's base chemosensors for sulde, based on deprotonation, have recently
been reported. Although these reported systems perform well,
they still require multiple syntheses. Kaushik et al.26 reported an
azine based chemosensor for the detection of sulde in biological uids such as human serum and mouse serum. Another
colorimetric chemosensor was proposed by Ryu et al.27 to detect
multiple targets: copper(II), cyanide, and sulde. Furthermore,
colorimetry provides a straightforward and convenient strategy
for developing paper-based sensor. Paper-based colorimetric
detection has become a popular strategy due to its advantages
such as portability, fast response, and minimal instrumentation. Most of all, the development of colorimetric paper strips
eliminates the requirement of light sources and enables signal
determination by the naked eye in the form of color change.28
This is an advantage for an in situ analysis with limited equipment in a low-resource environment.
Herein, we introduce a sensitive paper-based colorimetric
chemosensor for the rapid detection of sulde. By performing
absorption spectroscopy, we characterized various sensing
properties of the chemosensor, such as selectivity, sensitivity,
pH effect, stability, and reversibility. Furthermore, the colorimetric chemosensor was incorporated in designed paper strips
for rapid and selective visualization of sulde without sophisticated equipment. Moreover, we demonstrated the practicality
of the colorimetric chemosensor for quantitatively monitoring
sulde in water sample as a proof-of-concept test.
2 Experimental section
2.1
Materials and instrumentation
All reagents and solvents were of analytical grade and used
without further purication. Sulfanilamide, 2,3-dihydroxybenzaldehyde, sodium sulde, Bis–Tris, all anions in the form
of tetrabutylammonium and sodium salts and amino acids
(cysteine and glutathione) were purchased from Sigma-Aldrich,
Korea. Dimethyl sulfoxide (DMSO) solvent was supplied by
Biosesang (Seongnam, Korea). All solutions and buffers were
prepared in deionized water. Glass microber lters were
supplied by Chmlab Group (Barcelona, Spain). Proton Nuclear
Magnetic Resonance (1H NMR) spectra were recorded on a JEOL
JNM-LA400 instrument (JEOL Ltd., USA) at 400 MHz. Liquid
Chromatography-Mass Spectrometry (LC-MS) with the electrospray ionization (ESI) technique was performed with an API
3200™ system (AB Sciex Pte. Ltd., USA). Fourier-transform
infrared spectroscopy (FTIR) peaks were recorded on a JASCO
FT/IR-4600 spectrometer (JASCO International Co., Ltd., Japan).
UV-Vis absorption and uorescence spectra were recorded
using an Epoch™ Microplate and Synergy™ H1 Hybrid MultiMode spectrophotometer (BioTek Instruments, Inc., USA),
respectively. The pH level of a solution was measured using
a digital pH meter Istek pH-200L (Istek, Inc., Korea).
This journal is © The Royal Society of Chemistry 2021
Analytical Methods
2.2
Synthesis of chemosensor L
The chemosensor L was prepared by a reaction of 2,3-dihydroxybenzaldehyde (1 mmol) and sulfanilamide (1 mmol) in
absolute ethanol. The reaction mixture was stirred at room
temperature for 24 h. Then, the tangerine orange precipitate
was ltered, washed several times with cold ethanol, and dried
in a vacuum. The chemosensor L was characterized by standard
analytical techniques including mass spectrometry, 1H NMR
spectroscopy,
FTIR
spectroscopy,
and
UV-Vis
spectrophotometry.
2.3
General analytical procedure
Chemosensor L stock was initially prepared in DMSO, and the
solution was then diluted with the DMSO : Bis–Tris buffer (4 : 6,
10 mM, pH 7.0) solution to make a nal concentration of 50 mM.
All anions such as uoride (F), chloride (Cl), bromide (Br),
iodide (I), azide (N3), nitrite (NO2), nitrate (NO3), sulfate
(SO42), cysteine (Cys), glutathione (GSH), and sulde (S2)
were prepared in deionized water and the test solutions were
mixed in an Eppendorf tube. The spectral measurements were
recorded at room temperature for further analysis. The detailed
procedures for selective, competitive, sensitive, pH-dependent
and time-dependent, and reversible tests are described in the
ESI.†
2.4
Theoretical calculations
To understand the sensing mechanism of chemosensor L,
theoretical computation was performed using density functional theory (DFT) calculations. The optimized structures of
the ground state for chemosensor L with and without sulde
were generated using GaussView embedded in the Gaussian 09
program.29 DFT calculations were evaluated and optimized with
the B3LYP/6-31G* basis set.30 Moreover, the frontier molecular
orbital contributions including the HOMO (highest occupied
molecular orbital) and LUMO (lowest unoccupied molecular
orbital) and the corresponding energies of chemosensor L with
and without sulde were also simulated.
2.5 Fabrication of paper strips and detection of sulde in
water samples
The designed paper strip was fabricated using lter paper on
which chemosensor L was coated by immersion, as shown in
Fig. 1. The strip was rst designed in AutoCAD soware, and
then a glass microber lter (Chmlab Group, Spain) was used to
fabricate the paper strip by laser cutting. Next, the prepared
strip was immersed in the DMSO solution of chemosensor L
(100 mM) for 5 min allowing the chemosensor L to homogeneously distribute along the substrate. The substrate was then
dried in a vacuum to obtain a ready-to-use paper strip, which
could be used for the recognition of sulde. For water sample
analysis, water samples from laboratory taps were collected in
glass bottles. To mimic the sulde contaminated water
samples, the tap water samples were spiked with various
concentrations of sulde ranging from 200 to 800 mM. The
chemosensor L was prepared in Bis–Tris buffer as optimized
Anal. Methods, 2021, 13, 1332–1339 | 1333
View Article Online
Published on 09 February 2021. Downloaded by Istanbul Technical University on 5/27/2023 12:46:04 AM.
Analytical Methods
Paper
Fig. 1 Fabrication process of the paper strip coated with chemosensor
L for the visualization of sulfide.
conditions for performing a real sample test. For sample analysis, different sulde-spiked water samples were added to chemosensor L solution to perform quantitative measurements.
The UV-Vis absorption was monitored at the 400 nm wavelength, and the experiments were performed four times and
averaged for further calculation.
3 Results and discussion
3.1
Characterization of the synthesized chemosensor L
Generally, there are three forms of sulde that exist in aqueous
solution including hydrogen sulde (H2S), bisulde (HS), and
sulde anions (S2). Under physiological conditions, approximately one-h exists as H2S and the remainder largely as its
anions such as HS and S2.26 In this study, a mixture of H2S,
HS, and S2 in solution was referred to as “sulde”. A chemosensor L was designed and developed for application to detect
sulde in an aqueous solution. Chemosensor L ((Z)-4-((2,3-dihydroxybenzylidene)amino)benzenesulfonamide) was synthesized
with a 56% yield, from 2,3-dihydroxybenzaldehyde and sulfanilamide with a 1 : 1 molar ratio (Scheme 1). The mass of chemosensor L was determined from the dominant peak at 293.3 m/z in
the mass spectra (Fig. 2a). The content of C, H, N, O, and S
elements in chemosensor L was found to be 53.42%, 4.14%,
9.58%, 21.89%, and 10.97%, respectively. The proposed chemical
structure of chemosensor L is illustrated in Fig. 2b. Fig. 2c shows
Scheme 1
Schematics of the synthesis of chemosensor L.
1334 | Anal. Methods, 2021, 13, 1332–1339
Structural characterization of chemosensor L. (a) Liquid chromatography-mass spectra (LC-MS) of chemosensor L. (b) Proposed
chemical structure of chemosensor L. (c) 1H NMR of chemosensor L in
DMSO-d6. (d) FTIR spectra of chemosensor L.
Fig. 2
the 1H NMR chemical shis of chemosensor L using DMSO-d6
solvent. The peak at 8.92 ppm corresponded to benzylidene HC]
N, while the proton attached to hydroxyl groups was featured at
5.77 ppm.31 The carbon next to the hydroxyl group carried the
proton featured as a peak at 6.86–6.76 ppm. The peaks at 6.56–
6.54 and 6.96–6.94 ppm were related to other protons of the
benzylidene moiety. In addition, the peaks at 7.81–7.85, 7.54–
7.52, 7.41–7.37, and 7.12–7.10 ppm corresponded to the aromatic
protons of the benzenesulfonamide moiety.32 Furthermore, the
IR band resulting from the NH2 group of the chemosensor L was
observed at approximately 700 cm1. The sharp band at 1016
cm1 and doublet bands at around 1403 cm1 represented O]
S]O symmetric and HC]N vibrations, respectively. Additionally, the IR spectra showed a band at 3425 cm1, which was
related to aromatic O–H vibration (Fig. 2d).31 The typical aromatic
hydroxyl groups, like phenols, have been well known for being
deprotonated in the presence of basic anions.
3.2
Spectral properties and sensitivity of chemosensor L
The spectral properties of chemosensor L were examined by
both UV-Vis and uorescence titration towards different
concentrations of sulde in the DMSO: Bis–Tris buffer (4 : 6, 10
mM, pH 7.0) solution. Fig. 3a shows the UV-Vis absorption
spectra of chemosensor L in the wavelength range from 300 to
600 nm. The free chemosensor L displayed no certain absorption band; however, the solutions showed a gradual bathochromic shi in the successive addition of sulde. An
absorption peak at 400 nm appeared with the addition of
sulde, and the absorbance at 400 nm signicantly increased as
the concentration of sulde increased. Sulde also induced
a distinct color change of chemosensor L from colorless to
yellow (inset gure). In addition, the absorbance at 400 nm was
plotted as a function of sulde concentration. As a result,
This journal is © The Royal Society of Chemistry 2021
View Article Online
Published on 09 February 2021. Downloaded by Istanbul Technical University on 5/27/2023 12:46:04 AM.
Paper
Analytical Methods
Fig. 4 Results of selectivity of chemosensor L to sulfide. Absorbance
(a) and fluorescence intensity (b) change of chemosensor L (50 mM) in
the presence of various anions. The corresponding visible color (c) and
fluorescence color (d) change of chemosensor L (50 mM) in the
presence of various anions (60 equivalents) in the DMSO: Bis–Tris
buffer (4 : 6, 10 mM, pH 7.0) solution.
UV-Vis spectra (a) and emission spectra (c) of chemosensor L
(50 mM) in the presence of different concentrations of sulfide (0–60
equivalents). The inset figures show the visible and fluorescence color
change of chemosensor L after reacting with sulfide in the DMSO: Bis–
Tris buffer (4 : 6, 10 mM, pH 7.0) solution. Absorbance (b) and fluorescence intensity (d) plotted as a function of sulfide concentrations.
Fig. 3
a linear equation was obtained with the coefficient of determination R2 ¼ 0.9996. The limit of detection (LOD) was determined to be 25.8 mM (Fig. 3b).
Interestingly, during the experiment, we found that chemosensor L showed a uorescence response to sulde. The chemosensor L initially showed a non-uorescence signal when
excited at 400 nm. However, when sulde was introduced into
the chemosensor L solution, a new broad uorescence band
emerged at around 570 nm (Fig. 3c). The uorescence emission
was also recognized by orange uorescence under 365 nm UV
light (inset gure). Consequently, the emission intensity gradually enhanced with an increase of sulde concentration with
a detection limit of 15.6 mM (R2 ¼ 0.9843) (Fig. 3d). The LOD was
considerably lower than a secondary maximum contaminant
level set for sulde in water (7.8 mM) as established by the US
Environmental Protection Agency (EPA).33 Therefore, chemosensor L showed a strong colorimetric and uorescence signal,
which could be efficiently used for the sensitive detection of
sulde in environmental systems.
3.3
The selectivity of chemosensor L for sulde detection
Fig. 4 shows the selectivity of chemosensor L to sulde (S2)
among various anions, such as uoride (F), chloride (Cl),
bromide (Br), iodide (I), azide (N3), nitrite (NO2), nitrate
(NO3), sulfate (SO42), cysteine (Cys), and glutathione (GSH),
the nal two of which contained thiol. Chemosensor L (50 mM)
was essentially colorless and non-uorescent in the aqueous
solution; however, chemosensor L exhibited an absorption peak
at 400 nm in response to sulde upon the addition of each
anion. Aer the addition of sulde to chemosensor L, the
absorbance increased approximately 5.4-fold compared with
that of chemosensor L (Fig. 4a). In contrast, for other anions
This journal is © The Royal Society of Chemistry 2021
chemosensor L showed almost no absorption spectral change.
Accordingly, the chemosensor L solution showed a visible color
change from colorless to yellow with sulde, while with the other
anions it remained colorless (Fig. 4c). This can be explained by
the fact that sulde deprotonates chemosensor L to produce
a yellow color product, whereas the other anions cannot react
with chemosensor L and the solution remained colorless. On the
other hand, the addition of sulde to chemosensor L resulted in
emission spectra change accompanied by orange uorescence
and the uorescence intensity signicantly enhanced about 12.4fold compared with that of chemosensor L. However, no such
obvious change was observed in the presence of other anions
and thiols (Fig. 4b and d). These results demonstrate that chemosensor L was highly selective to sulde and also efficiently
discriminated sulde from thiol-containing analytes.
Moreover, chemosensor L exhibited the high selectivity to
sulde in the presence of other interferences in the mixture,
such as F, Cl, Br, I, N3, NO2, NO3, SO42, Cys, and GSH
(Fig. S1, ESI†). The sulde sensing had no considerable interference from coexistent anions. Although a slight decrease in
spectra was observed in competitive thiols such as Cys and GSH,
the absorbance and emission signals still presented the
signicant recognition of sulde in the solution. Therefore,
chemosensor L was able to effectively detect sulde in the
presence of foreign solutions. Sulde tends to compete for the
proton and deprotonates the hydroxyl (–OH) group from chemosensor L.24,26 Although uoride can deprotonate phenolic
hydroxyl in a quinolone-containing receptor,34 this has not
caused any interference to our proposed chemosensor L,
presumably because sulde is a stronger base than uoride.
These results indicate that chemosensor L is not only highly
selective for sulde, but also possesses an excellent capacity to
resist disturbance from other anions and thiols.
3.4
The sensing mechanism of chemosensor L
To better understand the interaction between chemosensor L
and sulde, FTIR spectra of chemosensor L were successively
Anal. Methods, 2021, 13, 1332–1339 | 1335
View Article Online
Published on 09 February 2021. Downloaded by Istanbul Technical University on 5/27/2023 12:46:04 AM.
Analytical Methods
(a) FTIR spectra of chemosensor L in the presence and absence
of sulfide. The optimized structures of the ground state for (b) chemosensor L and (c) chemosensor L + S2. HOMO – LUMO and energy
gaps of (d) chemosensor L and (e) chemosensor L + S2.
Fig. 5
recorded with and without sulde. As shown in Fig. 5a, the
absorbance band in the IR spectra of chemosensor L was
observed at 3425 cm1 as a result of O–H stretching. The broad
chemical band was possibly due to the overlap between two
aromatic –OH groups contained in chemosensor L. However,
upon reaction with sulde, as a result of chemical modication
at the –OH group, the O–H vibration band was converted to
a sharp band and concurrently shied to 3355 cm1. The
hydrogen atom in the –OH functional group was deprotonated
and the chemical shi changed due to the deprotonation of the
aromatic –OH by sulde. Besides, there was very little IR spectral change from the remaining structural features in chemosensor L when reacted with sulde. In addition, the change in
UV-Vis absorption of chemosensor L in the presence of strong
base OH anions, known as a deprotonation factor, was nearly
similar to that of chemosensor L obtained with sulde (Fig. S2,
ESI†). Moreover, a Job's plot was used to determine the binding
mode of the complex chemosensor L and sulde. The molar
fraction of sulde corresponding to the maximum absorbance
change amounted to 0.5, indicating 1 : 1 stoichiometry between
chemosensor L and sulde (Fig. S3, ESI†). These results
conrmed that the changes in the color and absorption were
due to the deprotonation of a hydroxyl functionality contained
in chemosensor L. The sensing mechanism of the chemosensor
L with sulde is proposed in Scheme 2. The presence of sulde
Anticipated sensing mechanism of chemosensor L for
sulfide detection.
Scheme 2
1336 | Anal. Methods, 2021, 13, 1332–1339
Paper
deprotonates the hydroxyl group in the benzylidene moiety of
chemosensor L to produce a yellow color as well as spectral
change (Fig. 3a).
Furthermore, the optimized structure (Fig. 5b and c) and
molecular orbital properties (HOMO and LUMO) (Fig. 5d and e)
of chemosensor L with and without sulde were simulated by
DFT calculations using the B3LYP/6-31G* basic level of theory.
From the DFT calculations, the p electrons were generally located
on the benzylidene moiety in the HOMO of chemosensor L both
with and without sulde. In the LUMO, however, the electrons
were transferred to rearrange around the HC]N linkage when
chemosensor L was excited with an energy gap of 0.146 eV, which
was much higher than the energy needed for deprotonating
chemosensor L (0.036 eV) by sulde (Fig. 5d and e). Therefore,
the tendency of releasing protons from the hydroxyl group was
considered as the LUMO state of chemosensor L when reacted
with sulde. The hydroxyl unit being adjacent to the Schiff's base
linkage tended to be deprotonated, which eventually enhanced
the intramolecular charge transfer (ICT) in the chemosensor L,
resulting in the spectral and color change.35 This demonstration
was in good agreement with the apparent change in both UV-Vis
and uorescence spectra conforming to the deprotonation
mechanism of chemosensor L toward sulde (Scheme 2).
3.5
L
pH and time effects and reversibility of the chemosensor
To investigate the effect of pH on sulde binding, the sensing
behavior of chemosensor L was demonstrated with a series of
solutions with pH values ranging from 2 to 12. Chemosensor L
efficiently monitored sulde in the solution with pH ranging
from 6 to 9 (Fig. 6a and S4, ESI†). In general, in acidic pH, the
deprotonation is less likely to occur due to the inuence of
a high concentration of H+. However, a high concentration of
OH above pH 10 caused interference to chemosensor L.
Moreover, H2S species are predominant at acidic pH, while an
appreciable number of dissociated anions such as HS and S2
exist at physiological pH.26,36 Therefore, the natural pH was
chosen as the optimum condition to evaluate the performance
of the chemosensor L. These results show that chemosensor L
can be effectively applied for the detection of sulde in real
samples under physiological conditions.
Since real-time sulde monitoring of sulde is important, we
also demonstrated the response time of chemosensor L. The
absorbance was measured aer adding sulde into chemosensor L solution and then recorded every minute. Fig. 6b
shows the stability of chemosensor-based colorimetric detection of sulde. A quick response was observed aer adding
sulde into the chemosensor L solution. In particular, a new
absorption band at 400 nm immediately reached the plateau
within 10 s and then remained stable for the next 30 min. The
proposed chemosensor L exhibited an impressively fast
response in comparison with reported chemosensors for the
detection of sulde (Table S1, ESI†). This result conrmed the
rapid and sustained detection ability of chemosensor L for
sulde, which is probably appropriate for real-time monitoring
of sulde.
This journal is © The Royal Society of Chemistry 2021
View Article Online
Published on 09 February 2021. Downloaded by Istanbul Technical University on 5/27/2023 12:46:04 AM.
Paper
Analytical Methods
Fig. 7 The chemosensor L-coated paper strips for sulfide detection.
(a) Visible color change and (b) mean intensity of chemosensor Lcoated paper strips (100 mM) in the presence of various anions (100
mM). (c) Visible color change and (d) mean intensity of chemosensor Lcoated paper strips in the presence of various concentrations of sulfide
(0–100 mM).
The absorbance of chemosensor L (50 mM) with sulfide at
different pH (a) and intervals time (b) in the DMSO: Bis–Tris buffer
(4 : 6, 10 mM) solution. (c) Reversibility of chemosensor L (50 mM) in
the DMSO: Bis–Tris buffer (4 : 6, 10 mM, pH 7.0) solution.
Fig. 6
Interestingly, the chemosensor L is also reversible for the
distinct recognition of sulde over cycles of the sequential
addition of sulde, followed by H+ ions (Fig. 6c). Chemosensor
L displayed a yellow color and the maximum absorption at 400
nm when reacted with sulde. Upon addition of HCl to the
yellow-colored chemosensor L – sulde complex, the absorption
band at 400 nm disappeared and, correspondingly, the color
changed from yellow to colorless. When sulde was added
again, the spectra and yellow color were recovered. The spectral
absorptions of chemosensor L can be reversed by the addition
of sulde and HCl in succession (Fig. S5, ESI†). Therefore, the
chemosensor L could be reused, and the sensing behavior could
be reversible at least 3 cycles for the distinct colorimetric
recognition of sulde.
3.6
Sulde detection using paper strips
Aer optimizing the experiment in aqueous solution, we
established the paper-based assay for the determination of
sulde. Fig. 7 shows the result of chemosensor L-coated paper
This journal is © The Royal Society of Chemistry 2021
strips used for colorimetric recognition of sulde. By using laser
cutting techniques, paper strips were fabricated and were
immersed in the chemosensor L solution to be ready-to-use
colorimetric paper strips to visualize sulde (Fig. 1). Consequently, the paper strips displayed an obvious color change to
yellow when sulde was present inside a solution containing
various anions (Fig. 7a). Paper strips were further analyzed by
using ImageJ to obtain the mean intensities. The result indicated that the intensity of the paper strip containing sulde was
signicantly higher than those of the paper strip containing
other anions (Fig. 7b). Additionally, the color intensity of the
paper strips increased as the sulde concentration increased
from 0 to 100 mM. The detection limit of paper-based chemosensor L for naked-eye visualization of sulde was as low as 30
mM (Fig. 7c and d). Therefore, the paper strips could be effectively used for simple, selective, and rapid colorimetric detection of sulde for environmental monitoring of water samples
in real time in remote areas with scarce analytical
instrumentation.
In addition, to further test the potential of the synthesized
chemosensor L for the detection of sulde in real samples,
chemosensor L was applied to quantitate sulde in water
contaminated with sulde. Tap water samples were collected
and spiked with sulde to mimic sulde-contaminated water.
Sample analysis was performed by applying the same method
used for linearity experiments, with the optimized conditions.
The sulde-contaminated water samples were added to chemosensor L prepared in Bis–Tris buffer to perform the quantitative measurement. Each sample was analyzed with four
replicates to evaluate the reliability and accuracy of the
proposed method. The results are illustrated in Table 1, and it
was found that sulde was recovered, from 94 to 99%, for tap
water samples. These values indicated that the introduced
colorimetric approach exhibited reliable and precise sulde
recovery as compared to reported sensing strategies that could
be potentially applicable for the direct determination of sulde
in environmental samples.37–40
Anal. Methods, 2021, 13, 1332–1339 | 1337
View Article Online
Analytical Methods
Published on 09 February 2021. Downloaded by Istanbul Technical University on 5/27/2023 12:46:04 AM.
Table 1
Paper
Determination of sulfide concentration in water samples
a
Sample
Spiked amount
(mM)
Detected amount
(mM)
Percentage
recovery
Tap water 1
Tap water 2
Tap water 3
200
500
800
188.0 9.2
478.3 27.6
791.7 73.7
94%
96%
99%
a
Mean standard deviation.
4 Conclusions
In this study, we have successfully developed a novel paperbased chemosensor for rapid colorimetric detection of sulde
with high selectivity based on the deprotonation mechanism.
Chemosensor L selectively detected sulde without any interference from other anions and thiols, through a color change
from colorless to yellow. The paper-based colorimetric chemosensor L has an impressively fast response time of 10 s
compared with previous studies (Table S1, ESI†) and a detection
limit of 25.8 mM and 30 mM in aqueous solution and paper
strips, respectively. A portable paper strip is potentially applied
for on-site recognition of sulde in hazardous environments.
Furthermore, chemosensor L was found to enable reliable
quantitative determination of sulde for potential use in
monitoring of sulde in environmental samples such as water.
Therefore, the method introduced in this study is highly
promising for on-site and real-time colorimetric monitoring of
sulde in environmental and biological samples.
Author contributions
Duc Anh Thai: methodology, formal analysis, investigation,
resources, and writing – original dra. Nae Yoon Lee: conceptualization, methodology, resources, data curation, project
administration, funding acquisition, investigation, writing –
review & editing, and supervision.
Conflicts of interest
There are no conicts to declare.
Acknowledgements
This work was supported by the National Research Foundation
of Korea (NRF) grant funded by the Korea government (MSIT)
(No. NRF-2020R1A2B5B01001971). This research was also supported by Korea Basic Institute (National Research facilities and
Equipment Center) grant funded by the Ministry of Education
(2020R1A6C103A050) and by the Gachon University research
fund of 2019 (GCU-2019-0817).
Notes and references
1 J. W. Morse, F. J. Millero, J. C. Cornwell and D. Rickard,
Earth-Sci. Rev., 1987, 24, 1–42.
1338 | Anal. Methods, 2021, 13, 1332–1339
2 L. Zhang, P. De Schryver, B. De Gusseme, W. De Muynck,
N. Boon and W. Verstraete, Water Res., 2008, 42, 1–12.
3 L. Tang, M. Cai, Z. Huang, K. Zhong, S. Hou, Y. Bian and
R. Nandhakumar, Sens. Actuators, B, 2013, 185, 188–194.
4 O. Kabil and R. Banerjee, J. Biol. Chem., 2010, 285, 21903–
21907.
5 K. Tanizawa, Journal of Biochemistry, 2011, 149, 357–359.
6 D. Boehning and S. H. Snyder, Annu. Rev. Neurosci., 2003, 26,
105–131.
7 G. Yang, L. Wu and R. Wang, FASEB J., 2006, 20, 553–555.
8 W. Yang, G. Yang, X. Jia, L. Wu and R. Wang, J. Physiol., 2005,
569, 519–531.
9 Y. J. Peng, J. Nanduri, G. Raghuraman, D. Souvannakitti,
M. M. Gadalla, G. K. Kumar, S. H. Snyder and
N. R. Prabhakar, Proc. Natl. Acad. Sci. U. S. A., 2010, 107,
10719–10724.
10 K. Eto, T. Asada, K. Arima, T. Makifuchi and H. Kimura,
Biochem. Biophys. Res. Commun., 2002, 293, 1485–1488.
11 S. Fiorucci, E. Antonelli, A. Mencarelli, S. Orlandi, B. Renga,
G. Rizzo, E. Distrutti, V. Shah and A. Morelli, Hepatology,
2005, 42, 539–548.
12 D. J. Elsey, R. C. Fowkes and G. F. Baxter, Cell Biochem.
Funct., 2010, 28, 95–106.
13 J. Radford-Knoery and G. A. Cutter, Anal. Chem., 1993, 65,
976–982.
14 S. K. Pandey and K. H. Kim, Environ. Sci. Technol., 2009, 43,
3020–3029.
15 M. Okumura, N. Yano, K. Fujinaga, Y. Seike and S. Matsuo,
Anal. Sci., 1999, 15, 427–431.
16 N. S. Lawrence, J. Davis and R. G. Compton, Talanta, 2000,
52, 771–784.
17 A. R. Lippert, E. J. New and C. J. Chang, J. Am. Chem. Soc.,
2011, 133, 10078–10080.
18 V. S. Lin, A. R. Lippert and C. J. Chang, Proc. Natl. Acad. Sci.
U. S. A., 2013, 110, 7131–7135.
19 Y. Qian, J. Karpus, O. Kabil, S. Y. Zhang, H. L. Zhu,
R. Banerjee, J. Zhao and C. He, Nat. Commun., 2011, 2, 1–7.
20 X. Li, S. Zhang, J. Cao, N. Xie, T. Liu, B. Yang, Q. He and
Y. Hu, Chem. Commun., 2013, 49, 8656–8658.
21 M. G. Choi, S. Cha, H. Lee, H. L. Jeon and S. K. Chang, Chem.
Commun., 2009, 47, 7390–7392.
22 Z. Guo, G. Chen, G. Zeng, Z. Li, A. Chen, J. Wang and L. Jiang,
Analyst, 2015, 140, 1772–1786.
23 F. Yu, X. Han and L. Chen, Chem. Commun., 2014, 50, 12234–
12249.
24 S. Y. Lee and C. Kim, RSC Adv., 2016, 6, 85091–85099.
25 A. K. Manna, J. Mondal, R. Chandra, K. Rout and G. K. Patra,
Anal. Methods, 2018, 10, 2317–2326.
26 R. Kaushik, A. Ghosh, A. Singh and D. A. Jose, Anal. Chim.
Acta, 2018, 1040, 177–186.
27 K. Y. Ryu, J. J. Lee, J. A. Kim, D. Y. Park and C. Kim, RSC Adv.,
2016, 6, 16586–16597.
28 W. Chen, X. Fang, H. Li, H. Cao and J. Kong, Sci. Rep., 2016,
6, 1–7.
29 M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria,
M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone,
This journal is © The Royal Society of Chemistry 2021
View Article Online
Paper
30
31
Published on 09 February 2021. Downloaded by Istanbul Technical University on 5/27/2023 12:46:04 AM.
32
33
34
35
B. Mennucci, G. A. Petersson and H. Nakatsuji, Gaussian 09,
Gaussian Inc. Wallingford, CT, 2009.
A. D. Becke, J. Chem. Phys., 1993, 98, 1372–1377.
J. W. Y. Liew, K. S. Loh, A. Ahmad, K. L. Lim and W. R. Wan
Daud, Energies, 2018, 11, 1910.
D. A. Jose, P. Kar, D. Koley, B. Ganguly, W. Thiel, H. N. Ghosh
and A. Das, Inorg. Chem., 2007, 46, 5576–5584.
Z. Li, S. Wang, L. Xiao, X. Li, X. Shao, X. Jing and L. Ren,
Inorg. Chim. Acta, 2018, 476, 7–11.
US Environmental Protection Agency, Fed. Regist., 1986, 51,
34014–34025.
Y. C. Cai, C. Li and Q. H. Song, ACS Sens., 2017, 2, 834–841.
This journal is © The Royal Society of Chemistry 2021
Analytical Methods
36 B. Tan, S. Jin, J. Sun, Z. Gu, X. Sun, Y. Zhu and Y. Z. Zhu, Sci.
Rep., 2017, 7, 1–12.
37 Y. Wang, H. Sun, L. Hou, Z. Shang, Z. Dong and W. Jin, Anal.
Methods, 2013, 5, 5493–5500.
38 A. Hatamie, B. Zargar and A. Jalali, Talanta, 2014, 121, 234–
238.
39 U. R. Kondekar, L. S. Walekar, A. H. Gore, P. V. Anbhule,
S. H. Han, S. R. Patil and G. B. Kolekar, Anal. Methods,
2015, 7, 2547–2553.
40 S. Yan, R. Song and Y. Tang, Anal. Methods, 2016, 8, 3768–
3773.
Anal. Methods, 2021, 13, 1332–1339 | 1339
Download