You are on page 1of 9

Sensors and Actuators B 255 (2018) 3170–3178

Contents lists available at ScienceDirect

Sensors and Actuators B: Chemical


journal homepage: www.elsevier.com/locate/snb

Research paper

A novel turn-on fluorescent probe for cyanide detection in aqueous


media based on a BODIPY-hemicyanine conjugate
Yanhua Yu a,∗ , Tingting Shu a , Bingjie Yu a , Yun Deng a , Cheng Fu a , Yangguan Gao a ,
Changzhi Dong a,c , Yibin Ruan b,∗
a
Institute for Interdisciplinary Research, Jianghan University, Wuhan 430056, China
b
Technology Center of China Tobacco Guizhou Industrial Co. Ltd., Guiyang 550003, China
c
University Paris Diderot, Sorbonne Paris Cité, ITODYS, UMR CNRS 7086, 15 rue J-A de Baïf, 75205 Paris Cedex 13, France

a r t i c l e i n f o a b s t r a c t

Article history: A novel BODIPY-hemicyanine based fluorescent probe 1 for cyanide (CN− ) detection was designed and
Received 3 July 2017 synthesized. Among the tested anions, the probe showed high selectivity towards CN− . Only addition
Received in revised form of CN− to probe 1 in aqueous solution (VH2O /VEtOH = 1/1) could result in a remarkable blue shift of the
19 September 2017
absorption band from 385 nm to 265 nm and induce 20-fold fluorescence enhancement at 515 nm. The
Accepted 20 September 2017
sensing mechanism was based on nucleophilic addition between CN− and indolium group, which was
Available online 21 September 2017
confirmed by 1 H NMR and mass spectrum analysis. Plot of fluorescence intensity as a function of CN−
concentrations exhibited a good linear relationship in the range of 0–10 ␮M, with a detection limit to
Keywords:
Cyanide
be 1.53 ppb. Moreover, the proposed sensing approach could work in a wide pH range from 5.0–9.0 and
BODIPY-hemicyanine conjugate reach an equilibrium instantaneously. Finally, probe 1 was successfully applied in detection of CN− in
Fluorescence enhancement natural water samples and on test paper strips with satisfactory results.
Michael addition © 2017 Elsevier B.V. All rights reserved.

1. Introduction injection [9], colorimetric and fluorescent sensors [10–31]. Among


these methods, fluorescent sensors have drawn much attention due
Cyanide compounds are one of the most toxic species to human to their various advantages over other techniques, such as sim-
health and environment [1], due to their powerful interaction with ple operation, versatile adaptation, low-cost, portability, excellent
ferric iron atoms in metalloenzymes, which can disorder some sensitivity and rapid response. The reported CN− fluorescent sen-
important biological processes like cellular respiration and oxida- sors were mainly based on metal complex ensemble displacement
tive metabolism [2]. Still they are widely utilized in chemical and [32–39], nucleophilic addition [40], nanoparticles and polymeric
industrial processes, such as plastics production, gold mining, met- frameworks [41,42]. Among them, nucleophilic-reactions based on
allurgy, synthetic fibers and the resins industry [3]. Any accidental dicyano-vinyl [43–47], pyridinium ring [48], oxazines [49], acri-
release of cyanide into the environment could cause serious health dinium [50], aldehyde [51,52], boronic acid [53], indolium groups
problems. It was reported that in 2015 a large amount of cyanide [54–61], generally exhibit satisfactory selectivity due to the excep-
was leaked out in Tianjin Port by accident, which was considered tional nucleophilicity of CN− . Indolium has recently emerged as
to be a serious environmental disaster. Moreover, according to the one of most promising reactive group for CN− because of its pos-
World Health Organization (WHO), the maximum permissive level itive charge character which results in a strong attractive force
of cyanide in the drinking water is 1.9 ␮M [4]. Therefore, it is great between indolium group and CN− . The detailed computational cal-
of importance to develop novel approaches for cyanide anion detec- culations further demonstrated that solvation of anions by H2 O
tion. finally give the unique selectivity for CN− [62]. However, many of
The strategies for detecting cyanide include mass spectrometry them still existed several drawbacks in aqueous environment such
[5], electrochemistry [6], ion chromatography [7], headspace sorp- as low sensitivity, narrow pH working range, which cannot fulfill
tive solid phase microextraction with a spectrophotometry [8], flow the requirement for the real sample analysis [63–65].
Borondipyrromethene (BODIPY) derivatives have been widely
used in CN− detection due to its distinct advantages such as large
molar excitation coefficient, high fluorescence quantum yield, nar-
∗ Corresponding authors. row emission bandwidth, high photostability, small stokes shifts
E-mail addresses: hpyyh@aliyun.com (Y. Yu), ybruan@xmu.edu.cn (Y. Ruan).

https://doi.org/10.1016/j.snb.2017.09.142
0925-4005/© 2017 Elsevier B.V. All rights reserved.
Y. Yu et al. / Sensors and Actuators B 255 (2018) 3170–3178 3171

Scheme 1. Design and synthesis of BODIPY-hemicyanine dye 1.

and tunable fluorescence characteristic [66–74]. In this work, we surement. 1 H NMR, 13 C NMR spectra were collected on a Bruker
herein report a novel fluorescent BODIPY–hemicyanine probe 1 (as Advance 400 MHz spectrometer in CDCl3 , tetramethylsilane (TMS)
shown in Scheme 1) as a highly selective and sensitive turn-on fluo- as internal standard, chemical shifts are given in ppm related to
rescent probe for CN− detection in aqueous solution. Hemicyanine the protonated solvent as internal reference (1 H: CHCl3 in CDCl3 ,
unit containing indolium group which was incorporated to BODIPY 7.26 ppm, 13 CDCl3 in CDCl3 , 77.16 ppm), coupling constants (J) are
not only acted as the binding unit for CN− , but also promoted the given in Hz, the following abbreviations were used to explain
intramolecular charge transfer (ICT) process and quenched fluores- the multiplicities: s = singlet; d = doublet; m = multiplet. MS spectra
cence of BODIPY. When hemicyanine was attacked by CN− through were recorded on a Bruker amaZon SL instrument using stan-
nucleophilic addition, ICT process was blocked which then led to dard conditions (ESI). All the measurement experiments were
turn-on fluorescence. As expected, this BODIPY-hemicyanine based performed at room temperature.
probe showed excellent selectivity to CN− over other tested anions.
Furthermore, It’s worthy of notice that the BODIPY group was not
a good electron donor, which could lead to a more positive charge 2.3. Synthesis and characterization
distribution on indolium group and give a fast kinetics and low
detection limit for CN− detection. In our protocol the response time 2.3.1. Synthesis of 1,3,5,6-tetramethyl-
was within 3 s in aqueous media and the detection limit was cal- 8-(4-(hydroxymethyl)phenyl)BODIPY 2
culated to be 59 nM, far lower than the permissive level of WHO 1,3,5,6-tetramethyl- 8-(4-(hydroxymethyl)phenyl)BODIPY 2
in drinking water. Finally, this probe was successfully applied to was synthesized according to the literature reported pro-
detect CN− in real water samples and on the test paper strips. cedures [52]. To a solution of 2,4-dimethylpyrrole (2.2 mL,
20.0 mmol) in deoxygenated CH2 Cl2 (150 mL) was added 4-
hydroxymethylbenzaldehyde (1.4 g, 10.0 mmol) and one drop of
2. Materials and methods
TFA at room temperature. The mixture was stirred over 12 h under
argon, then treated with DDQ (2.3 g, 10.0 mmol), the mixture was
2.1. Materials
continued stirring for 1 h followed by addition of Et3 N (30 mL).
After stirring for 15 min, the solution was cooled to 0 ◦ C, then
4-Hydroxymethylbenzaldehyde, 2,4-dimethylpyrrole, 2,3-
BF3 ·OEt2 (30 mL) was added to the solution, the mixture was
dicyano-5,6-dichlorobenzoquinone (DDQ), trifluoroacetic acid
stirred at room temperature for further 3 h. After the completion
(TFA), boron trifluoride ether complex (BF3 ·OEt2 ) were purchased
of reaction, washed with saturated NaHCO3 solution, the organic
from Sigma-Aldrich without further purification. Triethylamine,
phase was separated, dried over MgSO4, filtered, and concentrated.
manganese dioxide, piperidine and solvents were purchased
The residue was purified by silica gel column chromatography
from Aladdin (Shanghai, China). In anhydrous reaction, ethanol
(CH2 Cl2 /petroleum ether: 1/2) to give compound 2 (0.9 g, 2.5 mmol)
was dried by distillation over sodium according to the stan-
as a red solid in 50% yield. 1 H NMR (400 MHz, CDCl3) ı 7.49 (d,
dard procedures. Solvents used for extraction treatment and
J = 7.8 Hz, 2H), 7.27 (d, J = 7.8 Hz, 2H), 5.98 (s, 2H), 4.80 (s, 2H), 2.55
column chromatography such as petroleum ether, ethyl acetate,
(s, 6H), 1.38 (s, 6H). 13 C NMR (101 MHz, CDCl3 )ı155.49, 143.10,
dichloromethane and methanol were purchased from Sinopharm
141.89, 141.56, 134.20, 131.48, 128.18, 127.39, 121.22, 64.66, 14.48.
Chemical Reagent Co., Ltd (Shanghai, China) in commercial grade
HRMS: calculated for [M + H]+ : 355.1793, measured: 355.1712.
and distilled before use. All column chromatography was carried
out using silica gel (200–300 mesh). Thin layer chromatography
(TLC) was performed on silica gel coated on aluminum plates. 2.3.2. Synthesis of 1,3,5,6-tetramethyl-
Milli-Q water was used in all aqueous analytical experiments. All 8-(4-(formylphenyl))BODIPY 3
anions were prepared from their tetrabutylammonium (TBA+ ) To a solution of compound 2 (0.9 g, 2.5 mmol) in CH2 Cl2 (150 mL)
salts. was added MnO2 (5.4 g, 62.5 mmol), the mixture was refluxed
over 12 h. After the completion of reaction, the mixture was fil-
2.2. Apparatus tered through celite, the filtrate was evaporated, then the solid
residue was further purified by silica gel column chromatography
Ultraviolet-visible (UV-vis) spectra were measured on a Perkin (CH2 Cl2 /petroleum ether: 1/2) to give compound 3 (0.7 g, 2.0 mmol)
Elmer Lambda 25 spectrometer. Fluorescence spectra measure- as a red solid in 80% yield. 1 H NMR (400 MHz, CDCl3 ) ı 10.14 (s,
ments were performed on a Perkin Elmer LS 55 spectrometer. 1H), 8.03 (d, J = 7.8 Hz, 2H), 7.51 (d, J = 7.8 Hz, 2H), 6.00 (s, 2H), 2.56
Perkin Elmer quartz cells with an inner path length of 10 mm (s, 6H), 1.36 (s, 6H). 13 C NMR (101 MHz, CDCl3 ) ı 191.46, 156.26,
were used for absorption and fluorescence spectroscopy mea- 142.76, 141.41, 139.69, 136.68, 130.83, 130.34, 129.16, 121.64,
3172 Y. Yu et al. / Sensors and Actuators B 255 (2018) 3170–3178

14.63, 14.52. HRMS: calculated for [M + Na]+ : 375.1456, measured:


375.1475.

2.3.3. Synthesis of BODIPY-hemicyanine conjugate 1


To a solution of compound 3 (0.5 g 1.4 mmol) in anhydrous
ethanol was added compound 4 (0.4 g, 1.4 mmol) and piperidine (3
drops). The mixture was refluxed over 12 h, After cooling the solid
was collected, washed with anhydrous ethanol, then dried, giving
a solid. The solid was purified by silica gel column chromatogra-
phy (CH2 Cl2 /MeOH: 100/1) to give compound 1 (0.5 g, 0.78 mmol)
as a dark purple solid in 56% yield. 1 H NMR (400 MHz, CDCl3 ) ı
8.36-8.23 (m, 3H), 7.96 (d, J = 16.3 Hz, 1H), 7.62-7.59 (m, 4H), 7.49-
7.47 (m, 2H), 5.99 (s, 2H), 4.54 (s, 3H), 2.56 (s, 6H), 1.91 (s, 6H),
1.42 (s, 6H).13 C NMR (101 MHz, CDCl3 ) ı 182.66, 174.25, 156.14,
152.78, 143.08, 142.89, 141.53, 140.51, 139.82, 134.37, 131.70,
130.84, 130.43, 129.86, 129.48, 128.94, 125.39, 122.65, 122.01,
121.57, 115.35, 114.66, 110.90, 105.02, 52.92, 48.86, 37.66, 32.77,
28.11, 26.70, 14.88, 14.66. HRMS: calculated for [M]+ : 508.2730,
measured: 508.2751.
Fig. 1. Absorption spectra of probe 1 (10 ␮M) in aqueous solution (VH2O /VEtOH = 1/1)
with increasing CN− concentrations; inset: the absorption intensity at A385nm versus
2.4. Photophysical property study
concentration of CN− from 0 to 60 ␮M.

The stock solution of probe 1 (1 mM) was prepared in ethanol.


The UV–vis absorption spectra of the solutions were measured in 3. Results and discussions
a 1 cm quartz cell from 250 nm to 600 nm at room temperature.
The fluorescence spectra were measured from 475 nm to 625 nm 3.1. Design and synthesis of probe 1
at room temperature using an excitation wavelength at 385 nm.
The spectra were acquired from 10 ␮M solutions, diluted from the Synthesis of probe 1 was illustrated in Scheme 1.
stock solution with their respective solvents. BODIPY 2 was synthesized through a condensation of 4-
hydroxymethylbenzaldehyde and 2,4-dimethylpyrrole in the
presence of trifluoroacetic acid (TFA) as catalyst in CH2 Cl2 solution,
2.5. UV–vis absorption and fluorescence spectra study
followed by oxidation with 2,3-dicyano-5,6-dichlorobenzoquinone
(DDQ). The boron difluoride bridge was formed by treatment with
All the UV–vis absorption and fluorescence spectra were
boron trifluoride diethyl etherate (BF3 ·Et2 O) and triethylamine.
acquired from ethanol and water (V/V = 1/1). The stock solutions
Hydroxyl group was oxidized by MnO2 to afford BODIPY 3 in 80%
of the tested anions (F− , Cl− , Br− , I− , ClO4 − , AcO− , H2 PO4 − , NO3 − ,
yields. BODIPY-hemicyanine conjugate 1 was prepared in 56%
HSO4 − , CN− ) (5 mM) were prepared by dissolving their tetrabuty-
yields from a Knoevenagel condensation between BODIPY 3 with
lammonium (TBA+ ) salts in Milli-Q water. For selectivity study, the
1-methyl-2,3,3-trimethyl-3H-indolium, which was synthesized
volumes of the mixtures were adjusted to 2.5 mL by ethanol and
according to the reported procedures [53]. Compounds 2, 3 and 1
water (V/V = 1/1) to afford the final concentration of 10 mM for the
were well characterized by 1 H NMR, 13 C NMR and HRMS.
fluorophores and 80 mM for the anions. The mixture was shaken for
3 s before recording absorption and fluorescence spectra. For CN−
3.2. Optical properties of probe 1
titration study, to a solution of probe 1 (10 ␮M) was added CN− (0.1
equiv.) and the mixture was shaken for 3 s before recording absorp-
With probe 1 in hand, we firstly examined its sensing ability
tion and fluorescence spectra. This step was repeated until reaching
towards CN− in aqueous solution (VH2O /VEtOH = 1/1) by UV–vis and
the desired amount of CN− . The fluorescence quantum yield (ФF )
fluorescence spectroscopy. As shown in Fig. 1, probe 1 exhibits
was calculated by the standard method using Rhodamine 6G in
three main absorption bands peaked at 385 and 502, 545 nm, which
ethanol as a reference.
are respectively attributed to intramolecular charge transfer from
BODIPY to hemicyanine and BODIPY’s typical absorption. With the
2.6. Job’s plot measurements incremental additions of CN− , the absorption band of probe 1 at
385 nm decreases progressively with appearance of a new band at
A series of solutions containing probe 1 and CN− in the mixture 256 nm; and the absorption bands at 502 nm and 545 nm decreased
of ethanol and water (V/V = 1/1) were prepared, which the sum of slightly. An isosbestic point at 295 nm was observed, which indi-
the concentrations remained constant (10 ␮M). The molar fraction cated formation of a new species. Plot of change in the absorption
of probe 1 was varied from 0 to 1. After measuring fluorescent band at 385 nm of probe 1 as a function of the added CN− con-
intensity of each solution,F(F = F(probe 1) − F(probe 1 + CN− )) centrations ranging from 0 to 11 mM gave a good linear regression
was plotted against the molar fractions of probe 1. (R2 = 0.991) and reached a plateau in the presence of 2.0 equiv. of
CN− (Fig. 2).
2.7. Paper-based sensor Fluorescence response of probe 1 towards CN− was also inves-
tigated. As shown in Fig. 3, probe 1 exhibits an emission band
The solution of probe 1 (0.1 mM, 20 ␮L) in aqueous solution centered at 515 nm upon excitation at 385 nm, which could be
(VH2O /VEtOH = 1/1) was pipetted onto the filter paper and then dried attributed to emission of BODIPY moiety. With increasing addi-
by air. Various anions and CN− at different concentrations (20 ␮L) tion of CN− , fluorescence intensity at 515 nm increased gradually
were pipetted onto the surface of the spots of probe 1. After air and reached a plateau in the presence of 20 ␮M CN− . Meanwhile,
drying, the images of the filter paper under 365 nm UV lamp were the fluorescence quantum yield (ФF ) increased from 0.015 to 0.33
photographically recorded. (SI). The initial weak fluorescence emission of probe 1 was proba-
Y. Yu et al. / Sensors and Actuators B 255 (2018) 3170–3178 3173

Table 1
Comparison of CN− detection limit of probe 1 with other reported sensors.

Sensors Solvent Detection limit Ref.

Coumarin-based MeCN/buffer (V/V = 1/1) 9.37 ␮M [58]


Naphthaldehyde-based EtOH/H2 O (V/V = 95/5) 1.6 ␮M [75]
Benzofurazan-based CH3 CN/H2 O (V/V = 95/5) 1.5 ␮M [76]
Indanedione-based THF/H2 O (V/V = 1/9) 0.94 ␮M [16]
BODIPY-salicylaldehyde-based DMSO/buffer (V/V = 9/1) 0.88 ␮M [51]
Diketopyrrolopyrrole-based THF 0.36 ␮M [77]
Gold nanodots-based Water 0.15 ␮M [18]
Dicyanovinyl-based DMSO/H2 O (V/V = 2/8) 0.14 ␮M [78]
Carbazole-based CH3 CN/H2 O (V/V = 95/5) 0.13 ␮M [79]
BODIPY-hemicyanine conjugate EtOH/H2 O (V/V = 1/1) 0.059 ␮M present work

Fig. 2. Plot of change in absorption band of probe 1 at 358 nm as a function of Fig. 4. Linear relationship between fluorescence intensity ratio (I/I0 ) of probe
concentrations of CN− from 0 to 11 ␮M in aqueous solution (VH2O /VEtOH = 1/1). 1 and the low concentrations of CN− from 0 to 10 ␮M in aqueous solution
(VH2O /VEtOH = 1/1).

(1.53 ppb) according to the equation 3s/k, where s is the standard


deviation of the blank measurements and k is the slope of the inten-
sity ratio versus the sample concentration plot. The detection limit
in present work is much lower than other reported CN− -selective
sensors (Table 1) and three times lower than maximum permissi-
ble level of CN− in drinking water proposed by the World Health
Organization (WHO).

3.3. Selectivity of probe 1 for CN− detection

Selectivity of probe 1 to CN− was evaluated by screening a series


of anions such as F− , Cl− , Br− , I− , ClO4 − , AcO− , H2 PO4 − , NO3 − ,
HSO4 − , CN− in aqueous solution (VH2O /VEtOH = 1/1). As shown in
Figs. 5 and 6, only addition of CN− can lead to a considerable absorp-
tion spectral change and turn-on fluorescence emission, while
other tested anions show no influence. This demonstrated that
probe 1 showed excellent selectivity for CN− over the other tested
anions. Emission color changes from blank to green were easily
Fig. 3. Fluorescence spectra of probe 1 (10 ␮M) in aqueous solution observed by the naked-eye under the illumination with a 365 nm
(VH2O /VEtOH = 1/1) as a function of CN− concentration; inset: fluorescence UV-lamp (Fig. 7). While color changes from orange to pink under
intensity ratio (I/I0 ) versus concentration of CN− from 0 to 60 ␮M. daylight could be only be observed with increasing the solution
concentration to 0.1 mM (Fig. S1)
bly due to the ICT from BODIPY unit to the strong electron acceptor
indolium group through its ␲-conjugated bridge. After nucleophilic 3.4. Competition studies of probe 1 for CN− detection
attack of CN− to indolium group, the electron accepting ability
of indolium group was strongly reduced and the ICT process was To further confirm practical applicability of probe 1 for CN−
blocked, which then led to turn-on fluorescent signal of BODIPY. detection, a competition experiment was carried out. As shown in
Plot of fluorescence intensity (I/I0 ) as a function of the concentra- Fig. 8, introducing the interfering anions including F− , Cl− , Br− , I− ,
tions of CN− also exhibited a good linear relationship (R2 = 0.993) ClO4 − , AcO− , H2 PO4 − , NO3 − , HSO4 − , CN− to the solution of probe
in the range of 0–10 ␮M (Fig. 4). The limit of detection for CN− 1 did not cause any fluorescence emission changes. A significant
based on fluorescence spectral changes was calculated to be 59 nM fluorescence enhancement was observed after the addition of CN−
3174 Y. Yu et al. / Sensors and Actuators B 255 (2018) 3170–3178

Fig. 5. Absorption spectra of probe 1 (10 ␮M) in aqueous solution (VH2O /VEtOH = 1/1)
in the presence of 8 equiv. different anions.
Fig. 8. Fluorescence intensity ratio (I/I0 ) of probe 1 (10 ␮M) in aqueous solution
(VH2O /VEtOH = 1/1) in the presence of different anions (80 ␮M) without/with CN−
(40 ␮M).

Fig. 6. Fluorescence spectra of probe 1 (10 ␮M) in aqueous solution


(VH2O /VEtOH = 1/1) in the presence of 8 equiv. different anions. (lex = 385 nm,
slit: 5 nm/5 nm).
Fig. 9. Fluorescence intensity at 515 nm of probe 1 (10 ␮M) in the absence and
presence of CN− (80 ␮M) as a function of pHs.
in all the interfering samples. This phenomenon demonstrated that
probe 1 had excellent selectivity for CN− detection over the other
anions even in the complicated environment. hydroxide buffer for pH 9.0–10.6). As depicted in Fig. 9, no
significant fluorescence changes could be observed at pH values
3.5. pH influence lower than 9 in the absence of CN− , while a 4-fold fluorescence
enhancement was detected when pH was higher than 10. This
It’s well known that the existing forms of CN− are highly pH- might be attributed to the nucleophilic attack of OH− to probe 1. In
dependent, which would then strongly affect the nucleophilic presence of CN−, almost no spectral change could be observed at pH
addition reaction. Therefore, pH influence on spectral response of below 4, which indicated that no reaction occurred between probe
probe 1 to CN− was investigated at different pH values (10 mM 1 and HCN. While in the pH range from 5 to 9, the fluorescence
NaH2 PO4 -citric acid buffer for pH 2.2–8.0 and 10 mM glycine- intensity at 515 nm of probe 1 increased remarkably in presence

Fig. 7. A visual fluorescence changes of probe 1 (10 ␮M) in aqueous solution (VH2O /VEtOH = 1/1) in the presence of 8 equiv. different anions under illumination with a 365 nm
UV lamp.
Y. Yu et al. / Sensors and Actuators B 255 (2018) 3170–3178 3175

Scheme 2. The proposed mechanism of probe 1 for CN− detection.

Table 2
Recovery data for CN− detection in spiked water samples.

Samples Spiked (␮M) Found (␮M) Recovery (%) RSD (%)

Lake water 5.0 5.23 ± 0.11 104.05 ± 2.03 1.81


8.0 7.87 ± 0.12 98.39 ± 1.51 1.35

Minerals water 5.0 5.13 ± 0.18 102.56 ± 3.73 3.64


8.0 8.17 ± 0.20 102.16 ± 2.28 2.79

Drinking water 5.0 5.27 ± 0.19 105.46 ± 3.67 1.67


8.0 7.83 ± 0.14 97.93 ± 1.81 1.68

shift of the absorption spectrum and fluorescence enhancement.


To confirm this plausible mechanism, 1 H NMR titration with CN−
was carried out in CDCl3 . As illustrated in Scheme 2 and Fig. 11,
upon addition of TBACN to probe 1, the vinyl protons (Ha and
Hb ) displayed upfield shifts from 8.25 ppm, 8.00 ppm to 6.62 ppm,
Fig. 10. Time-dependent fluorescence intensity at 515 nm of probe 1 (10 ␮M) in 6.33 ppm and the N-methyl protons (He ) displayed an upfield shift
aqueous solution (VH2O /VEtOH = 1/1) in the presence CN− (20 ␮M). from 4.55 ppm to 3.36 ppm. Moreover, the singlet peak correspond-
ing to C-methyl protons (Hc and Hd ) in probe 1 shifted from
of CN− . The results indicated that the probe 1 was stable in the pH 1.90 ppm (s, 6H) to 1.58 ppm (s, 3H) and 1.33 ppm (s, 3H), which
range from 2 to 9 and could successfully detect CN− in a range of indicated C-methyl was adjacent to a chiral center. All these infor-
pH values from 5 to 9.This is more applicable in real sample when mation supported the above assumption that nucleophilic attack
compared with other reported results. of CN− to hemicyanine moiety of probe 1 led to formation of 1–CN
adduct, in which the positive charged indolium group changed
into a neutral indole group with poor electron-withdrawing prop-
3.6. Effect of reaction time
erty. Job’s plot analysis also revealed that the probe 1 formed a 1:
1 adduct with CN− (Fig. S2). Furthermore, 1–CN adduct was fur-
Using chemodosimeters as a sensing approach usually takes
ther confirmed by mass spectrometry analysis, where a peak at
minutes or hours to complete the reactions, which restricts their
m/z 573.73 was observed, corresponding to [1-CN-K]+ (Fig. S3). All
applications in the real-time detection. As shown in Fig. 10, we use
these results strongly support the proposed mechanism that CN−
fluorescence spectroscopy to monitor the reaction between probe 1
attacks the hemicyanine moiety of probe 1.
and CN− in aqueous solution (VH2O /VEtOH = 1/1). Quite surprisingly,
this reaction was very fast and reached a plateau instantaneously,
which might be attributed to the poor electron-donating ability of
BODIPY unit in probe 1. Due to this fast kinetics, it showed promis- 3.8. CN− detection in spiked water samples
ing applications in real-time detection for CN− .
To investigate the practical application of probe 1 for CN− detec-
3.7. Mechanism study tion in complicated environment, probe 1 was applied to determine
CN− in the real water samples, such as lake water, mineral water
Based on the data collected from the absorption and fluores- and drinking water. The water samples were spiked with set
cence spectra, it’s deduced that nucleophilic addition between CN− amount of CN− (5 and 8 ␮M). As shown in Table 2, CN− was tested
and hemicyanine moiety in probe 1 was responsible for the blue with a satisfactory analytical accuracy and precision (RSD < 4%).
3176 Y. Yu et al. / Sensors and Actuators B 255 (2018) 3170–3178

Fig. 11. 1
H NMR spectra of probe 1 in CDCl3 in the absence and presence of CN− (1.0 equiv.).

Fig. 12. The photos of the filter paper containing probe 1 after exposure to different anions under UV light (365 nm).

Fig. 13. The photos of the filter paper containing probe 1 after exposure to CN− solution at different concentrations under UV light (365 nm).

3.9. CN− detection on filter paper HSO4 − , CN− . Only CN− could switch on the emission color from dark
to green. Moreover, CN− at different concentrations were tested by
To verify if this approach could be utilized as a portable sens- the paper-based sensor. As depicted in Fig. 13, dropping the solu-
ing device, the paper test strips were constructed. As shown in tion of CN− at different concentrations from 5 ␮M to 100 ␮M on the
Fig. 12, after dropping the solution of probe 1 (100 ␮M) on the neu- filter paper containing probe 1, the green emission can be clearly
tral filter paper and drying them by air, no fluorescence emission observed, the more concentrated of the CN− solutions, the brighter
was observed under UV light (365 nm). The filter paper contain- colors are observed on the filter paper. These results demonstrated
ing probe 1 were carefully exposed to different anions solutions that probe 1 could selectively detect CN− based on a portable device
(100 ␮M), such as F− , Cl− , Br− , I− , ClO4 − , AcO− , H2 PO4 − , NO3 − , by naked eye.
Y. Yu et al. / Sensors and Actuators B 255 (2018) 3170–3178 3177

4. Conclusions and selective probes for cyanide detection, Nanotechnology 27 (2016)


475505–475513.
[19] K. Prakash, P.R. Sahoo, S. Kumar, A substituted spiropyran for highly sensitive
In summary, a BODIPY-hemicyanine based turn-on fluorescent and selective colorimetric detection of cyanide ions, Sens. Actuators B 237
probe 1 has been designed and easily synthesized for sensitive and (2016) 856–864.
selective CN− detection. The detection limit was calculated to be [20] L.G. Nandia, C.R. Nicoleti, V.G. Marinia, I.C. Bellettini, S.R. Valandro, C.C.S.
Cavalheiro, V.G. Machado, Optical devices for the detection of cyanide in
59 nM in aqueous media. Nucleophilic addition of CN− to indolium water based on ethyl(hydroxyethyl)cellulose functionalized with perichromic
interrupted the ␲-conjugation system and restore fluorescence of dyes, Carbohydr. Polym. 157 (2017) 1548–1556.
BODIPY unit. Combined its fast kinetics and wide pH range, probe [21] A.J. Beneto, A. Siva, Highly selective colorimetric detection of cyanide anions
in aqueous media by triphenylamine and phenanthro(9,10-d)imidazole based
1 could be a good approach for the real-time, on-spot, selective
probes, Photochem. Photobiol. Sci. 16 (2017) 255–261.
and sensitive detection of CN− in the complicated systems. More- [22] F. Wang, L. Wang, X.Q. Chen, J.Y. Yoon, Recent progress in the development of
over, probe 1 was successfully applied to detect CN− in real water fluorometric and colorimetric chemosensors for detection of cyanide ions,
Chem. Soc. Rev. 43 (2014) 4312–4324.
samples and on paper test strips with satisfactory results.
[23] N. Maurya, S. Bhardwaj, A.K. Singh, A modest colorimetric chemosensor for
investigation of CN- insemi-aqueous environment with high selectivity and
sensitivity, Sens. Actuators B 229 (2016) 483–491.
Acknowledgement [24] N. Kumari, S. Jha, S. Bhattacharya, Colorimetric probes based on
anthraimidazolediones for selective sensing of fluoride and cyanide ion via
This research work was financially supported by National Nat- intramolecular charge transfer, J. Org. Chem. 76 (2011) 8215–8222.
[25] S. Kumar, P. Singh, G. Hundal, M.S. Hundal, S. Kumar, A chemodosimeter for
ural Science Foundation of China (No. 21708015). ratiometric detection of cyanide in aqueous media and human blood serum,
Chem. Commun. 49 (2013) 2667–2669.
[26] L.Y. Wang, L.H. Zhu, D.R. Cao, A colorimetric probe based on
Appendix A. Supplementary data diketopyrrolopyrrole and tert-butyl cyanoacetate for cyanide detection, New
J. Chem. 39 (2015) 7211–7218.
[27] G.M. Zhang, Y.Y. Qiao, T. Xu, C.H. Zhang, Y. Zhang, L.H. Shi, S.M. Shuang, C.
Supplementary data associated with this article can be found, in
Dong, Highly selective and sensitive nanoprobes for cyanide based on gold
the online version, at https://doi.org/10.1016/j.snb.2017.09.142. nanoclusters with red fluorescence emission, Nanoscale 7 (2015)
12666–12672.
[28] Y. Xu, X. Dai, B.X. Zhao, A coumarin-indole based colorimetric and turn on
References fluorescent probe for cyanide, Spectrochim. Acta. A 138 (2015) 164–168.
[29] A. Kumar, H.S. Kim, A pyrenesulfonyl-imidazolium derivative as a selective
[1] G.D. Muir, Hazards in Chemical Laboratory, The Royal Chemical Society, cyanide ion sensor in aqueous media, New J. Chem. 39 (2015) 2935–2942.
London, 1977. [30] G.J. Park, Y.W. Choi, D. Lee, C. Kim, A simple colorimetric chemosensor bearing
[2] J. Hamel, A review of acute cyanide poisoning with a treatment update, Crit. a carboxylic acid group with high selectivity for CN− , Spectrochim. Acta. A
Care Nurse 31 (2011) 72–82. 132 (2014) 771–775.
[3] C. Young, L. Tidwell, C. Anderson, Cyanide: Social, Industrial, and Economic [31] T.D. Ashton, K.A. Jolliffe, F.M. Pfeffer, Luminescent probes for the bioimaging
Aspects, Minerals, Metals, and Materials Society, Warrendale, 2001. of small anionic species in vitro and in vivo, Chem. Soc. Rev. 44 (2015)
[4] WHO Guidelines for Drinking-Water Quality, World Health Organisation, 4547–4595.
Geneva, Switzerland, 2011, pp. p 342. [32] J.F. Xu, H.H. Chen, Y.Z. Chen, Z.J. Li, L.Z. Wu, C.H. Tung, Q.Z. Yang, A
[5] K. Minakata, H. Nozawa, K. Gonmori, I. Yamagishi, M. Suzuki, K. Hasegawa, K. colorimetric and fluorometric dual-modal chemosensor for cyanide in water,
Watanabe, O. Suzuki, Determination of cyanide in blood by electrospray Sens. Actuators B 168 (2012) 14–19.
ionization tandem mass spectrometry after direct injection of dicyanogold, [33] A.A. Biradar, A.V. Biradar, T. Sun, Y. Chan, X.X. Huang, T. Asefa, Bicinchoninic
Anal. Bioanal. Chem. 400 (2011) 1945–1951. acid-based colorimetric chemosensor for detection of low concentrations of
[6] M.K. Chahal, M. Sankar, Porphyrin chemodosimeters: synthesis, cyanide, Sens. Actuators B 222 (2016) 112–119.
electrochemical redox properties and selective ‘naked-eye’ detection of [34] X.D. Lou, D.X. Ou, Q.Q. Li, Z. Li, An indirect approach for anion detection: the
cyanide ions, RSC Adv. 5 (2015) 99028–99036. displacement strategy and its application, Chem. Commun. 48 (2012)
[7] O. Destanoglu, G.G.Y. Ilmaz, R. Apak, Selective determination of free cyanide 8462–8477.
in environmental water matrices by ion chromatography with suppressed [35] A. Bencini, V. Lippolis, Metal-based optical chemosensors for CN− detection,
conductivity detection, J. Liq. Chromatogr. R. T. 38 (2015) 1945–1951. Environ. Sci. Pollut. Res. 23 (2016) 24451–24475.
[8] H.M. Al-Saidi, S.A. Al-Harbi, E.H. Aljuhani, M.S. El-Shahawi, Headspace [36] J.H. Lee, A.R. Jeong, I.-S. Shin, H.-J. Kim, J.-I. Hong, Fluorescence turn-on sensor
sorptive solid phase microextraction (HS-SPME) combined with a for cyanide based on a cobalt(II)- coumarinylsalencomplex, Org. Lett. 12
spectrophotometry system: a simple glass devise for extraction and (2010) 764–767.
simultaneous determination of cyanide and thiocyanate in environmental [37] I. Bhowmick, D.J. Boston, R.F. Higgins, C.M. Klug, M.P. Shores, T. Gupta, Naked
and biological samples, Talanta 159 (2016) 137–142. eye detection of cyanide in water with CoII bis(terpyridine)complexes, Sens.
[9] J.M. Zhao, H.C. Bi, Y. Li, H.C. Bi, Flow injection on-line distillation method for Actuators B 235 (2016) 325–329.
determination of cyanide in water, Adv. Mater. Res. 1073–1076 (2015) [38] L.A. Greenawald, J.L. Snyder, N.L. Fry, M.J. Sailor, G.R. Boss, H.O. Finklea, S. Bell,
197–201. Development of a cobinamide-based end-of-service-life indicator for
[10] S.T. Wang, Y.W. Sie, C.F. Wan, A.T. Wu, A reaction-based fluorescent sensor for detection of hydrogen cyanide gas, Sens. Actuators B 221 (2015) 379–385.
detection of cyanide in aqueous media, J. Lumin. 173 (2016) 25–29. [39] H. Yoon, C.H. Lee, Y.H. Jeong, H.C. Gee, W.D. Jang, A zinc porphyrin-based
[11] T.B. Wei, G.T. Yan, H. Li, Y.R. Zhu, B.B. Shi, Q. Lin, H. Yao, Y.M. Zhang, A highly molecular probe for the determination of contamination in commercial
sensitive and selective ‘turn-on’ fluorescence sensor for rapid detection of acetonitrile, Chem. Commun. 48 (2012) 5109–5111.
cyanide ions in aqueous solution, Supramol. Chem. 28 (2016) 720–726. [40] M.E. Jun, B. Roy, K.H. Ahn, Turn-on fluorescent sensing with reactive probes,
[12] Y. Shiraishi, M. Nakamura, N. Hayashi, T. Hirai, Coumarin-spiropyran dyad Chem. Commun. 47 (2011) 7583–7601.
with a hydrogenated pyran moiety for rapid, selective, and sensitive [41] Y.Q. Dong, R.X. Wang, W.R. Tian, Y.W. Chi, G.N. Chen, Turn-on fluorescent
fluorometric detection of cyanide anion, Anal. Chem. 88 (2016) 6805–6811. detection of cyanide based on polyamine-functionalized carbon quantum
[13] W.J. Qu, T.B. Wei, Q. Lin, W.T. Li, J.X. Su, G.Y. Liang, Y.M. Zhang, A recyclable dots, RSC Adv. 4 (2014) 3701–3705.
probe for highly selective and sensitive detection of cyanide anion in aqueous [42] C. Cheng, H.Y. Chen, C.S. Wu, J.S. Meena, T. Simon, F.H. Ko, A highly sensitive
medium by fluorescent and colorimetric changes, Sens. Actuators B 232 and selective cyanide detection using a gold nanoparticle-based dual
(2016) 115–124. fluorescence-colorimetric sensor with a wide concentration range, Sens.
[14] C.X. Liu, X.Y. Wang, H.W. Wu, Y.Y. Chen, Turn-on fluorescence detection of Actuators B 227 (2016) 283–290.
cyanide with large stokes shift: activation of latent intramolecular Br/HeC [43] Y.B. Chen, W. Shi, Y.H. Hui, X.H. Sun, L.X. Xu, L. Feng, Z.F. Xie, A new highly
hydrogen bonding, Dyes Pigm. 133 (2016) 255–260. selective fluorescent turn-on chemosensor for cyanide anion, Talanta 137
[15] R. Kaushik, A. Ghosh, A. Singh, P. Gupta, A. Mittal, D.A. Jose, Selective (2015) 38–42.
detection of cyanide in water and biological samples by an Off-the-Shelf [44] W.C. Lin, J.W. Hu, K.W. Chen, A ratiometric chemodosimeter for highly
compound, ACS Sens. 1 (2016) 1265–1271. selective naked-eye and fluorogenic detection of cyanide, Anal. Chim. Acta
[16] J.W. Hu, W.C. Lin, S.Y. Hsiao, Y.H. Wu, H.W. Chen, K.Y. Chen, An 893 (2015) 91–100.
indanedione-based chemodosimeter for selective naked-eye and fluorogenic [45] B. Garg, L.Y. Yan, T. Bisht, C.Y. Zhu, Y.C. Ling, A phenothiazine-based
detection of cyanide, Sens. Actuators B 233 (2016) 510–519. colorimetric chemodosimeter for the rapid detection of cyanide anions in
[17] A.D.S. Schramm, R. Menger, V.G. Machado, Malononitrile-derivative organic and aqueous media, RSC Adv. 4 (2014) 36344–36349.
chromogenic devices for the detection of cyanide in water, J. Mol. Liq. 223 [46] Q.S. Zhang, J. Zhang, H.J. Zuo, C.Y. Wang, Y.J. Shen, A novel colorimetric and
(2016) 811–818. fluorescent sensor for cyanide anions detection based on triphenylamine and
[18] N. Vasimalai, M.T. Fernandez-Arguelles, Novel one-pot and facile room benzothiadiazole, Tetrahedron 72 (2016) 1244–1248.
temperature synthesis of gold nanodots and application as highly sensitive
3178 Y. Yu et al. / Sensors and Actuators B 255 (2018) 3170–3178

[47] L.Y. Wang, L.Q. Li, D.R. Cao, A BODIPY-based dye with red fluorescence in solid [71] S. Madhu, S.K. Basu, S. Jadhav, M. Ravikanth, 3,
state and used as a fluorescent and colorimetric probe for highly selective 5-Diformyl-borondipyrromethene for selective detection of cyanide anion,
detection of cyanide, Sens. Actuators B 239 (2017) 1307–1317. Analyst 138 (2013) 299–306.
[48] S.H. Mashraqui, R. Betkar, M. Chandiramani, C. Estarellas, A. Frontera, Design [72] Z. Ekmekci, M.D. Yilmaz, E.U. Akkaya, A. Monostyryl-boradiazaindacene,
of a dual sensing highly selective cyanide chemodosimeter based on (BODIPY) derivative as colorimetric and fluorescent probe for cyanide ions,
pyridinium ring chemistry, New J. Chem. 35 (2011) 57–60. Org. Lett. 3 (2008) 461–464.
[49] M. Tomasulo, S. Sortino, A.J.P. White, F.M. Raymo, Chromogenic oxazines for [73] N. Niamnont, A. Promchat, C. Siangm, C. Pramaulpornsatit, M.
cyanide detection, J. Org. Chem. 71 (2006) 744–753. Sukwattanasinitt, A novel phenylacetylene −indolium fluorophore for
[50] P. Wang, Y. Yao, M. Xue, A novel fluorescent probe for detecting paraquat and detection of cyanide by the naked eye, RSC Adv. 5 (2015) 64763–64768.
cyanide in water based on pillar[5] arene/10-methylacridinium iodide [74] C.H. Lee, H.J. Yoon, J.S. Shim, W.D. Jang, A boradiazaindacene-based turn-on
molecular recognition, Chem. Commun. 50 (2014) 5064–5067. fluorescent probe for cyanide detection in aqueous media, Chem. Eur. J. 18
[51] R. Sukato, N. Sangpetch, T. Palaga, S. Jantra, V. Vchirawongkwin, C. Jongwohan, (2012) 4513–4516.
M. Sukwattanasinitt, S. Wacharasindhu, New turn-on fluorescent and [75] Y. Jhong, W.H. Hsieh, J.L. Chir, A.T. Wu, A highly selective and turn-on
colorimetric probe for cyanide detection based on BODIPY-salicylaldehyde fluorescence sensor for detection of cyanide, J. Fluoresc. 24 (2014) 723–1726.
and its application in cell imaging, J. Hazard. Mater. 314 (2016) 277–285. [76] Z.P. Lin, X.Q. Wang, Z.H. Yang, W.J. He, Rational design of a dual chemosensor
[52] Y.Q. Hao, W.S. Chen, L.Q. Wang, B.B. Zhou, Q.G. Zang, S. Chen, Y.N. Liu, A for cyanide anion sensing based on dicyanovinyl-substituted benzofurazan, J.
naphthalimide-based azo colorimetric and ratiometric probe: synthesis and Org. Chem. 76 (2011) 10286–10290.
its application in rapid detection of cyanide anions, Anal. Methods 6 (2014) [77] L.Y. Wang, J.Q. Du, D.R. Cao, A colorimetric and fluorescent probe containing
2478–2483. diketopyrrolopyrroleand 1,3-indanedione for cyanide detection based on
[53] M. Jamkratoke, V. Ruangpornvisuti, G. Tumcharern, T. Tuntulani, B. exciplex signaling mechanism, Sens. Actuators B 198 (2014) 455–461.
Tomapatanaget, A-D-A sensors based on naphthoimidazoledione and boronic [78] Q. Li, Y. Cai, H. Yao, Q. Lin, Y.R. Zhu, H. Li, Y.M. Zhang, T.B. Wei, A colorimetric
acid as turn-on cyanide probes in water, J. Org. Chem. 74 (2009) 3919–3922. and fluorescent cyanide chemosensor based on dicyanovinyl derivatives:
[54] A. Promchat, P. Rashatasakhon, M. Sukwattanasinitt, A novel indolium salt as utilization of the mechanism of intramolecular charge transfer blocking,
a highly sensitive and selective fluorescent sensor for cyanide detection in Spectrochim. Acta. A 136 (2015) 1047–1051.
water, J. Hazard. Mater. 329 (2017) 255–261. [79] R.K. Konidena, K.R.J. Thomas, Selective naked-eye cyanide detection in
[55] X.H. Huang, X.G. Gu, G.X. Zhang, D.Q. Zhang, A highly selective fluorescence aqueous media using a carbazole-derived fluorescent dye, RSC Adv. 4 (2014)
turn-on detection of cyanide based on the aggregation of tetraphenylethylene 22902–22910.
molecules induced by chemical reaction, Chem. Commun. 48 (2012)
12195–12197.
[56] J.B. Chao, Z.Q. Li, Y.B. Zhang, F.J. Huo, C.X. Yin, H.B. Tong, Y.H. Liu, A ratiometric Biographies
fluorescence probe for monitoring cyanide ion in live cells, Sens. Actuators B
228 (2016) 192–199.
[57] H.J. Kim, K.C. Ko, J.H. Lee, J.Y. Lee, J.S. Kim, KCN sensor: unique chromogenic Dr. Yanhua Yu has earned her MS degree under the supervision of Prof. Jie Tang at
and ‘turn-on’ fluorescent chemodosimeter: rapid response and high East China Normal University in 2010 and PhD degree under the supervision of Prof.
selectivity, Chem. Commun. 47 (2011) 2886–2888. Joanne Xie at École normale supérieure de Cachan (France) in 2013. She then moved
[58] M.J. Peng, Y. Guo, X.F. Yang, F. Suzenet, J. Li, C.W. Li, Y.W. Duan, to Jianghan University as Assistant Researcher. Her research field refers to analyt-
Coumarin-hemicyanine conjugates as novel reaction-based sensors for ical chemistry and organic chemistry. Her research subject of PhD focused on the
cyanide detection: convenient synthesis and ICT mechanism, RSC Adv. 4 rational design and synthesis of fluorescent molecules based on benzothiadiazole,
(2014) 19077–19085. coumarin, BODIPY and DCM with click reaction and investigation on their applica-
[59] F.J. Huo, J. Kang, C.X. Yin, J.B. Chao, Y.B. Zhang, A turn on fluorescent sensor for tions in biology and analytical chemistry. In Jianghan University, her research work
cyanide based on ICT off in aqueous and its application for bioimaging, Sens. is focused on fluorescent probes and fluorescent peptide synthesis.
Actuators B 215 (2015) 93–98.
Dr. Yibin Ruan, currently a staff of Technology Center of China Tobacco Guizhou
[60] Y.K. Yue, F.J. Huo, C.X. Yin, J.B. Chao, Y.B. Zhang, A new donor-two-acceptor
Industrial Co. Ltd., worked with Prof. Yun-Bao Jiang at Xiamen University for his MS
red emission fluorescent probe for highly selective and sensitive detection of
degree. He received his PhD degree under the supervision of Dr. Isabelle Leray and
cyanide in living cells, Sens. Actuators B 212 (2015) 451–456.
Prof. Joanne Xie at École normale supérieure de Cachan (France). And then he spent 6
[61] N. Niamnont, A. Promchat, C. Siangm, C. Pramaulpornsatit, M.
months as a post-doc to work in Laboratoire de Spectrochimie Infrarouge et Raman
Sukwattanasinitt, A novel phenylacetylene −indolium fluorophore for
with Dr. Stéphane Aloïse. His research interests are the development of molecular
detection of cyanide by the naked eye, RSC Adv. 5 (2015) 64763–64768.
sensors for anionic and cationic species and photophysics of metal-complex in the
[62] L. Li, Y. Zhang, Z.M. Chang, F.Q. Bai, H.X. Zhang, J.K. Ferri, W.F. Dong,
excited state.
Theoretical study on fluorescent probes for cyanide based on the indolium
functional group, Org. Electron. 30 (2016) 1–31. Prof. Changzhi Dong has earned his BS and MS degrees from Beijing University in
[63] Y.H. Yu, T.T. Shu, C. Fu, B.J. Yu, D.D. Zhang, H.X. Luo, J.J. Chen, C.Z. Dong, A 1983 and 1986 respectively. He moved then to Yantai University as Assistant Pro-
novel colorimetric sensor based on BODIPY-coumarin dye for simultaneous fessor. In 1990, he was awarded of an one-year scholarship from the French Atomic
detection of cyanide and fluoride, J. Lumin. 186 (2017) 212–218. Energy Commissariat (CEA) to work as a visiting scholar in CEA Cadarache Research
[64] D. Udhayakumari, S. Velmathi, M.S. Boobalan, Novel chemosensor for Centre. He pursued then in France his education and defended successfully his PhD
multiple target anions: the detection of F− and CN− ion via different thesis in 1996 in the School of Pharmacy, Paris V University with the major in medic-
approach, J. Fluorine Chem. 175 (2015) 180–184. inal chemistry under the supervision of Dr. Bernard P. Roques. He spent then one
[65] L.Y. Wang, L.Q. Li, D.R. Cao, Synthesis photoluminescence, chromogenic and year as a post-doc to work in Dr. Marc Julia’s lab in the Chemistry Department of
fluorogenic discrimination of fluoride and cyanide based on a “Ecole Normale Supérieure de Paris (ENS Paris)”. He was appointed as assistant pro-
triphenylamine-tri(2-formyl BODIPY) conjugate, Sens. Actuators B 241 (2017) fessor in 1997 and full professor in 2011 in Paris VII University. From 2001–2003,
1224–1234. he has worked as a research associate in Ziwei Huang’s lab in UIUC, USA.S ince
[66] A. Loudet, K. Burgess, BODIPY dyes and their derivatives: syntheses and 2011, he has been the PI of the Peptide and Protein Chemistry Center of the Institute
spectroscopic properties, Chem. Rev. 107 (2007) 4891–4932. for Interdisciplinary Research, Jianghan University. The research interests of Prof.
[67] N. Boens, V. Leen, W. Dehaen, Fluorescent indicators based on BODIPY, Chem. Dong mainly focus on two domains: drug development and peptide/protein chem-
Soc. Rev. 41 (2012) 1130–1172. istry. In the first one, he has worked on design, synthesis and SAR study of small
[68] G. Ulrich, R. Ziessel, A. Harriman, The chemistry of fluorescent bodipy dyes: organic compounds for the treatment of different diseases, such as AIDS, AD, cancer
versatility unsurpassed, Angew. Chem. Int. Ed. 47 (2008) 1184–1201. and inflammation.In the second one, he has performed total chemical synthesis of
[69] J.S. Lee, H.K. Kim, S. Feng, M. Vendrell, Y.T. Chang, Accelerating fluorescent certain small proteins with full biological activities and their derivatives and these
sensor discovery: unbiased screening of a diversity-oriented BODIPY library, works provided powerful tools for the structure-function relationship study of these
Chem. Commun. 47 (2011) 2339–2341. proteins. He is recently interested also in the development of chemical sensors for
[70] Y.H. Yu, N. Bogliotti, J. Tang, X. Juan, Synthesis and properties of carbohydrate detecting ions, small molecules or peptides.
based BODIPY-functionalized, Eur. J. Org. Chem. (2013) 7749–7760.

You might also like