You are on page 1of 5

Article

Cite This: Anal. Chem. XXXX, XXX, XXX−XXX pubs.acs.org/ac

Colorimetric and Fluorescent Detecting Phosgene by a Second-


Generation Chemosensor
Ying Hu,†,‡ Xin Zhou,*,†,§ Hyeseung Jung,‡ Sang-Jip Nam,‡ Myung Hwa Kim,*,‡ and Juyoung Yoon*,‡

Department of Chemistry and Nano Science, Ewha Womans University, Seoul 120-750, Korea
§
College of Chemistry and Chemical Engineering, Shandong Sino-Japanese Center for Collaborative Research of Carbon
Nanomaterials, Laboratory of Fiber Materials and Modern Textile, The Growing Base for State Key Laboratory, Qingdao University,
Shandong 266071, People’s Republic of China
*
S Supporting Information

ABSTRACT: Because of the current shortage of first-


generation phosgene sensors, increased attention has been
given to the development of fluorescent and colorimetric based
methods for detecting this toxic substance. In an effort focusing
on this issue, we designed the new, second-generation
phosgene chemosensor 1 and demonstrated that it undergoes
a ring-opening reaction with phosgene in association with color
and fluorescent changes with a detection limit of 3.2 ppb.
Notably, in comparison with the first-generation sensor RB-
OPD, 1 not only undergoes a much faster response toward phosgene with an overall response time within 2 min, but it also
generates no byproducts during the sensing process. Finally, sensor 1 embedded nanofibers were successfully fabricated and used
for accurate and sensitive detection of phosgene.

P hosgene is a highly toxic gas that is widely used in the


chemical industry for the production of isocyanate-based
polymers or pharmaceutical compounds.1,2 This substance is
Furthermore, sensor 1 embedded nanofibers were successfully
fabricated and used for concise detecting phosgene via a
fluorescent and colorimetric manner.
also receiving great attention owing to its potential use as a
bioterrorism agent. Thus, a strong need exists to develop
concise methods for sensing phosgene gas. Fluorescence and
■ EXPERIMENTAL SECTION
General Methods. Unless otherwise stated, all materials
color-based chemical sensors are of current interest owing to were obtained from commercial suppliers and used without
several advantageous characteristics such as high sensitivity and further purification. Thin layer chromatography (TLC) was
rapid rates of response.3−20 However, in contrast to the wide carried out using Merck 60 F254 plates with a thickness of 0.25
number of fluorescence and color sensors for nerve gas agents mm. Preparative TLC was performed using Merck 60 F254
that have been reported,21−42 only a few efforts have focused on plates with a thickness of 1 mm. Flash chromatography was
developing sensors for phosgene.43−53 carried out on silica gel (230−400 mesh). The 1H and 13C
Previously, we described the sensor PY-OPD (Scheme 1a), NMR spectra were recorded using Bruker 300 MHz or Varian
which utilizes an o-phenylenediamine moiety (OPD) as a 500 MHz spectrometer. Chemical shifts were given in ppm and
universal reacting group, that enables discrimination between the coupling constants (J) in Hz. The mass spectra were
phosgene and nerve gas mimics.54 Inspired by this strategy, obtained using a JMS-HX 110A/110A Tandem Mass
many robust sensors including NBD-OPD, NAP-OPD, and Spectrometer (JEOL). The UV absorption spectra were
RB-OPD (Scheme 1a) were designed for detecting phosgene obtained on a UVIKON 933 Double Beam UV/vis
by using nanofibers and test papers.47−51 However, these first spectrometer. The fluorescence emission spectra were obtained
generation sensors share the common disadvantage that they using an RF-5301/PC spectrofluorophotometer (Shimadzu).
generate hydrochloric acid (HCl) upon reaction with phosgene Synthesis of 1. To a solution of RB-OPD (0.9 mmol) in
(Scheme 1b). This product is a secondary pollutant that could dry THF (10 mL) in an ice bath was added lithium aluminum
cause detection errors like, for example, promotion of a ring- hydride (4.5 mmol), stirring at room temperature for 24 h
open reaction of the rhodamine dye moiety RB-OPD.55,56 under N2 protection. After quenching by adding 1-butanol, the
Thus, to overcome this shortage of the first-generation mixture was concentrated under vacuum. The residue was
phosgene sensors, we presented here our effort on design of the diluted with water and extracted with ethyl acetate, and the
second-generation phosgene sensor 1. As shown in Scheme 1c,
sensor 1, a benzimidazole-fused rhodamine dye, can undergo a Received: December 2, 2017
ring-open process with phosgene, resulting in a colorimetic and Accepted: February 7, 2018
fluorescent response with a detection limit as low as 3.2 ppb. Published: February 7, 2018

© XXXX American Chemical Society A DOI: 10.1021/acs.analchem.7b05011


Anal. Chem. XXXX, XXX, XXX−XXX
Analytical Chemistry Article

Scheme 1. (a) Chemical Structures of the First-Generation aluminum plate located 15 cm below the tip of the needle. The
Sensors; (b) Reaction between RB-OPD and Phosgene; (c) flow rate was carefully kept at 0.5 mL/h. After collection, the
Strategy for the Design of the Second-Generation Sensor 1 polymer fibers were dried in a 40 °C vacuum oven for 1 d to
prevent agglomeration. The structures and morphologies of the
fibers were assessed using electron microscopy (FE-SEM, JEOL
JSM-6700F).

■ RESULTS AND DISCUSSION


Synthesis of Sensor 1. Sensor 1 was synthesized in a total
yield of 67% using a route, which was initiated by preparation
of intermediate RB-OPD from rhodamine B, employing a
reported procedure.3 Then RB-OPD was reacted with lithium
aluminum hydride in dry THF to generate 1 as a light pink
solid after chromatographic purification (detailed procedure
and 1H and 13C NMR, as well as mass data are given in the
Supporting Information).
Phosgene Promoted Color and Fluorescence
Changes. The fluorescent response of 1 toward phosgene
was investigated first by using triphosgene, a nonvolatile
precursor of this gaseous substance. Inspection of the spectra
displayed in Figure 1a shows that addition of triphosgene to a

extracts were dried over anhydrous MgSO4 and concentrated


under a vacuum to give a residue that was subjected to column
chromatography to give sensor 1 as pale pink in 80% yield. 1H
NMR (CDCl3, 300 MHz) δ (ppm): 8.13 (d, J = 7.2 Hz, 1H);
7.83 (d, J = 8.4 Hz, 1H); 7.51 (t, J = 7.2 Hz, 1H); 7.40 (t, J =
7.5 Hz, 1H); 7.26 (d, J = 7.5 Hz, 1H); 7.18 (t, J = 8.4 Hz, 1H);
7.04 (t, J = 7.2 Hz, 1H); 6.97 (d, J = 7.2 Hz, 1H); 6.51 (s, 2H);
6.33 (d, J = 9.0 Hz, 2H); 6.18 (d, J = 6.3 Hz, 2H); 3.34 (q, J =
7.2 Hz, 8H); 1.16 (t, J = 7.2 Hz, 12H). 13C NMR (CDCl3, 75
MHz) δ (ppm): 156.87, 156.53, 153.30, 149.18, 149.06, 131.15,
130.66, 128.84, 128.52, 128.19, 125.08, 122.75, 122.14, 121.59,
120.35, 110.44, 108.37, 106.53, 98.11, 44.57, 12.84. HRMS
(ESI) m/z = 515.2823 [M + H]+, calcd for C34H36N4O2 =
515.2733. Figure 1. (a) Fluorescence spectra of sensor 1 in chloroform (10 μM)
Fluorescent Study. Stock solutions of sensor 1 (1 mM) upon gradual addition of a solution of triphosgene (0−1.0 equiv) in
were prepared in chloroform. Triphosgene stock solution (10 chloroform (λex = 530 nm, slits: 3 nm × 1.5 nm). (b) Time-dependent
mM) in chloroform were prepared. The test solutions were spectra of sensor 1 (red) and RB-OPD (black) in chloroform solution
prepared by placing 30 μL of the sensor stock solution into a (10 μM) following addition of a chloroform solution of triphosgene
test tube, adding an appropriate aliquot of triphosgene, and (0.5 equiv), respectively (λex = 530 nm, λem = 578 nm, Slits: 3 nm ×
diluting these stock solutions to 3 mL with chloroform. 1.5 nm). (c) Absorbance spectra of sensor 1 in chloroform (10 μM)
For all of the measurements of the fluorescence spectra, upon gradual addition of a solution of triphosgene (0−1.0 equiv) in
chloroform. (d) Colorimetric responses of sensor 1 toward different
excitation was performed at 530 nm with slit widths for
equivalents of triphosgene.
excitation and emission are 1.5 and 3.0 nm, respectively. UV/
vis and fluorescence titration experiments were performed
using 10 μM of 1 in chloroform with varying concentrations of
triphosgene at room temperature. CHCl3 solution of 1 causes a significant concentration
Preparation of Sensor 1 Embedded Electrospun dependent enhancement in the fluorescence intensity at 578
Fibers. To prepare a precusor solution for synthesizing the nm accompanied by a small red shift in the emission
electrospun fibers, 1.0 mg of sensor 1 was dissolved in 2.5 mL wavelength maximum. Accordingly, the fluorescence quantum
of acetonitrile. In each case, when the solute was completely yields (Φf) of sensor 1 enhanced from 0.06 to 0.22. The
dissolved, 100.0 mg of poly(ethylene oxide) (PEO, Mw = fluorescence increase is due to the unique spirolactam ring-
600000) as a matrix polymer was added, and the resulting opening process. The detection limit of 1 for phosgene was
mixture was stirred for 1 h and loaded into a stainless steel determined by a fit of the emission titration data to be 3.2 ppb
syringe connected to a needle of gauge 23. The distance (Figure S1). Notably, in comparison with the first-generation
between the end of the needle and grounded plate was adjusted sensor RB-OPD, 1 undergoes a much faster response upon
to be 15 cm before applying 5.0 kV. The precursor solution was addition of triphosgene with the overall sensing process being
ejected to form fine fibers on the surface of a grounded accomplished within 2 min (Figure 1b). This data indicates that
B DOI: 10.1021/acs.analchem.7b05011
Anal. Chem. XXXX, XXX, XXX−XXX
Analytical Chemistry Article

the second-generation sensor 1 possesses a higher reactivity 3, a unique peak at 13.51 ppm exists in the spectrum of 3 that is
that its precursor RB-OPD. to the NH proton. Meanwhile, the aromatic protons of 3
Inspection of the absorption spectrum of a CHCl3 solution
of 1 upon treatment with triphosgene (Figure 1c) shows that a
new UV band appears at 560 nm. As can be seen by viewing the
images displayed in Figure 1d, addition of triphosgene to the
CHCl3 solution of 1 causes a distinct color change from
colorless to pink, which can be easily observed by using the
naked eye.
The response of this sensor to the nerve agent mimic, DCP,
was investigated next. Unlike in the case of phosgene, the
addition of even 70 equiv DCP induces only a slight
fluorescence intensity increase (Figure S2). However, sensor
1 does display similar color and fluorescence responses to that
promoted by phosgene when treated with exceptionally high
doses of DCP (200 equiv; Figures S3 and S4).
Reaction Mechanisms. Based on the UV and fluorescent
observations, we propose the plausible mechanism shown in
Scheme 2 for the reaction of sensor 1 with triphosgene. The Figure 3. Partial 1H NMR spectra of 2 and 3 (CDCl3, 300 MHz).

Scheme 2. Mechanism of the Reaction of Sensor 1 with resonate at dramatically upfield positions in contrast to those of
Triphosgene compound 2. In addition, 13C-labeled triphosgene was used in
reaction of 1 to gain further evidence for the assignment of 2 as
the product of the reaction. As can be seen by viewing the 13C
NMR spectra in Figure 4, a remarkably high intensity peak at

product of this process was separated from the reaction


mechanism and identified in order to gain evidence to support
this mechanistic proposal. Analysis of the 1H NMR spectrum of
the product showed that it possesses the structure represented
by the ring opened N-chloroformyl-benzimadazole 2. As can be
seen by viewing the spectra of 1 and 2 shown in Figure 2, the

Figure 4. Partial 13C NMR spectra of 2 and 3 (CDCl3, 75 MHz).

143.36 ppm is present in the spectrum of 2, which is


unambiguously assigned to the 13C-labeled chloroformyl
carbonyl carbon. The results showing that 2 is formed in the
reaction of phosgene with 1 also demonstrate that, unlike
reactions with the first-generation sensors shown in Scheme 1a,
HCl is not generated.
Figure 2. Partial 1H NMR spectra of sensor 1 (top) and compound 2
Sensor 1-Embedded Polymer Fibers. To generate a
(CDCl3, 300 MHz).
format in which 1 can be employed for sensing phosgene gas,
the electrospinning technique was used to fabricate sensor 1-
signals for H8′, H9′, H10′, and H11′ in 2 are downfield shifted embedded fibers. For this purpose, a mixture containing 1 and
in comparison to their counterparts in 1, as a consequence of the matrix polymer poly(ethylene oxide) (PEO, Mw = 600000)
the presence of the electron withdrawing formyl chloride group. was jet ejected from a syringe containing a needle to which a
To exclude the possibility that the product of the reaction of voltage of 5.0 kV is applied. This procedure produced
1 with phosgene is simply a ring-opened protonated species, uniformly distributed fibers containing a minimum number of
sensor 1 was reacted with TFA to form the nonchloroformy- bead structures. The polymer fibers containing 1 display a
lated-benzimidazole 3. This substance was characterized by highly sensitive color as well as fluorescence response upon
using 1H NMR, 13C NMR, and ESI-mass spectroscopy. As exposure to phosgene. As shown in Figure 5, the white colored
shown by comparing the 1H NMR spectra displayed in Figure and nonfluorescent (under a UV lamp, 365 nm) sensor 1-
C DOI: 10.1021/acs.analchem.7b05011
Anal. Chem. XXXX, XXX, XXX−XXX
Analytical Chemistry Article

Figure 5. Color/fluorescent responses of sensor 1 based PEO


nanofiber upon exposure to phosgene (0.8 mg/L phosgene gas).
Figure 7. Confocal microscope images of sensor 1-based PEO
embedded fibers upon exposure to phosgene for several nanofiber upon exposure to phosgene (0.8 mg/L phosgene gas).
seconds change color to dark pink and emit pink fluorescence.
In Figure 6 are shown scanning electron microscope (SEM)
images of sensor 1-embedded nanofibers before and after to design 1 will accelerate the development of new sensors
protocols.


*
ASSOCIATED CONTENT
S Supporting Information
The Supporting Information is available free of charge on the
ACS Publications website at DOI: 10.1021/acs.anal-
chem.7b05011.
The NMR and MS spectra of sensor 1 and other
materials (PDF).

■ AUTHOR INFORMATION
Corresponding Authors
*Fax: (+82) 2-3277-2384. E-mail: jyoon@ewha.ac.kr.
Figure 6. Scanning electron microscope images of sensor 1-based PEO *E-mail: myungkim@ewha.ac.kr.
nanofibers before and after exposure to phosgene (0.8 mg/L phosgene *E-mail: zhouxin@qdu.edu.cn.
gas). ORCID
Myung Hwa Kim: 0000-0001-7254-2886
exposure to phosgene. As can be seen, the initial embedded Juyoung Yoon: 0000-0002-1728-3970
fibers have a uniform structure with a diameter of about 1.0 μm. Author Contributions
After exposure to phosgene, the surfaces of the fibers become †
These authors contribute equally to this work.
rough and their diameters become irregular. These observations Notes
indicate that reaction of 1 with phosgene on the surface of The authors declare no competing financial interest.


nanofibers causes partial deformation of the in PEO polymer
matrix. ACKNOWLEDGMENTS
Confocal microscopy was employed to examine the sensor 1-
embedded nanofiber before and after phosgene exposure J.Y. acknowledges a grant from the National Creative Research
(Figure 7). The fibers exposed to phosgene were observed to Initiative programs of the National Research Foundation of
display bright red fluorescence using this imaging technique. Korea (NRF) funded by the Korean government (MSIP; No.


2012R1A3A2048814). M.H.K. thanks to the National Research
CONCLUSION Foundation of Korea (NRF) funded by the Korean government
(MSIP; No. 2016R1D1A1B03934962). X.Z. acknowledges the
In conclusion, in the studies described above we developed the Natural Science Foundation of China (NSFC. 21662037,
new second-generation sensor 1 for color and fluorescence 21762045), the Youth Science Foundation of Jilin Province
sensing of phosgene. Sensor 1 was found to undergo a (20160520003JH), for supporting this work. Mass spectral data
spirocyclic ring-open reaction with phosgene, which is were obtained from the Korea Basic Science Institute (Daegu)
associated with both a color change and fluorescence using a Jeol JMS 700 high-resolution mass spectrometer.


enhancement, and a detection limit as low as 3.2 ppb. Notably,
in comparison with that of the first-generation sensor RB-OPD, REFERENCES
reaction of 1 with phosgene is much more rapid corresponding (1) Ryan, T. A.; Seddon, E. A.; Seddon, K. R.; Ryan, C. Phosgene and
to a time for complete reaction of 2 min. More importantly, Related Carbonyl Halides; Elsevier: Amsterdam, 1996.
reaction of sensor 1 with phosgene does not generate HCl. (2) Zhou, X.; Lee, S.; Xu, Z.; Yoon, J. Chem. Rev. 2015, 115, 7944−
Finally, sensor 1 embedded nanofibers were successfully 8000.
fabricated and used for detecting phosgene in a fluorescent (3) Yang, Y. C.; Baker, J. A.; Ward, J. R. Chem. Rev. 1992, 92, 1729−
and colorimetric manner. We believe that the strategy utilized 1743.

D DOI: 10.1021/acs.analchem.7b05011
Anal. Chem. XXXX, XXX, XXX−XXX
Analytical Chemistry Article

(4) Bartelt-Hunt, S. L.; Knappe, D. R. U.; Barlaz, M. A. Crit. Rev. (38) Dale, T. J.; Rebek, Jr J. Am. Chem. Soc. 2006, 128, 4500−4501.
Environ. Sci. Technol. 2008, 38, 112−136. (39) Zhu, R.; Azzarelli, J. M.; Swager, T. M. Angew. Chem., Int. Ed.
(5) Palit, M.; Pardasani, D.; Gupta, A. K.; Dubey, D. K. Anal. Chem. 2016, 55, 9662−9666.
2005, 77, 711−717. (40) Chen, S.; Ruan, Y.; Brown, J. D.; Gallucci, J.; Maslak, V.; Hadad,
(6) Mäkinen, M. A.; Anttalainen, O. A.; Sillanpäa,̈ M. E. T. Anal. C. M.; Badjić, J. D. J. Am. Chem. Soc. 2013, 135, 14964−14967.
Chem. 2010, 82, 9594−9600. (41) Belger, C.; Weis, J. G.; Egap, E.; Swager, T. M. Macromolecules
(7) Anzai, J.-i. In Handbook of Toxicology of Chemical Warfare Agents, 2015, 48, 7990−7994.
2nd ed.; Gupta, R. C., Ed.; Academic Press: Boston, 2015; pp 925− (42) Zhang, W. Q.; Cheng, K.; Yang, X.; Li, Q. Y.; Zhang, H.; Ma, Z.;
934. Lu, H.; Wu, H.; Wang, X. Org. Chem. Front. 2017, 4, 1719−1725.
(8) Richardt, A.; Jung, M.; Niemeyer, B. CBRN Protection; Wiley- (43) Zhang, H.; Rudkevich, D. M. Chem. Commun. 2007, 1238−
VCH Verlag GmbH & Co. KGaA, 2013; pp 179−209. 1239.
(9) Kim, K.; Tsay, O. G.; Atwood, D. A.; Churchill, D. G. Chem. Rev. (44) Wu, X.; Wu, Z.; Yang, Y.; Han, S. Chem. Commun. 2012, 48,
2011, 111, 5345−5403. 1895−1897.
(10) Jang, Y. J.; Kim, K.; Tsay, O. G.; Atwood, D. A.; Churchill, D. G. (45) Kundu, P.; Hwang, K. C. Anal. Chem. 2012, 84, 4594−4597.
Chem. Rev. 2015, 115, PR1−PR76. (46) Liu, Z.; Zhou, X.; Miao, Y.; Hu, Y.; Kwon, N.; Wu, X.; Yoon, J.
(11) Jang, Y. J.; Tsay, O. G.; Murale, D. P.; Jeong, J. A.; Segev, A.; Angew. Chem., Int. Ed. 2017, 56, 5812−5816.
Churchill, D. G. Chem. Commun. 2014, 50, 7531−7534. (47) Hu, Y.; Chen, L.; Jung, H.; Zeng, Y.; Lee, S.; Swamy, K. M.;
(12) Jang, Y. J.; Mulay, S. V.; Kim, Y.; Jorayev, P.; Churchill, D. G. Zhou, X.; Kim, M. H.; Yoon, J. ACS Appl. Mater. Interfaces 2016, 8,
New J. Chem. 2017, 41, 1653−1658. 22246−22252.
(13) Zhang, S. W.; Swager, T. M. J. Am. Chem. Soc. 2003, 125, 3420− (48) Wang, S. L.; Zhong, L.; Song, Q. H. Chem. Commun. 2017, 53,
3421. 1530−1533.
(14) Eubanks, L. E.; Dickerson, T. J.; Janda, K. D. Chem. Soc. Rev. (49) Wang, X. J.; Zhang, W. Q.; Cheng, K.; Yang, X.; Li, Q. Y.;
2007, 36, 458−470. Zhang, H.; Ma, Z.; Lu, H.; Wu, H. Org. Chem. Front. 2017, 4, 1719.
(15) Ashley, J. A.; Lin, C.; Wirsching, H. P.; Janda, K. D. Angew. (50) Xia, H. C.; Xu, X. H.; Song, Q. H. ACS Sensors. 2017, 2, 178−
Chem., Int. Ed. 1999, 38, 1793−1795. 182.
(16) Burnworth, M.; Rowan, S. J.; Weder, C. Chem. - Eur. J. 2007, 13, (51) Xia, H. C.; Xu, X. H.; Song, Q. H. Anal. Chem. 2017, 89, 4192−
7828−7836. 4197.
(17) Hewage, H. S.; Wallace, K. J.; Anslyn, E. V. Chem. Commun. (52) Zhang, Y.; Peng, A.; Jie, X.; Lv, Y.; Wang, X.; Tian, Z. ACS Appl.
2007, 3909−3911. Mater. Interfaces 2017, 9, 13920−13927.
(18) Kim, T. I.; Maity, S. B.; Bouffard, J.; Kim, Y. Anal. Chem. 2016, (53) Xie, H.; Wu, Y.; Zeng, F.; Chen, J.; Wu, S. Chem. Commun.
88, 9259−9263. 2017, 53, 9813−9816.
(19) Knapton, D.; Burnworth, M.; Rowan, S. J.; Weder, C. Angew. (54) Zhou, X.; Zeng, Y.; Chen, L.; Wu, X.; Yoon, J. Angew. Chem., Int.
Chem., Int. Ed. 2006, 45, 5825−5829. Ed. 2016, 55, 4729−4733.
(20) Chen, L.; Wu, D.; Yoon, J. ACS Sens. 2018, 3, 27−43. (55) Chen, X.; Pradhan, T.; Wang, F.; Kim, J. S.; Yoon, J. Chem. Rev.
(21) Lei, Z.; Yang, Y. J. Am. Chem. Soc. 2014, 136, 6594−6597. 2012, 112, 1910−1956.
(22) Diaz de Grenu, B.; Moreno, D.; Torroba, T.; Berg, A.; Gunnars, (56) Kim, H. N.; Lee, M. H.; Kim, H. J.; Kim, J. S.; Yoon, J. Chem.
J.; Nilsson, T.; Nyman, R.; Persson, M.; Pettersson, J.; Eklind, I.; Soc. Rev. 2008, 37, 1465−1472.
Wasterby, P. J. Am. Chem. Soc. 2014, 136, 4125−4128.
(23) Kumar, V.; Anslyn, E. V. J. Am. Chem. Soc. 2013, 135, 6338−
6344.
(24) Barba-Bon, A.; Costero, A. M.; Gil, S.; Martinez-Manez, R.;
Sancenon, F. Org. Biomol. Chem. 2014, 12, 8745−8751.
(25) Gotor, R.; Costero, A. M.; Gavina, P.; Gil, S. Dyes Pigm. 2014,
108, 76−83.
(26) Royo, S.; Martinez-Manez, R.; Sancenon, F.; Costero, A. M.;
Parra, M.; Gil, S. Chem. Commun. 2007, 4839−4847.
(27) Goswami, S.; Manna, A.; Paul, S. RSC Adv. 2014, 4, 21984−
21988.
(28) Sarkar, S.; Shunmugam, R. Chem. Commun. 2014, 50, 8511−
8513.
(29) Sarkar, S.; Mondal, A.; Tiwari, A. K.; Shunmugam, R. Chem.
Commun. 2012, 48, 4223−4225.
(30) Xuan, W.; Cao, Y.; Zhou, J.; Wang, W. Chem. Commun. 2013,
49, 10474−10476.
(31) Jang, Y. J.; Murale, D. P.; Churchill, D. G. Analyst 2014, 139,
1614−1617.
(32) Hiscock, J. R.; Piana, F.; Sambrook, M. R.; Wells, N. J.; Clark, A.
J.; Vincent, J. C.; Busschaert, N.; Brown, R. C.; Gale, P. A. Chem.
Commun. 2013, 49, 9119−9121.
(33) Goswami, S.; Das, S.; Aich, K. RSC Adv. 2015, 5, 28996−29001.
(34) Xuan, W.; Cao, Y.; Zhou, J.; Wang, W. Chem. Commun. 2013,
49, 10474−10476.
(35) Climent, E.; Biyikal, M.; Gawlitza, K.; Dropa, T.; Urban, M.;
Costero, A. M.; Martínez-Máñez, R.; Rurack, K. Sens. Actuators, B
2017, 246, 1056−1065.
(36) Bandyopadhyay, I.; Kim, M. J.; Lee, Y. S.; Churchill, D. G. J.
Phys. Chem. A 2006, 110, 3655−3661.
(37) Kim, Y.; Jang, Y. J.; Mulay, S. V.; Nguyen, T. T.; Churchill, D. G.
Chem. - Eur. J. 2017, 23, 7785−7790.

E DOI: 10.1021/acs.analchem.7b05011
Anal. Chem. XXXX, XXX, XXX−XXX

You might also like