You are on page 1of 13

Dyes and Pigments 147 (2017) 400e412

Contents lists available at ScienceDirect

Dyes and Pigments


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

Pyreneamide-based dipodal probes for ultra-sensitive and selective


detection of 3,5-dinitrosalicylic acid in an aqueous solution
Ashwani Kumar*, Pil Seok Chae**
Department of Bionanotechnology, Hanyang University, Ansan, 155-88, Republic of Korea

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

Article history: Pyrene-appended dipodal probes (probes 1 and 2) with differences in conformational rigidity were
Received 9 August 2017 synthesized for sensitive detection of 3,5-dinitrosalicylic acid (DNSA) in an aqueous solution. Both
Received in revised form dipodal probes had two pyrene units and exhibited quenching caused by aggregation (ACQ) at high water
23 August 2017
content (>60%) in DMSO. As the pyrene dimer (e.g., excimer) was a dominant conformation in 99%
Accepted 23 August 2017
aqueous solution, both probes showed sufficient excimer emission intensity at 486 nm (probe 1) or
Available online 25 August 2017
480 nm (probe 2) upon excitation. The excimer emission of the probes (probes 1 and 2) was sensitively
quenched in the presence of DNSA. Probe 1 selectively responded to DNSA among various carboxylic
Keywords:
Pyreneamide
acids (CAs), while probe 2 was responsive to both DNSA and 5-NSA. Density functional theory (DFT)
Excimer emission calculations suggest that, in the presence of DNSA, the probes undergo structural changes, leading to
DNSA formation of new pep interactions between each pyrene unit of the probe and the benzene ring of the
Fluorescence quenching guest molecule. This new pep interaction enabled energy transfer from pyrene to the non-radiative
guest molecule (DNSA), resulting in fluorescence quenching in the probes. The limit of detection
(LOD) of probe 1 with DNSA was observed to be 1 nM, the lowest value reported so far. This ultra-
sensitivity toward DNSA detection was retained even in the case of a probe 1-coated TLC strip. Probe
3, which had only one pyrene unit, gave produced a monomeric emission spectrum with lmax ¼ 406 nm,
along with aggregation-induced emission enhancement (AIEE) behavior. The interactions of DNSA with
individual probes were examined by UV-visible, fluorescence, and 1H NMR spectroscopies and were
further supported by DFT calculations.
© 2017 Published by Elsevier Ltd.

1. Introduction biological samples.


The medicinal properties of salicylic acid (SA), particularly for
Aliphatic and aromatic carboxylic acids play significant roles in a fever reduction, have been known since ancient times. SA also has
large number of metabolism/biological processes [1] and are anti-inflammatory and analgesic effects. These days, SA and its
extensively used in various fields and industries. For example, derivatives are used as constituents of skin-care products and in
aliphatic carboxylic acids such as citric, gluconic, lactic, malic, and some drugs. For instance, methyl salicylate is a liniment to reduce
tartaric acids are used as additives in the food industry [2]. Aro- muscle and joint pain, while choline salicylate is used topically to
matic carboxylic acids are the main ingredients of paints and wood relieve the pain of mouth ulcers and to treat several diseases such
preservatives. Perfluorocarboxylic acids are used as components of as psoriasis, calluses, corns, acne, seborrheic dermatitis, acanthosis
fire-retardant foams and fluoroelastomers, in addition to their nigricans, ichthyosis, and warts [6]. SA is an ingredient in shampoos
widespread applications in process control [3], medical diagnosis because it is used for dandruff treatment [7]. Derivatives of SA are
[4], and environmental monitoring [5]. Thus, there is growing in- widely used in medical diagnosis, pharmaceuticals, cosmetics, food
terest in sensing these acidic molecules in environmental and technology, and environmental monitoring [8]. SA also plays
important roles in plants including defense regulation against in-
sects and pathogens and in various physiological processes such as
seed germination, fruit yield, protein synthesis, and nutrient uptake
* Corresponding author.
** Corresponding author.
[9]. SA can also cause one of the most prevalent forms of immu-
E-mail addresses: ashwanirubal@gmail.com (A. Kumar), pchae@hanyang.ac.kr notoxicity found in humans (i.e., allergic contact dermatitis) [10].
(P. Seok Chae).

http://dx.doi.org/10.1016/j.dyepig.2017.08.039
0143-7208/© 2017 Published by Elsevier Ltd.
A. Kumar, P. Seok Chae / Dyes and Pigments 147 (2017) 400e412 401

One SA derivative, 3,5-dinitrosalicylic acid (DNSA), is used in a crude solid was dissolved in H2O and extracted with CH2Cl2, and
colorimetric biomedical assay for the detection of reducing sugars the organic layer was washed with H2O (2  50 mL) and a 5%
[11,12]. Consequently, selective monitoring and quantitative aqueous NaCl solution (100 mL). The collected organic layer was
detection of SA and its derivatives in aqueous solutions are neces- then dried over anhydrous sodium sulfate and filtered. The filtrate
sary for detailed studies of their effects, functions, and physiological was evaporated under reduced pressure, and the residue was pu-
properties [13]. rified by column chromatography using CH2Cl2 as the eluent to
Among various detection methods, fluorescence-based tech- afford compound 4 as a white solid (2.3 g, 90% yield). 1H NMR
niques have many advantages over others. The high sensitivity, (400 MHz, DMSO-d6): d 1.40 (s, 18H, 6  CH3), 3.69 (d, J ¼ 4.8 Hz, 4H,
specificity, and real-time monitoring with rapid response and easy 2  CH2), 3.83 (s, 2H, CH2), 7.01 (t, J ¼ 4.8 Hz, 2H, 2  NHCOOBut),
portability make this method irreplaceable and enable in-situ 7.13 (d, J ¼ 6.8 Hz, 4H, 4  ArH), 7.48 (d, J ¼ 6.8 Hz, 4H, 4  ArH), 9.85
monitoring of analytes of interest in real time and space. (s, 2H, 2  NHCO).
In the design of fluorescence-based chemical probes, versatile
fluorophores have been connected with distinct binding units such 2.1.1.2. 2,2'-((methylenebis(4,1-phenylene))bis(azanediyl))bis(2-
as a-aminopyridines/a-amidopyridines, ureas, amides/sulfon- oxoethan-1-aminium) (compound 5). To a CH2Cl2 (30 mL) solution
amides, and quaternary ammonium/imidazolium salts to develop of compound 4 (512 mg, 1 mmol), trifluoroacetic acid (TFA, 4 mmol)
carboxylic acid sensors [13e17]. However, detection systems for a was added drop-wise, and the mixture was stirred overnight at
CA in aqueous solutions are rare because most sensors have been room temperature under an argon atmosphere. After completion of
effective in non-polar or polar organic solvents. Recently, pyrene-1- the reaction (analyzed by TLC), the reaction mixture was evapo-
sulfonamide-based probes having neutral or ionic character have rated under reduced pressure and precipitated by the addition of
been used for the recognition of SA and its derivatives in non-polar diethyl ether. The suspension was filtered, and the solid was
or polar to mixed solvent systems [13]. In continuation of our washed with diethyl ether to afford compound 5 as a white solid
previous study of optical molecular probes [18], the aim of this (300 mg, 95% yield). 1H NMR (400 MHz, DMSO-d6): d 3.76 (br s, 4H,
study is to develop chemical probes with enhanced selectivity to- 2  CH2), 3.86 (s, 2H, CH2), 7.19 (t, J ¼ 8.4 Hz, 4H, 4  ArH), 7.49 (d,
ward DNSA and to approach a pure aqueous system, which is an J ¼ 8.4 Hz, 4H, 4  ArH), 8.16 (br s, 6H, 2 x 2NHþ 3 ), 10.45 (s, 2H, 2 
essential feature for practical applications. NHCO); 13C NMR (100 MHz, DMSO-d6): d 40.94, 119.33, 129.13,
In the current study, we synthesized pyreneamide-based dipo- 136.22, 136.87, 164.61.
dal probes 1 and 2 for sensitive and selective detection of an SA
derivative in pure aqueous solution. Probes 1 and 2 commonly bear 2.1.1.3. N,N'-(((methylenebis(4,1-phenylene))bis(azanediyl))bis(2-
a central 4,40 -methylenediphenylene (MDP) unit flanked by two oxoethane-2,1-diyl))bis(pyrene-1-carboxamide) (probe 1). A
glycine residues attached to a pyrene fluorophore. Probe 1 dis- mixture of pyrene-1-carboxylic acid (270 mg, 1.1 mmol), hydrox-
played high selectivity and sensitivity in DNSA sensing, with a limit ybenzotriazole (HOBt, 162 mg, 1.2 mmol), and 1-ethyl-3-(3-
of detection (LOD) of 1 nM, while probe 2 was responsive to both dimethylaminopropyl)carbodiimide (EDC, 230 mg, 1.2 mmol) was
DNSA and 5-nitrosalicylic acid (5-NSA) with reduced sensitivity. To dissolved in (CH2Cl2: DMF (1:1), 20 mL) solution and stirred
the best of our knowledge, this is the lowest LOD (1 nM) observed vigorously at 0  C for 20 min under an argon atmosphere. Com-
for DNSA detection. The ultra-sensitive detection observed for pound 5 (157 mg, 0.5 mmol) and diisopropylethylamine (DIPEA)
probe 1 was attributed to strong intermolecular p-p interactions (155 mg, 1.2 mmol) were added to the reaction mixture and sub-
between pyrene units of the probe and DNSA, allowing energy sequently stirred at room temperature for another 8 h. After
transfer from the probe to the guest molecule. The conformational completion of the reaction (as detected via TLC analysis), the sol-
rigidity of probe 1 appeared to be associated with the high selec- vent of the reaction mixture was removed under vacuum. The
tivity of the probe toward DNSA over 5-NSA. We discuss in- crude solid was dissolved in H2O and extracted with CH2Cl2, and
teractions of these dipodal probes and their monopod analog the organic layer was washed with H2O (2  50 mL) and a 5%
(probe 3) with DNSA in terms of aggregate size and absorption, aqueous NaCl solution (100 mL). The collected organic extract was
fluorescence and nuclear magnetic resonance (NMR) spectros- then dried over anhydrous sodium sulfate and filtered. The filtrate
copies in addition to DFT calculations frontier molecular orbitals was evaporated under reduced pressure, and the residue was pu-
(FMOs) of the probes and their complexes with the guest molecule. rified by column chromatography using CH2Cl2:CH3OH (8:2) as the
eluent to afford probe 1 as a white solid (325 mg, 85% yield);
2. Experimental section m.p. ¼ 302  C. 1H NMR (400 MHz, DMSO-d6): d 3.91 (s, 2H, CH2),
4.23 (d, J ¼ 6.0 Hz, 4H, 2  CH2), 7.22 (d, J ¼ 8.4 Hz, 4H, 4  ArH),
2.1. Details for general procedure and theoretical calculations were 7.61 (d, J ¼ 8.4 Hz, 4H, 4  ArH), 8.14 (t, J ¼ 8.0 Hz, 2H, 2  ArH),
given in ESI 8.21e8.31 (m, 8H, 8  ArH), 8.36e8.38 (m, 6H, 6  ArH), 8.71 (d,
J ¼ 9.2 Hz, 2H, 2  ArH), 9.03 (t, J ¼ 6.0 Hz, 2H, 2  PyCONH), 10.17
2.1.1. Detailed procedures for probe preparation (synthesis of probes (s, 2H, 2  ArNH); 13C NMR (100 MHz, DMSO-d6): d 43.26, 119.30,
1e3) 123.62, 123.76, 124.42, 124.94, 125.34, 125.62, 125.79, 126.58,
2.1.1.1. Di-tert-butyl (((methylenebis(4,1-phenylene))bis(azanediyl)) 127.22, 127.92, 128.09, 128.33, 128.98, 130.23, 130.71, 131.59, 131.67,
bis(2-oxoethane-2,1-diyl))dicarbamate (compound 4). A mixture of 136.31, 137.03, 167.61, 169.41; HRMS (EI): calc'd. for C51H37N4O4
(tert-butoxycarbonyl)glycine (1.93 g, 11 mmol), hydroxybenzo- (M þ Hþ): 769.2809; found: m/z 769.2812.
triazole (HOBt, 1.62 g, 12 mmol), and 1-ethyl-3-(3-
dimethylaminopropyl)carbodiimide (EDC, 2.30 g, 12 mmol) was 2.1.1.4. N,N'-(((methylenebis(4,1-phenylene))bis(azanediyl))bis(2-
dissolved in dichloromethane-dimethylformamide (CH2Cl2: DMF oxoethane-2,1-diyl))bis(4-(pyren-1-yl)butanamide) (probe 2).
(1:1), 20 mL) and stirred vigorously at 0  C for 20 min under an The synthetic procedure for this preparation is similar to that used
argon atmosphere. Then, 4,40 -methylenedianiline (0.99 g, 5 mmol) for probe 1 using pyrene-1-butyric acid (317 mg, 1.1 mmol) instead
and diisopropylethylamine (DIPEA) (1.55 g, 12 mmol) were added of pyrene-1-carboxylic acid. Probe 2 was obtained as a white solid
to the reaction mixture and stirred at room temperature for another (400 mg, 93% yield); m.p. ¼ 246  C. 1H NMR (400 MHz, DMSO-d6):
2 h. After completion of the reaction (detected by TLC analysis), the d 1.99e2.06 (m, 4H, 2  CH2), 2.34 (t, J ¼ 7.2 Hz, 4H, 2  CH2), one
solvent of the reaction mixture was reduced under vacuum. The signal of two CH2 groups merged with the signal of H2O, 3.82 (s, 2H,
402 A. Kumar, P. Seok Chae / Dyes and Pigments 147 (2017) 400e412

CH2), 3.88 (d, J ¼ 6.0 Hz, 4H, 2  CH2), 7.12 (d, J ¼ 8.4 Hz, 4H, 4  DMSO-d6 exhibited two aliphatic peaks: one from CH2 groups
ArH), 7.49 (d, J ¼ 8.4 Hz, 4H, 4  ArH), 7.96 (d, J ¼ 8.0 Hz, 2H, 2  (singlet at d 3.91 ppm) and the other from two NCH2 groups of the
ArH), 8.05 (t, J ¼ 7.6 Hz, 2H, 2  ArH), 8.10e8.26 (m, 14H, 12  glycine units (doublet at d 4.23 ppm). The amide protons of the two
ArH þ 2  PyCONH), 8.41 (d, J ¼ 9.2 Hz, 2H, 2  ArH), 9.94 (s, 2H, 2  PyCONH and two ArNHCO groups appeared at d 9.03 and
ArNH); 13C NMR (100 MHz, DMSO-d6): d 27.61, 32.20, 34.85, 42.67, d 10.17 ppm, respectively, which are far downfield. This is pre-
119.25, 123.62, 124.16, 124.24, 124.78, 124.91, 124.96, 126.14, 126.50, sumably due to intramolecular hydrogen bond formation. In addi-
127.20, 127.48, 127.57, 128.18, 128.89, 129.30, 130.44, 130.90, 136.25, tion to the 1H NMR spectrum, the 13C NMR spectrum and mass data
136.68, 136.93, 167.78, 172.58; HRMS (EI): calc'd. for C57H49N4O4 are consistent with the chemical structure of probe 1 (see ESI).
(M þ Hþ): 853.3748; found: m/z 853.3751. Similarly, the 1H NMR spectrum of probe 2 in DMSO-d6 showed
three aliphatic signals arising from the propylene chain. One signal
2.1.1.5. Tert-butyl (2-((4-methoxyphenyl)amino)-2-oxoethyl)carba- appeared as a multiplet at d 1.99e2.06 ppm, corresponding to the
mate (compound 6). The synthetic procedure for this preparation is central CH2 group of the chain. The signal for one CH2 group of the
similar to that used for compound 4 using p-anisidine (370 mg, chain close to the carbonyl group appeared as a triplet at
3 mmol) instead of 4,40 -methylenedianiline. The procedure affor- d 2.34 ppm, while the signal for the other CH2 group close to pyrene
ded compound 6 as a thick liquid (800 mg, 95% yield). 1H NMR was obscured by overlapping with a H2O signal. The signals for the
(400 MHz, CDCl3): d 1.47 (s, 9H, 3  CH3), 3.79 (s, 3H, OCH3), 3.92 (d, amide protons close to pyrene merged with pyrene signals, giving a
J ¼ 6.0 Hz, 2H, CH2), 5.28 (br s, 1H, NHCOOBut), 6.86 (d, J ¼ 10.4 Hz, multiplet peak at d 8.10e8.26 ppm, and the peak for the amide
2H, 2  ArH), 7.40 (d, J ¼ 10.0 Hz, 2H, 2  ArH), 8.04 (br s, 1H, protons of two ArNHCO appeared as a singlet in a far downfield
NHCO); 13C NMR (100 MHz, CDCl3): d 28.51, 55.68, 77.55, 144.38, position (d 9.94 ppm). The 13C NMR spectrum and mass data also
122.04, 130.67, 156.79, 167.77. supported the successful formation and isolation of probe 2.
Similarly, the synthesis of probe 3 was also confirmed by 1H and 13C
2.1.1.6. 2-((4-methoxyphenyl)amino)-2-oxoethan-1-aminium (com- NMR spectra and mass data.
pound 7). The synthetic procedure for this preparation is similar to The UV-visible absorption spectra of the probes (probes 1, 2, and
that used for compound 5. The procedure afforded compound 7 as a 3) dissolved in DMSO exhibited a typical pyrene signature with
brownish-white solid (500 mg, 91% yield). 1H NMR (400 MHz, absorption maxima (lmax) at 330, 345, and 372 nm. Due to the
DMSO-d6): d 3.73 (s, 3H, OCH3), 3.74 (d, J ¼ 5.6 Hz, 2H, CH2), 6.94 (d, presence of two pyrene units, absorbances of probes 1 and 2 were
J ¼ 6.8 Hz, 2H, 2  ArH), 7.50 (d, J ¼ 6.8 Hz, 2H, 2  ArH), 8.10 (br s, found to be roughly two-fold higher than that of probe 3 (Fig. S1).
3H, NHþ 3 ), 10.31 (s, 1H, NHCO). The optimal excitation wavelength for maximum fluorescence in-
tensity was obtained using a range of wavelengths from 300 to
2.1.1.7. N-(2-((4-methoxyphenyl)amino)-2-oxoethyl)pyrene-1- 360 nm with a 5 nm interval. Probes 1 and 2 dissolved in aqueous
carboxamide (probe 3). The synthetic procedure for this prepara- solution (99%) produced the maximum fluorescence intensity upon
tion is similar to that used for probe 1 using compound 7 (181 mg, excitation at 340 nm (Fig. S2-S3). Under these conditions, the
1.0 mmol) instead of compound 5. The procedure afforded probe 3 dipodal probes gave rise to green fluorescence with lmax at
as a white solid (360 mg, 88% yield); m.p. ¼ 254  C. 1H NMR 480e488 nm, indicative of the formation of a pyrene excimer via
(400 MHz, DMSO-d6): d 3.75 (s, 3H, CH3), 4.21 (d, J ¼ 6.0 Hz, 2H, the intramolecular p-p stacking of two pyrene units. An emission
CH2), 6.94 (d, J ¼ 9.6 Hz, 2H, 2  ArH), 7.60 (d, J ¼ 9.6 Hz, 2H, 2  maximum at 383 nm associated with monomeric pyrene units was
ArH), 8.14 (t, J ¼ 8.0 Hz, 1H, ArH), 8.22e8.31 (m, 4H, 4  ArH), 8.37 also observed. To further investigate the solvent effect on the
(t, J ¼ 8.0 Hz, 3H, 3  ArH), 8.71 (d, J ¼ 9.6 Hz, 1H, ArH), 9.03 (t, fluorescence emission of these probes, 1 mM probe solutions were
J ¼ 6.0 Hz, 1H, PyCONH), 10.07 (s, 2H, ArNH); 13C NMR (100 MHz, prepared in various solvent systems, and their fluorescence spectra
DMSO-d6): d 43.18, 55.17, 113.94, 120.65, 123.63, 123.77, 124.44, were measured. The solvents used for this experiment include
124.97, 125.36, 125.64, 125.81, 126.60, 127.24, 127.93, 128.09, 128.35, water (99%), dimethylsulfoxide (DMSO), dimethylformamide
130.25, 130.73, 131.62, 131.67, 132.21, 155.19, 167.28, 169.41; HRMS (DMF), ethanol (EtOH), methanol (MeOH), t-butanol (tBuOH),
(EI): calc'd. for C26H21N2O3 (M þ Hþ): 409.1547; found: m/z acetonitrile (ACN), acetone, tetrahydrofuran (THF), ethyl acetate
409.1550. (EA), dichloromethane (DCM), chloroform (CHCl3), dioxane, and
toluene. Under an illumination of 365 nm, probe 1 showed a high
3. Results and discussion blue emission in most solvent cases. This emission was from the
pyrene monomer. In contrast, this monomeric fluorescence in-
The dipodal probes (1 and 2) share a rigid MDP unit with a bent tensity sharply decreased when water was used as a solvent. This
structure in the central region, making it possible to form a pyrene was accompanied by the emergence of a green excimer emission
excimer in an appropriate environment (see Fig. 1). However, these with relatively low intensity (Fig. 2). As for probe 2, both excimer
two probes differ from each other in terms of conformational ri- and monomer emissions (designated as I480 and I380, respectively)
gidity because different pyrene acid derivatives were used for appeared in all tested solvent systems, but the relative intensities
probe preparation. Probe 2 is less rigid than probe 1 because the (I480/I380) varied significantly depending on the solvent used for the
pyrene-1-butyric acid used for probe 2 preparation has an addi- study (Fig. S4). Of the various solvent systems, 99% water gave the
tional flexible propylene chain compared to pyrene-1-carboxylic highest value in this emission ratio (I480/I380), and this was
acid used for probe 1 preparation. A short glycine linker and two responsible for the green emission of the probe solution under
rigid amide bonds along with the rigid MDP unit strongly restrict illumination at 365 nm (Fig. S4). This behavior was similar to that of
the conformational flexibility of probe 1. We also prepared probe 3 probe 1, although the change in the emission ratio was less dra-
with a single pyrene unit, which is a monopod analog of probe 1 matic than probe 1 (Fig. S5). This result implies that the solvent
(see Fig. 1). Probes 1 and 2 were readily synthesized by EDC- plays a crucial role in the formation of intramolecular p-p stacking
supported coupling of 2,2'-((methylenebis(4,1-phenylene))bis(a- of the two pyrene rings present in these dipodal probes. Next, we
zanediyl))bis(2-oxoethan-1-aminium) (compound 5) with respec- measured fluorescence spectra with increasing probe concentra-
tive acid components (Scheme 1). The structures of probes 1 and 2 tion in DMSO. Probe 1 produced pyrene monomer emission peaks
were supported by 1H and 13C NMR spectroscopies as well as by FAB with little contribution from the pyrene excimer emission (Fig. S6).
mass spectrometry (see ESI). The 1H NMR spectrum of probe 1 in Probe 1 has a low tendency to form intramolecular p-p interactions
A. Kumar, P. Seok Chae / Dyes and Pigments 147 (2017) 400e412 403

Fig. 1. Chemical structures of pyreneamide-based probes (probe 1, 2, and 3). In dipodal probes 1 and 2, two terminal pyrene fluorophores are connected to a central 4,40 -meth-
ylenediphenylene (MDP) unit through a glycine linker, while probe 3 is the monopod analog of probe 1. Pyrene-1-carboxylic acid and pyrene-1-butyric acid were used for the
preparation of probes 1 and 2, respectively, providing a substantial difference in conformational rigidity between these dipodal probes.

Scheme 1. Synthetic scheme for the preparation of pyreneamide-based molecular probes (probes 1, 2, and 3). (a) HOBt, EDC, DIPEA, 0  C / RT; (b) TFA, RT; (c) pyrene-1-carboxylic
acid, HOBt, EDC, DIPEA, 0  C / RT; (d) pyrene-1-butyric acid, HOBt, EDC, DIPEA, 0  C / RT.

in this solvent system, presumably due to its rigid structure. In probes 1e3 became abruptly broader, the absorbance significantly
contrast, probe 2, which has a rather flexible architecture, exhibited decreased, and a level-off tail was observed in a wavelength region
both pyrene monomer and excimer emissions under the same higher than 360 nm for each probe (Fig. S1). This result suggests the
conditions, indicating that the conformational flexibility of this formation of large probe aggregates at high concentrations of water
probe enables pyrene-pyrene interactions. As expected, probe 3, in DMSO.
which has a single pyrene unit, yielded only peaks assigned to Fluorescence properties of probes 1e3 also significantly varied
monomeric pyrene emission (Fig. S6c). These peak intensities with increasing water content in a binary (DMSO-water) solvent
tended to increase with increasing probe concentration in all probe system (Fig. 3). The fluorescence intensity (I384) of probe 1 pro-
cases, but reached individual maxima at different probe concen- duced by monomeric pyrene remained almost constant up to 40%
trations, 4 to 10e12 mM, for probes 1e3, respectively. For probe 2, water content, but decreased drastically with a further increase in
the increase in excimer emission intensity (I480) was more promi- water content from 60 to 100% (Fig. 3A), which suggests aggrega-
nent than that of the monomer counterpart (I380) (Fig. S6d). tion caused by quenching (ACQ) behavior. Due to this quenching
We investigated the solvent composition effect on the photo- behavior, an increase in excimer emission intensity (I484) of the
physical properties of probes 1e3 by increasing the concentration probe with increasing water content was not noticeable in the
of water from 0 to 99% in DMSO and performing UV-visible (Fig. S1) spectra (Fig. 3B), but could be observed using the emission intensity
and fluorescence spectroscopy (Fig. 3 & Fig. S7-8) measurements. ratio (I484/I384) (Fig. 3C). This result indicates the enhanced ten-
We also monitored the morphology and size changes for probe dency to form intramolecular pyrene-pyrene stacking under the
aggregates under the same conditions (Fig. 3). When water content condition tested (i.e., high water content in DMSO). To further
was increased more than 60%, the UV-visible absorption spectra of support the ACQ behavior of probe 1, we analyzed probe aggregates
404 A. Kumar, P. Seok Chae / Dyes and Pigments 147 (2017) 400e412

Fig. 2. (A) Solution color image of probe 1 dissolved in an indicated solvent. The image was obtained under UV light illumination at 365 nm, (B) corresponding fluorescence
emission spectra, and (C) normalized fluorescence emission spectra of the probe (1 mM) with lex ¼ 340 nm. In the normalized data, the fluorescence spectrum obtained from the
probe in water is quite different from those in other solvents, as expected from the corresponding solution colors in (A). (For interpretation of the references to colour in this figure
legend, the reader is referred to the web version of this article.)

via FE-SEM at different water contents (0, 50 and 90%) (Fig. 3D). At this FE-SEM study indicates that aggregates formed by probe 3
0% water content (DW0), the image showed very few small parti- mainly underwent morphological and size changes from nano-
cles with a diameter of about 400 nm. At 50% water content sized pillars to micro-sized cloud-shaped aggregates. This obser-
(DW50), the particle size increased to 3e5 mm. When 90% water vation is consistent with the AIEE behavior of this probe.
content was used, large round plate-shaped aggregates with a In summary, probes 1 and 2, with the dipodal pyrene units,
diameter of 30 mm were observed in a free or merged state. Thus, showed an increased tendency to form intramolecular pyrene-
this FE-SEM study indicates that aggregates formed by probe 1 pyrene stacking with increasing water concentration. They also
mainly underwent a morphological change from nano-sized par- showed ACQ phenomena in the high water concentration range of
ticles to micro-sized plate-shaped aggregates when water content 60e100%. In contrast, probe 3, having a single pyrene unit, dis-
was increased from 0 to 50 to 90%. This behavior is likely associated played AIEE behavior with increasing water concentration without
with a decrease in the monomer emission intensity of this probe at excimer emission. These results are consistent with the previously
a high water concentration (i.e., ACQ phenomenon). Unlike probe 1, reported pyrene sulfonamide-based probes having one or two
the monomer emission intensity (I380) of probe 2 decreased sub- pyrene units [18].
stantially with increasing water content from 0 to 40%, along with a The interesting photo-physical properties of probe 1 prompted
concomitant increase in excimer emission (I480) (Fig. S7), indicating us to investigate this probe for DNSA detection in an aqueous so-
that probe 2 has a greater tendency to form intramolecular pyrene- lution. For this purpose, a variety of CAs (each 50 mM) was indi-
pyrene interactions than does probe 1, probably due to its increased vidually added to solutions of probe 1 (1 mM). CAs used for this
flexibility. When water content was further increased to 60e100%, study includes both aliphatic and aromatic acids (e.g., fumaric acid
probe 2 showed a sharp decrease in both the monomer and excimer (Fum. A), indole-2-carboxylic acid (IND-2-A), maleic acid (Maleic
emission intensities (Figs. S7B,C), which was also observed for A), malonic acid (Malonic A), nicotinic acid (NA), picolinic acid (PA),
probe 1. This result is likely traced to the ACQ effect for both probes. terephthalic acid (T-PhA), isophthalic acid (Iso-PhA), phthalic acid
In this context, probe 3 showed a behavior quite different from (PhA), succinic acid (Succinic A), uric acid (UA), 3,5-dinitrobenzoic
probes 1 and 2. This monopod probe produced a continuous in- acid (3,5-DNBA), benzoic acid (BA), 2-thiophene carboxylic acid
crease in emission intensity with increasing water content up to (2-ThiopheneA), 3-chlorobenzoic acid (3-ClBA), 3-iodobenzoic acid
80%, and the emission maxima was a little changed from 383 nm to (3-IBA), 3-nitrobenzoic acid (3-NBA), 4-nitrobenzoic acid (4-NBA),
406 nm (Fig. S8). This result is a typical feature for a probe showing SA, 2,5-dihydroxybenzoic acid (2,5-DiOH-BA), 5-bromosalicylic
aggregation-induced emission enhancement (AIEE). We also acid (5-BrSA), 5-chlorosalicylic acid (5-ClSA), 5-methylsalicylic
analyzed probe 3 aggregates via FE-SEM at 0, 50, and 90% water acid (5-MSA), 5-iodosalicylic acid (5-ISA), 5-nitrosalicylic acid (5-
content in DMSO (designated as DW0, DW50, and DW90, respec- NSA), and 3,5-dinitrosalicylic acid (DNSA)). Among these CAs, the
tively) (Fig. S8C). At 0% water content (DW0), the image showed a fluorescence emission of probe 1 was selectively quenched by the
number of standing random pillars about 100 nm each in size; presence of DNSA, presumably due to its highly electron-deficient
while, at 50% water contents (DW50), these pillars moved closely nature (Fig. 4 and Fig. S9). The sample containing 5-NSA also
together to form large aggregates. Finally, at 90% water content, the showed fluorescence quenching, but with a much smaller degree
FE-SEM image showed large aggregates 20e35 mm in size. Thus, than DNSA. We also carried out an interference study by adding
A. Kumar, P. Seok Chae / Dyes and Pigments 147 (2017) 400e412 405

Fig. 3. (A) Changes in solution color of probe 1 (1 mM) with increasing water concentration in a mixed DMSO:water solvent system under 365 nm illumination, (B) corresponding
spectral changes in fluorescence emission (lex ¼ 340 nm), (C) bar graph of the fluorescence intensity ratio (I488/I388) of the probe, and (D) FE-SEM images for probe aggregates
prepared at 0, 50, and 90% water content in DMSO. D and W represent DMSO and water, respectively, and the numbers 0, 10, 20, 40, …, 100 indicate the percentage of water. (For
interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

10 mM of DNSA to a probe 1 solution including individual CAs at inflection point at 0.7 mol fraction of DNSA (Fig. S11). The limit of
50 mM. No interferences of these CAs were observed in DNSA detection (LOD) of probe 1 for DNSA was observed to be as low as
detection (Fig. S10). 1 nM. This value was obtained from the linear relationship between
A fluorescence titration of probe 1 with DNSA was carried out to the fluorescence intensity of probe 1 at 484 nm (I484) and [DNSA]
understand their binding characteristics. Each additional DNSA (inset of Fig. 4C). A similar behavior was observed for probe 2 with
aliquot gradually quenched the emission of probe 1 at 484 nm, with the same guest molecule, but the fluorescence quenching effect of
complete fluorescence quenching observed upon the addition of 2 5-NSA was also pronounced in this case (Fig. S12), indicating the
equivalents of DNSA (Fig. 4BeC). The fitting of this titration data limited selectivity of this probe in DNSA detection. Probe 2 showed
using a 1:2 binding model gave Ka ¼ 2.92  107 M1 with a small the same binding stoichiometry (1:2) as probe 1 and bound
error (±5%), which indicates a strong association between the strongly to DNSA, although the binding strength of this probe was
probe and DNSA (Table 1). A binding ratio of 1:2 was deduced from approximately two times weaker than that of probe 1
the titration data and was also confirmed by a Job's plot showing an (Ka ¼ 2.92  107 vs. 1.57  107). This probe was calculated to bind to
406 A. Kumar, P. Seok Chae / Dyes and Pigments 147 (2017) 400e412

Fig. 4. (a) Fluorescence quenching effect of various carboxylic acids (CAs) for probe 1 in water. The probe and individual CAs were used at 1 and 50 mM, respectively. Io and I
represent fluorescence intensities of probe 1 at 484 nm in the absence and presence of individual CAs, respectively. The inset shows the color change of probe 1 solution before and
after DNSA addition under illumination at 365 nm. (B) Fluorescence spectral change of probe 1 (1 mM, water) with increasing amount of DNSA ranging from 0 to 4 mM, and (C) a
graphical representation of the change in fluorescence intensity at 484 nm (I484) as a function of DNSA concentration, lex ¼ 340 nm. The inset shows a linear relationship between
I484 and [DNSA]. (D) FE-SEM images for aggregates formed by free probe 1, its DNSA complex, and DNSA alone in a 90% water solution in DMSO. (For interpretation of the references
to colour in this figure legend, the reader is referred to the web version of this article.)

DNSA 100 times stronger than to 5-NSA (Ka ¼ 1.57  107 vs. in detecting DNSA and 5-NSA. The association strengths of the
1.25  105) (Table 1). A control agent, probe 3, displayed a sensing probe with these acid derivatives were only moderately strong or
behavior similar to that of probe 2 in terms of the limited selectivity relatively weak (Fig. 5 and Fig. S13). It is noteworthy that we used
A. Kumar, P. Seok Chae / Dyes and Pigments 147 (2017) 400e412 407

Table 1 emission intensity at 406 nm instead of 484 nm to calculate the


Association constants (Ka (M1)) of probes 1e3 with DNSA or 5-NSA in water (99%). association constant for the [probe 3 þ DNSA or 5-NSA] system as
Probes 1 2 3 this probe did not show excimer emission.
DNSA 2.92  107 1.57  107 2.42  104
To investigate whether the complexation of probe 1 with DNSA
5-NSA e 1.25  105 9.83  103 affects the aggregate size, we performed FE-SEM analysis on [probe
1 þ DNSA (4 equiv.)] suspended in 99% water solution (Fig. 4D). FE-

Fig. 5. (a) Fluorescence quenching effects of various carboxylic acids (CAs) for probe 3 dissolved in water. The probe and individual CAs were used at 1 mM and 50 mM, respectively. Io
and I represent fluorescence intensities of probe 1 at 406 nm in the absence and presence of individual CAs, respectively. (B) Fluorescence spectral change of probe 3 (1 mM, water)
with increasing amount of DNSA, and (C) a graphical representation of the change in fluorescence intensity at 406 nm (I406) as a function of DNSA concentration, lex ¼ 340 nm. (D)
FE-SEM images for aggregates formed by free probe 3, its DNSA complex, and DNSA alone in a 90% water solution in DMSO.
408 A. Kumar, P. Seok Chae / Dyes and Pigments 147 (2017) 400e412

SEM images showed a decrease in aggregate size from 30 mm to (6:2) systems.


5 mm upon complexation of the probe with DNSA. The decrease in Further insight into the conformational changes of probes 1e3
aggregate size is likely due to the increased polarity of the probe 1- upon complexation with DNSA was obtained by calculating energy-
DNSA complex relative to that of probe 1 alone. This increase in minimized structures via density functional theory (DFT) at the
polarity was caused by the association with highly water-soluble B3LYP/6-31G* level. As can be seen in Fig. 8, the optimized struc-
DNSA. Such a large polarity increase can induce partial dissocia- tures for probe 1 or 2 alone showed intramolecular pyrene-pyrene
tion of aggregates in an aqueous solution. A similar de-aggregation interactions at a distance of 4.15e4.22 Å, which is comparable to a
behavior was also observed for probe 3 complexed with DNSA previous report [19]. In the presence of DNSA, probes 1 and 2 were
(Fig. 5D). calculated to undergo structural changes from a closed to an open
The fluorescence quenching effect of DNSA for each probe conformation. Consequently, the pyrene-pyrene interactions of the
(probe 1, 2, or 3) was further investigated by plotting the fluores- probes are converted into newly formed pyrene-DNSA interactions
cence intensity of the probe as a function of DNSA concentration at a distance of 3.53e3.68 Å (Fig. 8). These new pep interactions
(Stern-Volmer plot) (Fig. 6A). Stern-Volmer constants were calcu- enable energy transfer from the electron-rich pyrene ring to an
lated using the Stern-Volmer equation (Io/I ¼ 1 þ Ksv [Q]). Stern- electron-deficient DNSA upon excitation at 340 nm, leading to
Volmer constants (Ksv (M1)) of probes 1 and 2 with DNSA were quenching of fluorescence emission. In Fig. 5, we show that the
found to be higher (1.19  107 M1 and 7.73  106, respectively) fluorescence emission of probe 3 at 406 nm was also quenched in
than that of probe 3 (2.11  105 M1) (Fig. 6B), indicating a more the presence of DNSA. This result could also be explained by
effective decrease in the fluorescence intensity of these dipodal looking at the optimized structures of the probe-DNSA complex
probes in the presence of DNSA than probe 3. The relative order of obtained from the DFT calculation. Like the dipodal probes, the
Stern-Volmer constants (Ksv) for individual probe-DNSA systems pyrene ring of this probe forms pep interactions with DNSA at a
was similar to that of the association constants (Ka) for the same distance of 3.58 Å in the presence of the guest molecule (Fig. 8).
systems (Table 1). The high Stern-Volmer constants (Ksv) and as- The fluorescence emission of the probes and the emission
sociation constants (Ka) observed for the dipodal probes (1 and 2) quenching by their association with DNSA were further analyzed in
strongly suggest that pyrene excimer formation is a main reason for terms of electronic transitions from their HOMOs to LUMOs
the enhanced ability of these probes to sensitively detect DNSA (Fig. S17). DFT calculations showed that the frontier molecular or-
relative to probe 3. bitals (i.e., HOMOs and LUMOs) of these probes are mainly localized
To further address the interactions between the pyrene rings on the pyrene rings of the probes. Upon complexation with DNSA,
and DNSA, 1H NMR spectra of probes 1 and 2 were collected in their locations were changed as follows. For probe 1 or 2-DNSA
DMSO-d6 in the absence or presence of DNSA. The addition of 1 or 2 complexes, the HOMOs and LUMOs were localized on the pyrene
equiv. of DNSA to each probe did not result in any substantial ring and DNSA, respectively, while for the probe 3-DNSA complex,
changes in chemical shifts for the protons of pyrene and benzene these orbitals were localized on p-anisidine and DNSA, respectively.
rings or glycine units in DMSO-d6 (Fig. S14-S15). No detectable Thus, it is likely that energy transfer takes place from the pyrene
change in the chemical shift was observed for the DNSA signals. ring of each probe to DNSA upon excitation. Due to the non-
When the same NMR experiment was conducted in (DMSO- radiative character of DNSA, this energy transfer results in emis-
d6:D2O) (6:2) instead of DMSO only, we could not still observe any sion quenching of the complexes. The HOMO e LUMO energy gaps
significant changes in the chemical shifts of the proton signals of (DE) of all tested probes were significantly reduced upon
probe 1 in the presence of DNSA (Fig. 7 and Fig. S16). Thus, the complexation with DNSA compared to that of the respective probe
chemical shifts for the protons of both the probes and DNSA were alone; this is mainly due to the change in the LUMO location from
not significantly affected by the molecular interactions between the the pyrene ring to the DNSA in all probe cases. The HOMO e LUMO
probes and DNSA as observed for DMSO-d6 and (DMSO-d6:D2O) energy gap of the [probe 1(DNSA)2] complex was comparable to

Fig. 6. (a) Stern-Volmer plot and (b) Stern-Volmer constant (Ksv (M1)) of DNSA for probes 1e3. Emission intensities at 484 nm for probe 1, 480 nm for probe 2, and 406 nm for
probe 3 were measured during the stepwise addition of DNSA (i.e., quencher) to individual probes (1 mM) dissolved in 99% water. Emission intensity ratios (Io/I) were calculated and
plotted against quencher concentration ([Q]), where Io and I are the fluorescence intensities in the absence and presence of DNSA, respectively, and [Q] is DNSA concentration.
A. Kumar, P. Seok Chae / Dyes and Pigments 147 (2017) 400e412 409

that of [probe 2(DNSA)2], but substantially larger than that of of interest. For the current system of probe 1 and DNSA, both
[probe 3(DNSA)] (i.e., DE1(DNSA)2 ¼ 1.99 eV > DE2(DNSA) fluorescent intensity and lifetime decreased with increasing
2 ¼ 1.96 eV > DE3 DNSA ¼ 1.29 eV). The spectral overlap between quencher concentration (Figs. 6A and 9B), where the fluorescence
the absorption spectrum of DNSA and the fluorescence spectrum of lifetime was reduced from tint ¼ 13.25 to 9.27 ns by the addition of
probe 1 suggests energy transfer from the probe to the guest 2.0 equivalents of DNSA. The linear Stern-Volmer plot and the
molecule as a quenching mechanism (Fig. 9A). This was further substantially decreased excited state lifetime in the presence of the
supported by fluorescence lifetime experiments. Fluorescence guest molecule suggest that a dynamic process is responsible for
emission can be quenched by either static or dynamic process. the fluorescence quenching observed for probe 1 with DNSA.
Alternatively, both processes could take place together in fluores- To investigate the potential of the current sensing system as a
cence quenching. In static quenching, a fluorophore forms a stable convenient and inexpensive sensing tool, silica-coated test strips
non-fluorescent complex with an analyte in the ground state, while were prepared using probe 1. We wrote the letters ‘DNSA’ on a TLC
in dynamic quenching; an excited fluorophore transfers its energy plate using a probe 1 solution (10 mM, DMSO:H2O (1:9)). We then
to an analyte (i.e., a quencher) in close proximity without a need for dried the plate and then observed the green emission of these
stable complex formation. Whether fluorescence quenching occurs letters under 365 nm irradiation (Fig. 10a). After dipping this TLC
in a static or dynamic process can be determined based on (a) the plate into a 1 mM solution of DNSA in tap water, the fluorescence
Stern-Volmer plot and (b) the excited state lifetime (t) of a sample intensity of the probe was significantly reduced (Fig. 10b). The

Fig. 7. (i) Partial 1H NMR spectra of probe 1 in the absence of DNSA or (ii) in the presence of 1 equiv. of DNSA, (iii) 2 equiv. of DNSA, and (iv) DNSA alone in (DMSO-d6:D2O) (6:2).

Fig. 8. Energy-minimized structures of probes 1e3 in the absence (left) or presence of DNSA (right), pep interactions between two pyrene rings or between a pyrene ring and DNSA
were indicated by yellow dotted lines, along with inter-ring distances. The structures were obtained from DFT calculations at the B3LYP/6-31G* level. (For interpretation of the
references to colour in this figure legend, the reader is referred to the web version of this article.)
410 A. Kumar, P. Seok Chae / Dyes and Pigments 147 (2017) 400e412

(1:9)) (Fig. 10c). When these spots were individually in contact with
water and solutions of 1, 10 and 100 nM DNSA, a notable quenching
of the green emission was observed even in DNSA concentrations as
low as 1 nM. This demonstrates the potential of these probes in real
time sensing applications (Fig. 10d).
Considering that probe 1 was highly sensitive to DNSA, a highly
electron-deficient compound, it is interesting to investigate the
behavior of the same probe with other highly electron-deficient
compounds such as 2,4,6-trinitrotoluene (TNT) and 2,4,6-
trinitrophenol (TNP). For this purpose, we tested various phenol
derivatives including TNP and TNT with probe 1 under the same
conditions tested above. Among the various phenol derivatives and
TNT, only TNP exhibited the selective fluorescence quenching of
probe 1 (Fig. S18). However, the binding strength of probe 1 for TNP
was 100 times smaller than that of the probe with DNSA
(Ka ¼ 1.80  105 vs. 2.92  107 M1). Furthermore, an observed LOD
for TNP sensing was 500 nM, significantly larger than that (1 nM)
for DNSA sensing. Thus, probe 1 is an optimal chemical probe for
selective and sensitive detection of DNSA. DNSA, TNT and TNP
commonly have a highly electron-deficient ring system due to the
presence of multiple electron-withdrawing substituents on the
benzene ring. However, probe 1 was not responsive to TNT at all,
indicating that electron deficiency of a guest molecule is not suf-
ficient for the fluorescence quenching of the probe. DNSA and TNP
respectively contain a carboxylic acid and hydroxyl group with the
ability to bind to hydrogen-bond acceptor (e.g., carbonyl group)
while TNT doesn't have such a functional group. Based on the
selectivity on DNSA and TNP over TNT, we conceive that the pres-
ence of hydrogen-bond donor in a guest molecule (e.g., carboxylic
or and hydroxyl group) may play a key role in the sensing system of
probe 1 although this hydrogen-bond was not evident in the mo-
lecular structure of probe 1-DNSA complex obtained from the DFT
Fig. 9. (A) Spectral overlap between the normalized absorption spectrum of DNSA and calculations.
normalized fluorescence spectrum of probe 1 in water. (B) Fluorescence lifetime decay
profile of probe 1 (1 mM) alone or in the presence of 2.0 equiv. of DNSA in water. The
inset shows a change in the ratio of excited state lifetime (t0/t) of probe 1 induced by 4. Conclusions
the presence of 2.0 equiv. of DNSA. t0 and t represent excited state lifetimes of the
probe in the absence and presence of the guest molecule, respectively. In summary, the dipodal pyreneamide-based probes (probes 1
and 2) displayed ACQ behaviors, while the monopod counterpart
detection limit of this system was investigated by preparing a TLC (probe 3) showed AIEE character at a high water content (>60%) in
plate with four spots using a probe 1 solution (10 mM, DMSO:H2O DMSO. In addition, with increasing concentration of water in

Fig. 10. Fluorescence quenching behaviors of DNSA for (a,b) written letters or (c,d) dropped spots on a TLC plate prepared using a probe 1 solution (10 mM, DMSO:H2O (1:9)).
Fluorescence intensity of the letters ‘DNSA’ was detected under 365 nm irradiation, (a) before or (b) after dipping the TLC plate into a 1 mM DNSA solution. Fluorescence intensities
of the spots on a TLC plate were measured under 365-nm irradiation (c) before and (d) after exposure to a drop of H2O or different concentrations of a DNSA solution: 1, 10, or
100 nM. 5 mL of water or DNSA solution of indicated concentration was added to the TLC plate.
A. Kumar, P. Seok Chae / Dyes and Pigments 147 (2017) 400e412 411

DMSO, the dipodal probes gave strong fluorescence emission at 480 J Am Acad Dermatol 2014;70:788e92.
[8] (a) Huang JH, Wang G, Huang KL. Enhanced adsorption of salicylic acid onto a
or 484 nm upon excitation as a result of the high tendency to form
b-naphthol-modified hyper-cross-linked poly(styrene-co-divinylbenzene)
the pyrene dimer (i.e., excimer). This excimer emission was resin from aqueous solution. Chem Eng J 2011;168:715e21.
completely quenched in the presence of 2 equiv. of DNSA. A (b) Chou WL, Wang CT, Huang KY, Liu TC. Electrochemical removal of salicylic
conformational change of the probes from a closed to an open acid from aqueous solutions using aluminum electrodes. Desalination
2011;271:55e61.
structure and the resulting energy transfer from the probe to the (c) Niamlang S, Sirivat A. Electric field assisted transdermal drug delivery from
guest molecule upon excitation are mainly responsible for the high salicylic acid-loaded polyacrylamide hydrogels. Drug Deliv 2009;16:378e88.
sensitivity of these probes toward DNSA detection. The enhanced (d) Lück E, Jager M. Antimicrobial food additives: characteristics, uses, effects.
2nd rev. and enl. ed. Berlin; New York: Springer; 1997.
selectivity of probe 1 toward DNSA detection compared to probe 2 [9] (a) Shakirova FM, Sakhabutdinova AR, Bezrukova MV, Fatkhutdinova RA,
is, at least in part, attributed to its conformational rigidity caused by Fatkhutdinova DR. Plant Sci 2003;164:317.
a short linker between the central MDP unit and pendant pyrene (b) Raskin I, Annu. Rev Plant Physiol 1992;43:439.
(c) Fu ZQ, Yan SP, Saleh A, Wang W, Ruble J, Oka N, et al. Nature 2012;486:228.
rings. In the case of probe 3 (with no excimer emission), DNSA (d) Klessig DF, Malamy J. Plant Mol Biol 1994;26:1439.
association also resulted in fluorescence quenching, but with less [10] (a) Tsai JC, Chuang SA, Hsu MY, Sheu HM. Int J Pharm 1999;188:145.
sensitivity and selectivity than probe 1. Among the three probes, (b) Mikami E, Goto T, Ohno T, Matsumoto H, Nishida M. J Pharm Biomed
2002;28:261.
probe 1 exhibited the highest selectivity and sensitivity toward [11] (a) Gutmann H, Johnson C, Keyzer H, Molnar J. Charge transfer complexes in
DNSA detection, reaching an LOD of 1 nM in an aqueous solution. biochemistry systems. Marcel Dekker Inc; 1992.
This ultra-sensitivity was well retained in an application of a probe (b) Kidwai M, Saxena S, Rastogi S, Venkataramanan R. Pyrimidines as anti-
infective agents. Curr Med Chem Anti-Infect Agents 2003;2:269e86.
1-coated TLC strip for DNSA detection. DFT calculations supported
(c) Salem H. Spectrophotometric determination of beta-adrenergic blocking
that quenching the fluorescence emission of the probes by DNSA agents in pharmaceutical formulations. J Pharm Biomed Anal 2002;29:
was mostly caused by energy transfer from electron-rich pyrene to 527e38.
electron-deficient DNSA in an excited state. The current result is a (d) Bazzi SH, Mostafa A, Alqaradawi SY, Nour E. Synthesis and spectroscopic
structural investigations of the charge-transfer complexes formed in the re-
step toward the development of new sensing molecules for DNSA action of 2,6-diaminopyridine with p-acceptors TCNE, chloranil, and DDQ.
detection. J Mol Struct 2007;842:1e5.
(e) Mandal R, Lahiri SC. Interactions of L-amino acids with metronidazole and
tinidazole. J Indian Chem Soc 1999;76:347e9.
Acknowledgements [12] (a) Sumner JB. The estimation of sugar in diabetic urine using dinitrosalicyclic
acid. J Biol Chem 1924;62:287e90.
(b) Miller GL. Use of dinitrosalicylic acid reagent for determination of reducing
This work was supported by the National Research Foundation sugar. Anal Chem 1959:426e8.
of Korea funded by the Korean government (MSIP) (grant number (c) Lindsay H. A colorimetric estimation of reducing sugars in potatoes with
2008-0061891 to P.S.C. and A.K.). 3,5-dinitrosalicylic acid. Potato Res 1973;16:176e9.
(d) ??.
[13] (a) Kumar A, Pandith A, Kim HS. Pyrene-appended imidazolium probes as 3,5-
Appendix A. Supplementary data dinitrosalicylic acid sensors in 10% aqueous media. Dye Pig 2015;122:351e8.
(b) Kumar A, Ghosh MK, Choi CH, Kim HS. Selective fluorescence sensing of
salicylic acids using a simple pyrenesulfonamide receptor. RSC Adv 2015;5:
Supplementary data related to this article can be found at http:// 23613e21.
dx.doi.org/10.1016/j.dyepig.2017.08.039. (c) Croubels S, Maes A, Baert K, De Backer P. Quantitative determination of
salicylic acid and metabolites in animal tissues by liquid chromatogra-
phyetandem mass spectrometry. Anal Chim Acta 2005;529:179e87.
References (d) Kumar A, Pandith A, Kim HS. Pyrenebutylamidopropylimidazole as a
multi-analyte sensor for 3,5-dinitrosalicylic acid and Hg2þ ions. J Lumin
[1] (a) Afran M, Athar HR, Ashar M. Does exogenous application of salicylic acid 2016;172:309e16.
through the rooting medium modulate growth and photosynthetic capacity in [14] (a) Uzarevic K, Halasz I, Dilovic I, Bregovic N, Rubcic M, Matkovic-Calogovic D,
two differently adapted spring wheat cultivars under salt stress? J Plant et al. Dynamic molecular recognition in solid state for separating mixtures of
Physiol 2007;164:685e94. isomeric dicarboxylic acids. Angew Chem Int Ed 2013;52:5504e8.
(b) Styrer L, Biochemistry, Freeman WH, New York, third ed., p. 188, 373e394, (b) Chou HC, Hsu CH, Cheng YM, Cheng CC, Liu HW, Pu SC, et al. Multiple
376 and 575. hydrogen bonds tuning guest/host excited-state proton transfer reaction: its
[2] (a) Gunes A, Inal A, Alpaslan M, Eraslan F, Bagci EG, Cicek N. Salicylic acid application in molecular recognition. J Am Chem Soc 2004;126:1650e1.
induced changes on some physiological parameters symptomatic for oxida- (c) Saliuga P, Kaur N, Kang J, Singh N, Jang DO. Benzimidazole-based chro-
tive stress and mineral nutrition in maize (Zea mays L.) grown under salinity. mogenic chemosensor for the recognition of oxalic acid via counter ion
J Plant Physiol 2007;164:728e36. displacement assay in semi-aqueous medium. Tetrahedron 2013;69:9001e6.
(b) Maynor MS, Nelson TL, O'Sullivan C, Lavigne JJ. A food freshness sensor (d) Jackson SL, Rananaware A, Rix C, Bhosale SV, Latham K. Highly fluorescent
using the multistate response from analyte-induced aggregation of a cross- MetalOrganic framework for the sensing of volatile organic compounds.
reactive poly(thiophene). Org Lett 2007;9:3217e20. Cryst. Growth Des 2016;16:3067e71.
(c) Zeikus JG, Jain MK, Elankovan P. Biotechnology of succinic acid production [15] (a) Goswami S, Jana S, Fun HK. Recognition of dicarboxylic acids by 3,30-
and markets for derived industrial products. Appl Microbiol Biotechnol bipyridine amide based receptors and its supramolecular behavior in solid
1999;51:545e52. state. Cryst Eng Comm 2008;10:507e17.
[3] James D, Scott SM, Ali Z, O'Hare WT. Chemical sensors for electronic nose (b) Aakeroy CB, Desper J, Leonard B, Urbina JF. Toward high-yielding supra-
systems. Microchim Acta 2005;149:1e17. molecular Synthesis: directed assembly of ditopic imidazoles/benzimidazoles
[4] (a) MacFabe DF, Cain DP, Rodriguez-Capote K, Franklin AE, Hoffman JE, Boon F, and dicarboxylic acids into cocrystals via selective OH$$$N hydrogen bonds.
et al. Neurobiological effects of intraventricular propionic acid in rats: possible Cryst Growth Des 2005;5:865e73.
role of short-chain fatty acids on the pathogenesis and characteristics of (c) Kobaisi MA, Bhosale SV, Latham K, Raynor AM, Bhosale SV. Functional
autism spectrum disorders. Behav Brain Res 2007;176:149e69. naphthalene diimides: synthesis, properties, and applications. Chem Rev
(b) Gates AT, Fakayode SO, Lowry M, Ganea GM, Murugeshu A, Robinson JW, 2016;116:11685e796.
et al. Gold nanoparticle sensor for homocysteine thiolactone-induced protein [16] (a) Jadhav JR, Ahmad MW, Kim HS. New 2-aminoethylimidazole-based
modification. Langmuir 2008;24:4107e13. dicarboxylic acid receptor derived from cholestane. Tetra Lett 2010;51:
[5] (a) Raymond R. “Adipic acid”, handbook of pharmaceutical excipients. 2009. 5954e8.
p. 11e2. (b) Ahmad MW, Kim SH, Kim HS. Pyrenyl-appended imidazolium receptor for
(b) Verhoff FH. “Citric acid”, Ullmann's encyclopedia of industrial chemistry. selective fluorescence sensing of oxalic acid. Tetra Lett 2011;52:6743e7.
Weinheim: Wiley-VCH; 2005. [17] (a) Ghosh K, Saha I, Masanta G, Wang EB, Parish CA. Triphenylamine-based
(c) Adam V, Zehnalek J, Petrlova J, Potesil D, Sures B, Trnkova L, et al. Phy- receptor for selective recognition of dicarboxylates. Tetra Lett 2010;51:343e7.
tochelatin modified electrode surface as a sensitive heavy metal ion biosensor. (b) Moriuchi T, Yoshida K, Hirao T. Chirality-organized ferrocene receptor
Sensors 2005;5:70e84. bearing podand dipeptide chains (l-Ala-l-Pro-NHPyMe) for the selective
[6] TiloGrosser ES, FitzGerald GA. Anti-Inflammatory, antipyretic, and analgesic recognition of dicarboxylic acids. Org Lett 2003;5:4285e8.
agents. In: Brunton LL, editor. Goodman and Gilman's the pharmacological (c) Kusukawa T, Toyama K, Takeshita S, Tanaka S. Fluorescent detection of
basis of therapeutics. twelfth ed. New York: McGraw-Hill Co; 2011. p. 977e82. amidinium carboxylate and amidinium formation using anthracene-based
[7] Madan RK, Levitt J. A review of toxicity from topical salicylic acid preparations. diamidine: an application for the analysis of dicarboxylic acid binding.
412 A. Kumar, P. Seok Chae / Dyes and Pigments 147 (2017) 400e412

Tetrahedron 2012;68:9973e81. for TNP sensing in water. Sens Actuator B-Chem 2017;240:1e9.
[18] (a) Kumar A, Chae PS. A simple and dual responsive ultrasensitive thioether- [19] (a) Avasthi K, Shukla L, Kant R, Ravikumar K. Folded conformations due to
functionalized pyrenesulfonamide for the cascade detection of mercury ion arene interactions in dissymmetric and symmetric butylidene-linker models
and dithiouracil, a mimetic system for molecular logic gates. Sens Actuator B- based on pyrazolo[3,4-d]pyrimidine, purine and 7-deazapurine. Acta Cryst
Chem 2017;251:416e26. 2014;C70:555e61.
(b) Kumar A, Chae PS. Electronically tuned sulfonamide-based probes with (b) Avasthi K, Shukla L, Kant R. Synthesis, 1H NMR and X-ray crystallographic
ultra-sensitivity for Ga3þ or Al3þ detection in aqueous solution. Anal Chim studies of three isomeric butylidene linker protophanes. Chem Bio Interface
Acta 2017;958:38e50. 2014;4(5):292e300.
(c) Kumar A, Chae PS. New 1,8-naphthalimide-conjugated sulfonamide probes

You might also like