Water-Soluble Nonconjugated Polymer Nanoparticles with Strong

Jul 29, 2016 - Britton–Robinson (BR) buffer solutions (0.04 M) were prepared according to standard protocols. Instruments. Fluorescence spectra, inc...
0 downloads 11 Views 2MB Size
Subscriber access provided by Northern Illinois University

Article

Water-Soluble Nonconjugated Polymer Nanoparticles with Strong Fluorescence Emission for Selective and Sensitive Detection of Nitro-Explosive Picric Acid in Aqueous Media Shi Gang Liu, Dan Luo, Na Li, Wei Zhang, Jinglei Lei, Nian Bing Li, and Hong Qun Luo ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/acsami.6b07407 • Publication Date (Web): 29 Jul 2016 Downloaded from http://pubs.acs.org on July 30, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Applied Materials & Interfaces is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Water-Soluble Nonconjugated Polymer Nanoparticles with Strong Fluorescence Emission for Selective and Sensitive Detection of Nitro-Explosive Picric Acid in Aqueous Media Shi Gang Liu †, Dan Luo †, Na Li †, Wei Zhang ‡, Jing Lei Lei §, Nian Bing Li †*, and Hong Qun Luo †* Key Laboratory of Eco-environments in Three Gorges Reservoir Region (Ministry of



Education), School of Chemistry and Chemical Engineering, Southwest University, Chongqing 400715, P.R. China Chongqing Institute of Green and Intelligent Technology, Chinese Academy of



Sciences, Chongqing 400714, P.R. China School of Chemistry and Chemical Engineering, Chongqing University, Chongqing

§

400044, P.R. China

*

Corresponding Authors

*

Nian Bing Li

*

2, Tiansheng Road, BeiBei District, Chongqing, 400715, China. Tel: +86 23

68253237; fax: +86 23 68253237; E-mail address: [email protected] *

Hong Qun Luo

*

2, Tiansheng Road, BeiBei District, Chongqing, 400715, China. Tel: +86 23

68253237; fax: +86 23 68253237; E-mail address: [email protected]

1

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 37

ABSTRACT A kind of water-soluble nonconjugated polymer nanoparticles (PNPs) with strong fluorescence emission was prepared by hyperbranched polyethyleneimine (PEI) and D-glucose

via Schiff base reaction and self-assembly in aqueous phase. The preparation

of the PEI- D-glucose (PEI-G) PNPs was facile (one-pot reaction) and environmentally friendly under mild conditions. Also, the PEI-G PNPs showed a high fluorescence quantum yield in aqueous solution, and the fluorescence properties (such as concentration-dependent and solvent-dependent fluorescence) and the origin of intrinsic fluorescence was investigated and discussed. The PEI-G PNPs were then used to develop a fluorescent probe for fast, selective, and sensitive detection of nitro-explosive picric acid (PA) in aqueous media, because the fluorescence can be easily quenched by PA whereas other nitro-explosives and structurally similar compounds only caused negligible quenching. A wide linear range (0.05 – 70 µM) and a low detection limit (26 nM) were obtained. The fluorescence quenching mechanism was carefully explored, and it was due to a combined effect of electron transfer, resonance energy transfer, as well as inner filter effect between PA and the PEI-G PNPs, which resulted in the good selectivity and sensitivity for PA. Finally, the developed sensor was successfully applied to detection of PA in environmental water samples. KEYWORDS: nonconjugated polymer nanoparticles, polyethyleneimine, intrinsic fluorescence, picric acid, fluorescent sensor

2

ACS Paragon Plus Environment

Page 3 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

INTRODUCTION Fluorescent polymers and polymer nanoparticles have attracted much interest in chemical sensors and bioanalysis applications in recent decades.1,2 Polymers with intrinsic fluorescence emission, namely constructed without the need of entrapment or covalent conjugation of fluorescent agents in the polymer, are almost conjugated polymers (CPs). Conjugated polymers possess conjugated main chain or π-aromatic building blocks (such as benzene ring and thiophene), which results in their strong fluorescence emission.3 However, conjugated polymers are often poorly water-soluble and thus, functional modifications to their side chains with charged moieties are inevitable for enhancement of their water solubility to fulfill analysis and biological applications. Another strategy to promote dispersity in water is to fabricate conjugated polymer nanoparticles (CPNs) by miniemulsion method, reprecipitation method, and so on. In recent years, numerous water-soluble conjugated polymers (WSCPs) and conjugated polymer nanoparticles have been developed and their applications in sensors, fluorescence imaging, and gene and drug delivery have been extensively explored.3-7 Nevertheless, several drawbacks still exist for their preparation and application, such as sophisticated multistep synthetic pathways, the use of environmentally unfriendly organic solvents, and the possibility of fluorescence self-quenching in aqueous solution. Therefore, the development of autofluorescent polymer materials with good water solubility and easy-preparation are very challenging and still in pursuit. Fortunately, in recent years researchers have been discovered that a few kinds of polymers like 3

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 37

poly(amidoamine) (PAMAM) could emit strong fluorescence under proper conditions,8-10 which own nonconjugated polymer structures and also lack typical chromophores. The fluorescence mechanism of these polymers is still under investigation. Increasing studies show that the fluorescence-emitting moieties may be aliphatic tertiary amines and some heteroatom-containing double bonds (C=O, C=N, N=O), which traditionally are not regarded as fluorescent chromophores.11 Because of their special structures and ample terminal groups, these fluorescent nonconjugated polymers are very promising as novel fluorescent materials for applications in chemical and biological fields. Picric acid (PA), namely 2,4,6-trinitrophenol (TNP), is a kind of nitroaromatic compound. Although, 2,4,6-trinitrotoluene (TNT) is the most widely used nitro-explosive, PA is another superior explosive material that possesses low safety coefficient and high detonation velocity. Meanwhile, PA has wide applications in various industries such as antiseptics, pesticides, and dye industry. When exposed, PA is possible to contaminate soil and groundwater because of its good solubility in water. At a low concentration, PA can cause severe effects on the skin, eyes, liver and kidneys, respiratory system, and metabolism, etc.12-14 Thus, it is very necessary to develop effective and reliable methods for the detection of trace amounts of PA in aqueous solution. A variety of techniques such as X-ray diffraction,15 surface-enhanced Raman spectroscopy,16 mass spectrometry,17 surface plasmon resonance,18 colorimetry,19 and electrochemistry20 have been used for the detection of PA. However, most of these methods are not accessible because of the requirements for expensive and bulky

4

ACS Paragon Plus Environment

Page 5 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

equipment, complicated manipulations, and time-consuming procedures. As an alternative approach, fluorescent sensors have such advantages as high sensitivity, low cost, and convenience. Hence, many fluorescent sensors has been designed for PA detection based on conjugated polymers,21 semiconductor quantum dots,22 carbon dots,23 metal-organic frameworks,24,25 small molecular fluorophores,26-28 metal nanoclusters,29 and so on.30-32 But most of the sensors for PA detection still suffer from some obstacles hampering their application, for example, the interference from other nitroaromatics, the complex synthesis of conjugated polymers, and the environmental and health risks of semiconductor quantum dots and metal-organic frameworks. So it is desirable and challenging to develop new fluorescent materials with a simple preparation, low or non-toxic nature, and high sensitivity and selectivity for PA detection. In this work, a kind of water-soluble nonconjugated polymer nanoparticles with strong fluorescence emission was constructed by hyperbranched polyethyleneimine (PEI) and D-glucose via Schiff base reaction and self-assembly. The preparation was under a mild condition (80 oC) without the need of organic solvents. The fundamental properties (such as concentration-dependent and solvent-dependent fluorescence) of the PEI-D-glucose polymer nanoparticles (PEI-G PNPs) have been investigated and also, their fluorescence origin was explored. It can be found that the preparation of the fluorescent PEI-G PNPs is not a modification of fluorescent conjugated polymers or incorporation of fluorescent dyes into polymer matrix. So in the study we not only have created a new class of fluorescent polymer nanoparticles but also provide a new

5

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 37

strategy to synthesize fluorescent polymer materials. The fluorescence of the PEI-G PNPs can be easily quenched by PA whereas other nitro-explosives and structurally similar compounds only caused negligible quenching. Thus, we have used the PEI-G PNPs to develop a fluorescent probe for rapid, sensitive, and selective detection of PA in aqueous media. The possible mechanism of PA-induced fluorescence quenching was discussed and it was attributed to a combined effect of efficient electron transfer, resonance energy transfer, as well as inner filter effect between PA and the PEI-G PNPs.

EXPERIMENTAL SECTION Materials. Hyperbranched polyethyleneimine (PEI, Mw = 10 000, 99%) and D-glucose

(picric

were purchased from Aladdin Ltd., Shanghai, China. 2,4,6-Trinitrophenol

acid,

PA),

2,4,6-trinitrotoluene

(TNT),

m-dinitrobenzene

(m-DNB),

p-nitrotoluene (p-NT), nitrobenzene (NB), and 2,4-dinitrotoluene (2,4-DNT) were obtained from Tianjin Kermel Chemical Reagents Co., Ltd, Tianjin, China. Other reagents were of analytical reagent grade, and all the chemicals were used without further purification. Ultrapure water with a resistivity of 18.2 MΩ cm was used throughout the experiment. Britton-Robinson (BR) buffer solutions (0.04 M) were prepared according to standard protocols. Instruments. The fluorescence spectra including three-dimensional (3D) fluorescence spectra were collected with an F-2700 spectrofluorometer (Hitachi, Japan). The slits of both excitation and emission were fixed at 10 nm, and the scan wavelength speed was 1500 nm min−1. For 3D fluorescence measurement, the scan wavelength

6

ACS Paragon Plus Environment

Page 7 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

intervals were set as 10 nm. UV-vis absorption spectra were recorded using a UV-vis 2450 spectrophotometer (Shimadzu, Japan). Nuclear magnetic resonance (NMR) spectra were obtained on a Bruker AVANCE III 600 (600 MHz) (Bruker, Germany). Transmission electron microscopy (TEM) measurement was carried out with a JEM 1200EX transmission electron microscope (JEOL, Japan). The Fourier transform infrared (FT-IR) spectra were tested by the use of a Bruker IFS 113v spectrometer (Bruker, Germany). Zeta potential and hydrodynamic diameter measurements were performed on a Zetasizer Nano-ZS90 (Malvern Instruments Ltd., U.K.). Fluorescence lifetime decays were measured using an Edinburgh FL 920 fluorescence spectrometer (Edinburgh, U.K.). A PHS-3C pH meter (Shanghai Leici Instrument Company, Ltd., China) was utilized to detect pH values of solutions. Preparation of PEI-G Polymer Nanoparticles. The PEI-G PNPs were prepared by PEI and D-glucose via Schiff base reaction and self-assembly (Scheme 1). Typically, 50 µL of 0.1 g mL-1 PEI was first dissolved in 400 µL of Britton-Robinson (BR) buffer (0.04 M, pH 5.0) by stirring for about 1 min, and then 50 µL of 1 M D-glucose solution was added. Then, the mixture was stirred and heated at 80 oC for 4 h via hydrothermal treatment. Subsequently, the as-prepared PEI-G PNPs solution was dialyzed against ultrapure water through a dialysis membrane (molecular weight cutoff of 3000 Da) for 24 h. The product inside the dialysis bag was collected for further study. Procedures for Sensing Picric Acid. For a typical PA detection procedure, 5 µL of PEI-G PNPs was added to 395 µL of BR buffer (0.04 M, pH 7.0), and the solution was mixed. Then 100 µL of PA solution with various concentrations was added to the 7

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 37

mixtures, followed by shaking well. At last, the fluorescence emission spectra were recorded under excitation at 345 nm. All measurements were carried out at room temperature. For analysis of environmental water samples, tap water was straight collected from our laboratory and river water was obtained from the Jialing River (Chongqing, China). All raw water samples were filtered through a 0.22 µm membrane to remove the suspended impurities, and ethylenediaminetetraacetate (EDTA) with final concentration of 50 µM was added to the samples. Then, the water samples were spiked with ultrapure water or different concentrations of standard PA solutions. Other procedures were the same as described above.

RESULTS AND DISCUTION Preparation and Characterization of PEI-G Polymer Nanoparticles. The synthesis of the PEI-G copolymer is based on Schiff base reaction, which refers to the reaction between compounds containing amino groups and aldehydes (or ketones) resulting in a product containing C=N bonds.33 Hyperbranched PEI, one of dendritic polymers, contains abundant amino groups. When PEI reacted with D-glucose, not only single Schiff base bonds but also conjugated double Schiff base bonds would be formed because in addition to aldehyde group, α-hydroxyl group of D-glucose was also able to react with amino group to produce a C=N bond because the α-hydroxyl group can be transformed to carbonyl group via oxidation or Amadori rearrangement.34,35 Scheme 1 illustrates the principle of reaction between PEI and D-glucose. The Schiff base reaction can occur under a mild condition, and a weak acidic condition will be conducive to it as

8

ACS Paragon Plus Environment

Page 9 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

in the reaction, a proton attacks carbonyl oxygen firstly and then nucleophilic attack takes place on the carbonyl carbon by the primary amine.36,37 Both PEI and D-glucose own favorable water solubility and thereby, the synthesis was under an aqueous solution with a pH 5.0 at 80 oC.

Different from the colorless PEI or D-glucose, the PEI-G PNPs are deep yellow, and the diluted PEI-G PNPs solution emits strong blue fluorescence under a 365 nm UV-lamp (inset of Figure 1A). As exhibited in Figure 1A, the UV-vis absorption spectra show that the PEI-G PNPs have a new absorption peak located at 358 nm, whereas PEI and D-glucose have no significant absorption at above 260 nm. The new absorption band can be ascribed to n→π* transitions of C=N bonds. Figure 1B shows FT-IR spectra of PEI, D-glucose, and PEI-G PNPs, respectively. Several feature absorption bands at the region from 1283 to 1593 cm−1 in PEI are associated with the stretching vibration of N-H bond, and their intensity is decreased in PEI-G PNPs which reveals that some amine groups have reacted with D-glucose. The absorption peaks at 2930 and 2835 cm−1 corresponding to the stretching vibration of CH2 bonds were found in the spectra of both PEI and PEI-G PNPs. The PEI-G PNPs have characteristic absorption centred at 3427 cm−1 that is the overlap of the absorptions of N-H bond and O-H bond, and the stretching vibration of C-O is situated at 1055 cm−1, demonstrating the presence of C-OH in PEI-G PNPs. Compared to the spectra of PEI and D-glucose, another remarkable new peak at 1627 cm−1 is observed in the PEI-G PNPs spectrum, which is assigned to the C=N bond.38 In addition, the 1H NMR spectra of PEI and PEI-G PNPs are shown in Figure S1 in the Supporting Information. A new peak at 8.00 ppm 9

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 37

belonging to N=CH protons38 and an obvious increase in intensity of peaks centred at 2.65 ppm (contribution from CH, CH2 of D-glucose segment) on the 1H NMR spectrum of PEI-G PNPs are observed. The results demonstrate the formation of Schiff base bonds and PEI-G copolymer. The zeta potential of PEI-G PNPs at 25 oC in water was found to be 11.2 ± 0.7 mV, which was much lower than that of PEI (46.5 ± 1.3 mV), further indicating that amino groups of the PEI in the PEI-G PNPs had partly reacted with D-glucose. Figure 1C is a TEM image and reveals that the PEI-G PNPs are monodisperse with typical spherical shape and a size around 90 nm. The average size of the hydrodynamic size of the PEI-G PNPs in water was 342 nm measured by dynamic light scatting (DLS) (Figure 1D), which is larger than the diameter measured by TEM. The possible reason is that DLS gives a hydrodynamic diameter, while TEM image exhibits a size and shape of dry particles.39

The formation of the water-soluble nanoparticles is due to the following factors. In PEI-G copolymer, ample amine groups and hydroxyl groups are hydrophilic, whereas Schiff base bonds are hydrophobic. As a result, the hyperbranched structure of PEI-G copolymer trends to fold and collapse, shrinking and self-assembling into uniform polymer nanoparticles in aqueous media.40 And many hydrophilic groups on the surface of the PEI-G PNPs make the excellent water dispersity possible. The PEI-G PNPs can keep stable for at least six months when stored at 4 oC, suggesting satisfied long-term storage stability. Overall, the preparation of the PEI-G PNPs in this work is much facile and environmentally friendly in comparison with other methods for constructing polymer nanoparticles, such as emulsion polymerization and 10

ACS Paragon Plus Environment

Page 11 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

nano-reprecipitation that are complicated and often operated in organic solvents.7

Scheme 1. Synthetic strategy of PEI-G PNPs

Figure 1. (A) UV-vis absorption spectra of PEI (0.1 mg L-1), D-glucose (1 mM), and PEI-G PNPs

11

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 37

(2% v/v), respectively. Inset is photographs of PEI-G PNPs under visible light and UV light of 365 nm. (B) FT-IR spectra of D-glucose (a), PEI (b), and PEI-G PNPs (c), respectively. (C) TEM image of PEI-G PNPs. (D) Hydrodynamic size of PEI-G PNPs measured by dynamic light scatting.

Fluorescence Properties of PEI-G Polymer Nanoparticles. Extraordinarily bright fluorescence is exhibited from the PEI-G PNPs under ultraviolet lamp (365 nm), in contrast with the sole PEI or D-glucose that has insignificant fluorescence. Figure 2A displays the fluorescence excitation and emission spectra of the PEI-G PNPs solution (2% v/v) and the maximum excitation and emission wavelengths are 345 and 465 nm, respectively. Using quinine sulphate as a reference, the quantum yield was estimated to be as high as 46% in water, compared to that of other fluorescent materials obtained from PEI, such as PEI-capped Ag nanoclusters (3.8%) and PEI-functioned carbon dots (42.5%).41,42 When the excitation wavelength was varied from 320 to 400 nm, there was no dramatic change of the maximum emission wavelength (Figure 2B) which is different from the fluorescence of carbon dots prepared using PEI as a precursor that is excitation-dependent.43,44 But the fluorescence of the PEI-G PNPs shows apparently concentration-dependent. When the PEI-G PNPs are at a relatively low concentration ranges (0 to 3% v/v), their fluorescence intensity is proportional to the concentration and there is no obvious change of the excitation and emission wavelengths (Figure S2). However, when the concentration further increases, it is intriguing that the maximum excitation and emission wavelengths would be gradually red-shifted. The fluorescence can be tuned from blue to green by adjusting the concentration of the PEI-G PNPs. Figure S3A, S3B, and S3C shows the three-dimensional fluorescence spectra of 2%,

12

ACS Paragon Plus Environment

Page 13 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

5%, 10% (v/v) PEI-G PNPs solutions, respectively. It can be seen that the peaks of excitation and emission are red-shifted from λex = 345 nm and λem = 465 nm (2% v/v) to λex = 440 nm and λem = 508 nm (10% v/v), respectively. Photographs of the PEI-G PNPs solutions with different concentrations under visible light and UV light (365 nm) are

exhibited

in

Figure

S3D

and

S3E,

respectively,

from

which

the

concentration-dependent fluorescence can also be found. The phenomenon might be because of the formation of higher order aggregated species at high concentration.45,46 The red shift of fluorescence was also observed when the PEI-G PNPs were dispersed

in

organic

solvents

including

dimethylsulfoxide

(DMSO)

and

dimethylformamide (DMF). Conventional Schiff base polymers are usually insoluble in almost all solvents.47,48 In this work, the good water solubility of the PEI-G PNPs stems from the ample hydrophilic amine groups and hydroxyl groups on their surface. Accordingly, the solubility of the PEI-G PNPs may dramatically descend with the decrease of solvent polarity. As anticipated, the PEI-G PNPs are hardly soluble in most organic solvents, except for relatively strong polar solvents (for example, DMSO, DMF, and ethanol) in which the PEI-G PNPs exhibit slightly soluble. As shown in Figure 2C, it can be observed that the fluorescence intensities of PEI-G PNPs in DMSO and DMF are increased and also the emission wavelengths exhibit some red-shifts in comparison with those in water because of the formation of aggregated species induced by the poor solubility. Figure 2D are photographs of the PEI-G PNPs dispersed in different solvents under visible and UV light (365 nm), illustrating the various solubility in water, DMSO, DMF, ethanol, and tetrahydrofuran (THF), respectively. Figure S4 reveals that the 13

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 37

gradual changes of solubility and fluorescence in an aqueous alcoholic solution with increasing ethanol contents. The PEI-G PNPs with excellent solubility in water and poor solubility in organic solvents are different from water-soluble conjugated polymers which usually are amphiphilic through incorporating pendant charged functionalities on their hydrophobic backbones.3 Additionally, we also investigated the effects of preparation conditions (including concentration of reactants and reaction time) on the fluorescence. As shown in Figure S5A and S5B, increasing the concentration of D-glucose (or PEI), the fluorescence intensity would increase. But with higher concentration of reactants, some insoluble precipitates would be produced. As for the effect of reaction time, similarly, high fluorescence intensity would be obtained with the increase of heating time (Figure S5C), but precipitates also appeared when the reaction time exceeded 4 h. Therefore, the conditions of 0.1 M D-glucose and 0.01 g mL-1 PEI at 80 oC for 4 h were chosen for the preparation of PEI-F PNPs.

14

ACS Paragon Plus Environment

Page 15 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Figure 2. (A) Fluorescence excitation and emission spectra of PEI-G PNPs (1% v/v). (B) Fluorescence emission spectra of PEI-G PNPs (1% v/v) under different excitations. (C) Fluorescence spectra of PEI-G PNPs (2% v/v) dispersed in H2O, DMF, and DMSO, respectively. Excitation: 345 nm. (D) Photographs of PEI-G PNPs (2% v/v) dispersed in H2O, DMSO, DMF, ethanol, and THF under visible light (a) and UV light of 365 nm (b), respectively.

Fluorescence Origin of PEI-G Polymer Nanoparticles. It is noticed that the PEI-G PNPs show strong intrinsic fluorescence without the conjugation to any external fluorochrome, and also the PEI-G PNPs lack typical chromophores such as conjugated main chain or π-aromatic building blocks functioning as emitting units. The high fluorescence emission is unexpected. In this regard, the PEI-G PNPs are very similar to a few kinds of polymers discovered in recent years which are fluorescent but have not conventional chromophores.9,10 And the fluorescence mechanism of these polymers is still under investigation. Published literature discovered that small molecular Schiff 15

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 37

base compounds can exhibit strong fluorescence by the suppression of C=N isomerisation via coordinating with metal ions.49,50 In our recent work, we have demonstrated that the fluorescence from particles and gels formed by PEI and formaldehyde derives from the radiative decay from excited Schiff base bonds.51 Consequently, in this work, the fluorescence of the PEI-G PNPs also may originate from the n←π* transitions of Schiff base bonds. Even though the PEI-G PNPs lack of continuous conjugated chain, the formation of compact and rigid nanoparticles may confine the intramolecular rotations of Schiff base bonds which suppress the nonradiative pathway and results in the radiative decay, thus, making the PEI-G PNPs fluorescent. To verify the fluorescence mechanism, we utilized PEI with different molecular weights to react with D-glucose. When PEI with various molecular weights (Mw = 600, 1800 and 10 000) was used, the fluorescence intensity increased with increasing molecular weight, but the maximum emission wavelength was constant (Figure S6). The result demonstrates that a larger hyperbranched macromolecular structure, which is conducive to the formation of Schiff bases and nanoparticles, is crucial to the fluorescence. This is also confirmed by another control experiment in which

small-molecular

amine

compounds

(such

diethylenetriamine) were utilized to react with

as

ethylenediamine

D-glucose

and

and no significant

fluorescence was found in their products.

To further shed light on the fluorescence origin, another experiment was performed. Generally, Schiff base bonds would be easily reduced by sodium borohydride (NaBH4) at room temperature to secondary amines via reaction (1):52 16

ACS Paragon Plus Environment

Page 17 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

In our previous report,50 the fluorescence of the PEI-formaldehyde polymer particles containing only single Schiff base bonds would disappear after reduced by NaBH4 (Figure S7). However, a different result was found in the PEI-G PNPs. The PEI-G PNPs (1% v/v) was reduced by 0.1 M NaBH4 (reacting for 4 h at room temperature). As shown in Figure 3A, the UV-vis absorption peak at 358 nm of the PEI-G PNPs is replaced by a weaker absorption band after reduction, which is not similar to that of the PEI-formaldehyde polymer particles whose characteristic absorption of C=N disappears entirely (Figure S7A). Moreover, as shown in Figure 3B, the fluorescence intensity of the PEI-G PNPs increases remarkably and also, the emission spectrum is blue-shifted from 465 to 455 nm and the excitation spectrum is red-shifted from 345 to 355 nm. The results should be because of the existence of conjugated double Schiff base bonds in the PEI-G PNPs. When reduced by NaBH4, the conjugated double Schiff base bonds would turn to C=C bonds by reaction (2):

It is known that the C=C bonds are unable to be reduced by NaBH4. As a result, the fluorescence from the n←π* transition of C=N converts to the π←π* transitions of C=C, resulting in the increase of fluorescence intensity and the shifts of excitation and emission spectra.

17

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 37

Figure 3. (A) UV-vis absorption spectra of PEI-G PNPs (1% v/v) before (i) and after (ii) reduced by NaBH4. (B) Fluorescence excitation and emission spectra of PEI-G PNPs (1% v/v) before (a, b) and after (c, d) reduced by NaBH4.

Establishment of Fluorescence Method for Sensing Picric Acid. The PEI-G PNPs possess intrinsic fluorescence, good water solubility, and ample surface amine groups and hydroxyl groups, suggesting a variety of potential applications in chemical and biological fields. We found that the fluorescence of the PEI-G PNPs can be easily quenched by PA, which enables us to develop a sensor for the detection of PA. At the beginning, several tests were performed to optimize the sensing conditions including probe concentration, reaction time, pH, and ionic strength. The fluorescence-quenched efficiency

(∆F/F0,

which

is

a

common

representation

method

for

fluorescence-quenched efficiency.53) was used to optimize the probe (∆F = F0 – F, where F0 and F represent the fluorescence intensity of PEI-G PNPs in the absence and presence of PA, respectively). First, the concentration of probe was optimized. Generally, a low concentration of probe would result in an efficient fluorescence quenching in the presence of fluorescence quencher of a given concentration, and thus high sensitivity to analyte 18

ACS Paragon Plus Environment

Page 19 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

would be found. However, the linear range of the detection would be narrow with the decline of the concentration of fluorescent probe.41 Figure S8A clearly illustrates that the lower the PEI-G PNPs concentration, the higher the fluorescence-quenched efficiency is when the concentration of PA is fixed. The concentration of 1% (v/v) PEI-G PNPs was eventually chosen for the detection of PA in order to obtain the wider linear range and the lower limit of detection. Second, response time is another key factor in the practical application of a sensor. The response rate of the fluorescence signal of the PEI-G PNPs to reaction time upon addition of PA was monitored subsequently. Figure S8B reveals that the fluorescence quenching can be finished within 1 min. This result demonstrates that the quenching of the PEI-G PNPs fluorescence by PA is rapid and stable, implying a promising application in a fast and convenient sensing of PA without strict time control. Third, the effect of pH on fluorescence quenching was investigated. Figure S8C shows that the fluorescence intensities of PEI-G PNPs in the absence and presence of PA decrease with increasing pH value and then trend to keep constant under weak alkaline conditions. But the fluorescence-quenched efficiency of PA at various pH values has not dramatic change (Figure S8D). As a result, a neutral BR buffer with pH 7.0 was chosen to apply to the following experiments. Finally, we explored the effect of ionic strength on sensing PA. The fluorescence of PEI-G PNPs in different concentrations of NaCl solutions in the absence and presence of PA was tested, respectively. As presented in Figure S8E and S8F, in solutions with different NaCl concentrations, the fluorescence intensity can remain stable and the fluorescence-quenched efficiencies of PA are almost same, even

19

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 37

in 1 M NaCl solution, suggesting that the PA sensor may have the ability to fit a complicated environment. Under optimum conditions, the linearity and the sensitivity was examined by recording emission spectra of the PEI-G PNPs after the addition of different concentrations of PA. Figure 4A exhibits that increasing the PA concentration gradually quenches the fluorescence, but no obvious change in the maximum emission wavelength was observed. And Figure 4B shows the relationship between the fluorescence intensity of PEI-G PNPs and the concentration of PA, which is not linear. It is known that the fluorescence quenching data generally follow the Stern-Volmer equation: F0/F – 1 = K[Q],54 where F0 and F denote fluorescence intensity of fluorophore in the absence and presence of quencher, respectively, K is the Stern-Volmer quenching constant, and [Q] is the quencher concentration. As shown in the insets of Figure 4B, the intensity ratio (F0 – F)/F (namely F0/F – 1) displays two good linear relationships versus PA concentration ranging from 0.05 to 1 µM (R2 = 0.9937) and 2 to 70 µM (R2 = 0.9922), where F0 and F are the fluorescence intensity of the PEI-G PNPs in the absence and presence of PA. The limit of detection (LOD) was estimated to be 26 nM based on a signal-to-noise of 3, indicating a high sensitivity. For the effective and specific detection of PA, minimal or no interference from other substances is an essential requirement. Therefore, the selectivity of PEI-G PNPs to PA was evaluated. Figure 5 shows the fluorescence responses of PEI-G PNPs to PA, 2,4,6-trinitrotoluene (TNT), m-dinitrobenzene (m-DNB), p-nitrotoluene (p-NT), nitrobenzene

(NB),

o-nitrophenol

(o-NP),

2,4-dinitrotoluene

(2,4-DNT), 20

ACS Paragon Plus Environment

Page 21 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

p-dihydroxybenzene (HQ), aniline (AN), benzoic acid (BA), o-dihydroxybenzene (CC), phenol (CA), 4-methylphenol (4-MP), and p-chlorophenol (p-CP). Among all these potential interferences and PA, it can be seen that only PA exhibits a substantial fluorescence quenching. Other nitroaromatics and phenols are unable to lead to remarkable interference, suggesting the highly selective sensing of PA by PEI-G PNPs. When applied to the analysis of PA in real water samples, the developed sensor may need the ability to resist interference of some common ions. As a consequence, the effect of 19 kinds of cations, including Na+, K+, Li+, Ag+, Ba2+, Mg2+, Zn2+, Ni2+, Co2+, Cd2+, Fe2+, Mn2+, Hg2+, Ca2+, Pb2+, Cu2+, Cr3+, Fe3+, and Al3+, on the fluorescence response of PEI-G PNPs were investigated. As exhibited in Figure S9A, most metal ions have no obvious effect on the fluorescence intensity, except that Cu2+, Ni2+, Fe3+, and Hg2+ may quench the fluorescence of the PEI-G PNPs, to some extent. Nevertheless,

with

the

addition

of

a

strong

metal

ion

chelator,

ethylenediaminetetraacetate (EDTA), the fluorescence quenching by the four metal ions can be effectively avoided. Furthermore, we also measured the fluorescence of the PEI-G PNPs in the presence of common anions (F-, Cl-, Br-, I-, NO3-, NO2-, Ac-, SO32-, SO42-, S2-, and CO32-) and Figure S9B displays that the 11 common anions barely affect the fluorescence intensity of PEI-G PNPs. The results indicate that the developed sensor can serve as a favorable and reliable tool to probe PA in aqueous media. A comprehensive comparison with other PA sensors (e.g., carbon dots, conjugated polymer, and small molecule) reported by the literature in recent years is summarized in Table S1, from which we can find that our sensor has several prominent merits in both

21

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 37

preparation and detection, such as facile synthesis in aqueous media with a relatively low temperature, lower detection limit, and wider linear range.

Figure 4. (A) Fluorescence emission spectra of PEI-G PNPs after adding various concentrations of PA. (B) Fluorescence intensity changes of the PEI-G PNPs in the presence of increasing concentration of PA (insets: linear relationships of (F0 – F)/F versus the concentration of PA over the range from 0 to 1 µM and 2 to 70 µM, where F0 and F denote the fluorescence intensity of the PEI-G PNPs before and after the addition of PA, respectively). Conditions: PEI-G PNPs, 1% (v/v); BR buffer (pH 7.0); excitation, 345 nm.

22

ACS Paragon Plus Environment

Page 23 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Figure 5. Selectivity to PA using PEI-G PNPs as a sensor. ∆F = F0 – F, where F0 and F denote the fluorescence intensity of PEI-G PNPs before and after the addition of analytes, respectively. The concentration of all the analytes is 50 µM. Conditions: PEI-G PNPs, 1% (v/v); BR buffer, pH 7.0; excitation, 345 nm.

Mechanism of Picric Acid-Induced Fluorescence Quenching. As is known, nitro-explosive molecules are nitro-substituted electron-acceptor substrates, and they can bind to π-donor sites via donor-acceptor interactions.29 Fluorescence of most electron-rich fluorophores can be effectively quenched by the nitro-explosive molecules via electron-transfer, which has been extensively utilized to design fluorescence probes for explosives detection over the past years. Numerous electron-rich amine groups and hydroxyl groups are on the surface of the PEI-G PNPs. Thus, when electron-deficient PA is added to the PEI-G PNPs solution, it may be bound to the surface of PEI-G PNPs, which causes the fluorescence quenching because of electron-transfer. However, the good selectivity to PA and the weaker quenching caused by other nitro-explosives such as TNT indicate that the electron-transfer may not be the main factor in the quenching process. Figure 6A shows the fluorescence emission spectrum of PEI-G PNPs and UV-vis absorption spectrum of PA. It is observed that there is an evident overlap between the florescence emission spectrum of PEI-G PNPs and absorption spectrum of PA. In contrast, the absorption spectra of other latent interferences have insignificant overlap with the emission spectrum of PEI-G PNPs (Figure S10). Therefore, there is a strong possibility of resonance energy transfer from the fluorophores of PEI-G PNPs to the PA, which can effectively enhance the fluorescence-quenched efficiency and sensitivity.23 The UV-vis absorption spectrum of 23

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 37

PEI-G PNPs with the addition of PA further demonstrates the energy transfer process. As shown in Figure 6B, the absorption of the mixture of PEI-G PNPs and PA is different from that of sole PEI-G PNPs or PA solution, and the shoulder peak of PA at around 410 nm disappeared, indicating that the PA has successfully bound to the surface of the PEI-G PNPs. A close proximity between PEI-G PNPs and PA makes the fluorescence resonance energy transfer possible.55 To further verify the nature of the fluorescence quenching, the fluorescence lifetime measurement was carried out. Generally, fluorescence lifetimes of fluorophores vary for dynamic quenching and are invariant for static quenching as a function of quencher concentration.56 Figure 6C and 6D show the time-resolved fluorescence spectra of PEI-G PNPs in the absence and presence of PA, respectively. The fluorescence lifetime of PEI-G PNPs keeps the same for both two cases (τ1 = 1.3 µs, τ2 = 8.5 µs), suggesting that a static quenching mechanism is accountable for the fluorescence quenching. Also, the fluorescence-quenched efficiency of PA at different temperatures further demonstrated the static quenching. It is known that higher temperatures result in faster diffusion and hence larger amounts of collisional quenching, while higher temperature will cause the dissociation of weakly bound complexes, and thus smaller amounts of static quenching.56 As shown in Figure 6E, the fluorescence-quenched efficiency decreases gradually with increasing temperature, proving the static quenching process. These facts support the energy transfer from PEI-G PNPs to PA. In addition, we found that the absorption band centered at 360 nm of PA has

24

ACS Paragon Plus Environment

Page 25 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

significantly overlapped with the excitation spectrum (345 nm) of the PEI-G PNPs (Figure 6F). Published literature reported that the absorption of light at the excitation or emission wavelength by compounds could reduce the fluorescence intensity of fluorophore which was called inner filter effect.57 As a result, the fluorescence quenching of the PEI-G PNPs by PA may also partly result from the inner filter effect related to the specific absorption of PA.

Figure 6. (A) UV-vis absorption spectrum of PA (50 µM) (curve a) and fluorescence emission spectra of PEI-G PNPs (1% v/v) (curve b). (B) UV-vis absorption spectra of PA (50 µM), PEI-G PNPs (2% v/v), and PEI-G PNPs (2% v/v) added to 50 µM PA, respectively. (C) Time-resolved fluorescence spectrum of PEI-G PNPs (1% v/v). (D) Time-resolved fluorescence spectrum of PEI-G PNPs (1% v/v) with the addition of 50 µM PA. (E) Fluorescence-quenched efficiency of PA at

25

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 37

different temperatures. (F) UV-vis absorption spectrum of PA (50 µM) (curve a) and fluorescence excitation spectrum of PEI-G PNPs (1% v/v) (curve b).

Hence, on the basis of above discussion, it can be concluded that the highly selectivity and sensitivity of PEI-G PNPs for PA detection in aqueous media is attributed to the efficient electron and energy transfer process, as well as the inner filter effect. Detection of Picric Acid in Environmental Water Samples. It is very challenging for a prepared fluorescent sensor to apply in real samples analysis because of potential unknown interferences. To assess the practicality, we applied the developed method to the detection of PA in tap water and Jialing River water samples. The samples were tested and the results are summarized in Table S2. No PA was found in both tap water and Jialing River water. For the PA-spiked samples, it can be seen that the values measured by this proposed method are consistent with the spiked values for the concentration of PA. The recoveries of PA in the samples range from 100.4 to 112.0%. In addition, the relative standard deviation (RSD) was obtained by repeating the experiment 6 times under the same conditions. The RSDs are in the range of 1.05% to 4.52%, suggesting a good repeatability of the proposed method. The data of the recovery value and the RSD are satisfactory, indicating that the sensor is reliable and applicable for PA detection in environmental water samples.

CONCLUSIONS In conclusion, we created a kind of nonconjugated PNPs with strong fluorescence

26

ACS Paragon Plus Environment

Page 27 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

emission by PEI and D-glucose via Schiff base reaction and self-assembly under a mild condition. The fundamental properties of the PEI-G PNPs have been studied and the possible mechanism of fluorescence was discussed. Ample hydrophilic groups (amine groups and hydroxyl groups) and hydrophobic Schiff base bonds make the hyperbranched structure of PEI-G copolymer fold and collapse, shrinking and self-assembling into uniform polymer nanoparticles in aqueous media. The formation of the nanoparticles makes fluorescence emission from the n←π* transitions of Schiff base bonds possible because compact and rigid nanoparticles confine the intramolecular rotations of Schiff base bonds. Because of the specific structure, the PEI-G PNPs exhibit excellent water solubility but poor dispersity in organic solvents. Furthermore, we have used the PEI-G PNPs to develop a fluorescent sensor of PA on the basis of the fluorescence quenching induced by PA due to the combined effect of the electron transfer, resonance energy transfer, and inner filter effect. This PEI-G PNPs provides a favorable and reliable platform for the rapid, sensitive, and selective detection of nitro-explosive PA in 100% aqueous media.

ASSOCIATED CONTENT Supporting Information 1

H NMR spectra, concentration-dependent fluorescence, 3D fluorescence spectra,

effects of solvent polarity, preparation conditions, and molecular weights (PEI) on fluorescence, optimization of sensing conditions, comparison of fluorescent sensors,

27

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 37

UV-vis absorption spectra, and data of real water analysis. The Supporting Information is available free of charge on the Internet at http://pubs.acs.org.

AUTHOR INFORMATION Corresponding Authors *

E-mail address: [email protected], [email protected].

Notes The authors declare no competing financial interest.

ACKNOWLEDGMENT

This work was financially supported by the National Natural Science Foundation of China (No. 21273174), the Municipal Science Foundation of Chongqing City (No. CSTC-2013jjB00002, CSTC-2015jcyjB50001) and Youth Innovation Promotion Association of CAS (No. 2015316).

28

ACS Paragon Plus Environment

Page 29 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

REFERENCES (1) Kim, H. N.; Guo, Z.; Zhu, W.; Yoon, J.; Tian, H. Recent Progress on Polymer-Based Fluorescent and Colorimetric Chemosensors. Chem. Soc. Rev. 2011, 40, 79-93. (2) Yao, J.; Yang, M.; Duan, Y. Chemistry, Biology, and Medicine of Fluorescent Nanomaterials and Related Systems: New Insights into Biosensing, Bioimaging, Genomics, Diagnostics, and Therapy. Chem. Rev. 2014, 114, 6130-6178. (3) Zhu, C.; Liu, L.; Yang, Q.; Lv, F.; Wang, S. Water-Soluble Conjugated Polymers for Imaging, Diagnosis, and Therapy. Chem. Rev. 2012, 112, 4687-4735. (4) Liu, B.; Wang, S.; Bazan, G. C.; Mikhailovsky, A. Shape-Adaptable Water-Soluble Conjugated Polymers. J. Am. Chem. Soc. 2003, 125, 13306-13307. (5) Wang, C.; Tang, Y.; Liu, Y.; Guo, Y. Water-Soluble Conjugated Polymer as a Platform for Adenosine Deaminase Sensing Based on Fluorescence Resonance Energy Transfer Technique. Anal. Chem. 2014, 86, 6433-6438. (6) Feng, X.; Liu, L.; Wang, S.; Zhu, D. Water-Soluble Fluorescent Conjugated Polymers and Their Interactions with Biomacromolecules for Sensitive Biosensors. Chem. Soc. Rev. 2010, 39, 2411-2419. (7) Feng, L.; Zhu, C.; Yuan, H.; Liu, L.; Lv, F.; Wang, S. Conjugated Polymer Nanoparticles: Preparation, Properties, Functionalization and Biological Applications. Chem. Soc. Rev. 2013, 42, 6620-6633. (8) Wang, D.; Imae, T. Fluorescence Emission from Dendrimers and Its pH Dependence. J. Am. Chem. Soc. 2004, 126, 13204-13205. 29

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 37

(9) Zhao, E.; Lam, J. W. Y.; Meng, L.; Hong, Y.; Deng, H.; Bai, G.; Huang, X.; Hao, J.; Tang, B. Z. Poly[(maleic anhydride)-alt-(vinyl acetate)]: A Pure Oxygenic Nonconjugated Macromolecule with Strong Light Emission and Solvatochromic Effect. Macromolecules 2015, 48, 64-71. (10) Lu, H.; Feng, L.; Li, S.; Zhang, J.; Lu, H.; Feng, S. Unexpected Strong Blue Photoluminescence Produced from the Aggregation of Unconventional Chromophores in Novel Siloxane–Poly(amidoamine) Dendrimers. Macromolecules 2015, 48, 476-482. (11) Zhu, S.; Song, Y.; Shao, J.; Zhao, X.; Yang, B. Non-Conjugated Polymer Dots with Crosslink-Enhanced Emission in the Absence of Fluorophore Units. Angew. Chem. Int. Ed. 2015, 54, 14626-14637. (12) Wollin, K.; Dieter, H. H. Toxicological Guidelines for Monocyclic Nitro-, Aminoand Aminonitroaromatics, Nitramines, and Nitrate Esters in Drinking Water. Arch. Environ. Contam. Toxicol. 2005, 49, 18-26. (13) Malik, A. H.; Hussain, S.; Kalita, A.; Iyer, P. K. Conjugated Polymer Nanoparticles for the Amplified Detection of Nitro-Explosive Picric Acid on Multiple Platforms. ACS Appl. Mater. Interfaces 2015, 7, 26968-26976. (14) Toal, S. J.; Trogler, W. C. Polymer Sensors for Nitroaromatic Explosives Detection. J. Mater. Chem. 2006, 16, 2871-2883. (15) Soldate, A. M.; Noyes, R. M. X-Ray Diffraction Patterns for the Identification of Crystalline Constituents of Explosives. Anal. Chem. 1947, 19, 442-444. (16) Gaft, M.; Nagli, L. UV Gated Raman Spectroscopy for Standoff Detection of

30

ACS Paragon Plus Environment

Page 31 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Explosives. Opt. Mater. 2008, 30, 1739-1746. (17) Talaty, N.; Mulligan, C. C.; Justes, D. R.; Jackson, A. U.; Noll, R. J.; Cooks, R. G. Fabric Analysis by Ambient Mass Spectrometry for Explosives and Drugs. Analyst 2008, 133, 1532-1540. (18) Shankaran, D.; Gobi, K.; Matsumoto, K.; Imato, T.; Toko, K.; Miura, N. Highly Sensitive Surface Plasmon Resonance Immunosensor for Parts-Per-Trillion Level Detection of 2,4,6-Trinitrophenol. Sens. Actuators, B 2004, 100, 450-454. (19) Dong, M.; Wang, Y. W.; Zhang, A. J.; Peng, Y. Colorimetric and Fluorescent Chemosensors for the Detection of 2,4,6‐Trinitrophenol and Investigation of Their Co‐Crystal Structures. Chem. Asian J. 2013, 8, 1321-1330. (20) Forzani, E. S.; Lu, D.; Leright, M. J.; Aguilar, A. D.; Tsow, F.; Iglesias, R. A.; Zhang, Q.; Lu, J.; Li, J.; Tao, N. A Hybrid Electrochemical-Colorimetric Sensing Platform for Detection of Explosives. J. Am. Chem. Soc. 2009, 131, 1390-1391. (21) Sohn, H.; Calhoun, R. M.; Sailor, M. J.; Trogler, W. C. Detection of TNT and Picric Acid On Surfaces and in Seawater by Using Photoluminescent Polysiloles. Angew. Chem. Int. Ed. 2001, 40, 2104-2105. (22) Freeman, R.; Finder, T.; Bahshi, L.; Gill, R.; Willner, I. Functionalized CdSe/ZnS QDs for the Detection of Nitroaromatic or RDX Explosives. Adv. Mater. 2012, 24, 6416-6421. (23) Sadhanala, H. K.; Nanda, K. K. Boron and Nitrogen Co-doped Carbon Nanoparticles as Photoluminescent Probes for Selective and Sensitive Detection of Picric Acid. J. Phys. Chem. C 2015, 119, 13138-13143.

31

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 32 of 37

(24) Shi, Z. Q.; Guo, Z. J.; Zheng, H. G. Two Luminescent Zn(II) Metal-Organic Frameworks for Exceptionally Selective Detection of Picric Acid Explosives. Chem. Commun. 2015, 51, 8300-8303. (25) Hu, Z.; Deibert, B. J.; Li, J. Luminescent Metal-Organic Frameworks for Chemical Sensing and Explosive Detection. Chem. Soc. Rev. 2014, 43, 5815-5840. (26) Roy, B.; Bar, A. K.; Gole, B.; Mukherjee, P. S. Fluorescent Tris-Imidazolium Sensors for Picric Acid Explosive. J. Org. Chem. 2013, 78, 1306-1310. (27) Yadav, A.; Boomishankar, R. Concentration Dependent Ratiometric Turn-On Selective Fluorescence Detection of Picric Acid in Aqueous and Non-Aqueous Media. RSC Adv. 2015, 5, 3903-3907. (28) Bhalla,

V.;

Gupta,

A.;

Kumar,

M.

Fluorescent

Nanoaggregates

of

Pentacenequinone Derivative for Selective Sensing of Picric Acid in Aqueous Media. Org. Lett. 2012, 14, 3112-3115. (29) Enkin, N.; Sharon, E.; Golub, E.; Willner, I. Ag Nanocluster/DNA Hybrids: Functional Modules for the Detection of Nitroaromatic and RDX Explosives. Nano Lett. 2014, 14, 4918-4922. (30) Rong, M.; Lin, L.; Song, X.; Zhao, T.; Zhong, Y.; Yan, J.; Wang, Y.; Chen, X. A Label-Free Fluorescence Sensing Approach for Selective and Sensitive Detection of 2,4,6-Trinitrophenol (TNP) in Aqueous Solution Using Graphitic Carbon Nitride Nanosheets. Anal. Chem. 2015, 87, 1288-1296. (31) Wang, Y.; Ni, Y. Molybdenum Disulfide Quantum Dots as a Photoluminescence Sensing Platform for 2,4,6-Trinitrophenol Detection. Anal. Chem. 2014, 86,

32

ACS Paragon Plus Environment

Page 33 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

7463-7470. (32) Sun, X.; Ma, X.; Kumar, C. V.; Lei, Y. Protein-Based Sensitive, Selective and Rapid Fluorescence Detection of Picric Acid in Aqueous Media. Anal. Methods 2014, 6, 8464-8468. (33) Jia, Y.; Li, J. Molecular Assembly of Schiff Base Interactions: Construction and Application. Chem. Rev. 2015, 115, 1597-1621. (34) Shemyakin, M. M., Maimind, V. I., Ermolaev, K. M., Bamdas, E. M. The Mechanism of Osazone Formation. Tetrahedron, 1965, 21, 2771-2777. (35) El, K. H. Chemistry of Osazones. Adv. Carbohydr. Chem. 1965, 20, 139-181. (36) Peelen, D.; Smith, L. M. Immobilization of Amine-Modified Oligonucleotides on Aldehyde-Terminated Alkanethiol Monolayers on Gold. Langmuir 2005, 21, 266-271. (37) Zammatteo, N.; Jeanmart, L.; Hamels, S.; Courtois, S.; Louette, P.; Hevesi, L.; Remacle, J. Comparison between Different Strategies of Covalent Attachment of DNA to Glass Surfaces to Build DNA Microarrays. Anal. Biochem. 2000, 280, 143-150. (38) Guo, L.; Wu, S.; Zeng, F.; Zhao, J. Synthesis and Fluorescence Property of Terbium Complex with Novel Schiff-Base Macromolecular Ligand. Eur. Polym. J. 2006, 42, 1670-1675. (39) Ma, Y.; Pan, G.; Zhang, Y.; Guo, X.; Zhang, H. Narrowly Dispersed Hydrophilic Molecularly Imprinted Polymer Nanoparticles for Efficient Molecular Recognition in Real Aqueous Samples including River Water, Milk, and Bovine Serum. Angew. Chem. Int. Ed. 2013, 52, 1511-1514. (40) Mavila, S.; Eivgi, O.; Berkovich, I.; Lemcoff, N. G. Intramolecular Cross-Linking

33

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 34 of 37

Methodologies for the Synthesis of Polymer Nanoparticles. Chem. Rev. 2016, 116, 878-961. (41) Qu, F.; Li, N. B.; Luo, H. Q. Polyethyleneimine-Templated Ag Nanoclusters: A New Fluorescent and Colorimetric Platform for Sensitive and Selective Sensing Halide Ions and High Disturbance-Tolerant Recognitions of Iodide and Bromide in Coexistence with Chloride under Condition of High Ionic Strength. Anal. Chem. 2012, 84, 10373-10379. (42) Dong, Y.; Wang, R.; Li, G.; Chen, C.; Chi, Y.; Chen, G. Polyamine-Functionalized Carbon Quantum Dots as Fluorescent Probes for Selective and Sensitive Detection of Copper Ions. Anal. Chem. 2012, 84, 6220-6224. (43) Liu, C.; Zhang, P.; Zhai, X.; Tian, F.; Li, W.; Yang, J.; Liu, Y.; Wang, H.; Wang, W.; Liu, W. Nano-Carrier for Gene Delivery and Bioimaging Based on Carbon Dots with PEI-Passivation Enhanced Fluorescence. Biomaterials 2012, 33, 3604-3613. (44) Dong, Y.; Wang, R.; Li, H.; Shao, J.; Chi, Y.; Lin, X.; Chen, G. Polyamine-Functionalized Carbon Quantum Dots for Chemical Sensing. Carbon 2012, 50, 2810-2815. (45) Yoon, M. J.; Cheon, Y. J.; Kim, D. H. Absorption and Fluorescence Spectroscopic Studies on Dimerization of Chloroaluminum (III) Phthalocyanine Tetrasulfonate in Aqueous Alcoholic Solutions. Photochem. Photobiol. 1993, 58, 31-36. (46) Qu, F.; Dou, L. L.; Li, N. B.; Luo, H. Q. Solvatofluorochromism of Polyethyleneimine-Encapsulated Ag Nanoclusters and Their Concentration-Dependent Fluorescence. J. Mater. Chem. C 2013, 1, 4008-4013.

34

ACS Paragon Plus Environment

Page 35 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

(47) Jenekhe, S. A.; Yang, C. J.; Vanherzeele, H.; Meth, J. S. Cubic Nonlinear Optics of Polymer Thin-Films: Effects of Structure and Dispersion on the Nonlinear Optical-Properties of Aromatic Schiff-Base Polymers. Chem. Mater. 1991, 3, 985-987. (48) Khuhawar, M. Y.; Mughal, M. A.; Channar, A. H. Synthesis and Characterization of Some New Schiff Base Polymers. Eur. Polym. J. 2004, 40, 805-809. (49) Wang, L.; Qin, W.; Liu, W. A Sensitive Schiff-Base Fluorescent Indicator for the Detection of Zn2+. Inorg. Chem. Commun. 2010, 13, 1122-1125. (50) Wu, J. S.; Liu, W. M.; Zhuang, X. Q.; Wang, F.; Wang, P. F.; Tao, S. L.; Zhang, X. H.; Wu, S. K.; Lee, S. T. Fluorescence Turn on of Coumarin Derivatives by Metal Cations: A New Signaling Mechanism Based on C=N Isomerization. Org. Lett. 2007, 9, 33-36. (51) Liu, S. G.; Li, N.; Ling, Y.; Kang, B. H.; Geng, S.; Li, N. B.; Luo, H. Q. pH-Mediated Fluorescent Polymer Particles and Gel from Hyperbranched Polyethylenimine and the Mechanism of Intrinsic Fluorescence. Langmuir 2016, 32, 1881-1889. (52) Billman, J. H.; Diesing, A. C. Reduction of Schiff Bases with Sodium Borohydride. J. Org. Chem. 1957, 22, 1068-1070. (53) Han, B.; Yuan, J.; Wang, E. Sensitive and Selective Sensor for Biothiols in the Cell Based on the Recovered Fluorescence of the CdTe Quantum Dots-Hg(II) System. Anal. Chem. 2009, 81, 5569-5573. (54) Lu, W.; Qin, X.; Liu, S.; Chang, G.; Zhang, Y.; Luo, Y.; Asiri, A. M.; Al-Youbi, A. O.; Sun, X. Economical, Green Synthesis of Fluorescent Carbon Nanoparticles and

35

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 36 of 37

Their Use as Probes for Sensitive and Selective Detection of Mercury(II) Ions. Anal. Chem. 2012, 84, 5351-5357. (55) Yang, Y.; Huang, J.; Yang, X.; Quan, K.; Wang, H.; Ying, L.; Xie, N.; Ou, M.; Wang, K. FRET Nanoflares for Intracellular mRNA Detection: Avoiding False Positive Signals and Minimizing Effects of System Fluctuations. J. Am. Chem. Soc. 2015, 137, 8340-8343. (56) Lakowicz, J. R. Principles of Fluorescence Spectroscopy, 3rd ed.; Springer: New York, 2006. (57) Weert, M.; Stella, L. Fluorescence Quenching and Ligand Binding: A Critical Discussion of a Popular Methodology. J. Mol. Struct. 2011, 998, 144-150.

36

ACS Paragon Plus Environment

Page 37 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

For TOC only

37

ACS Paragon Plus Environment