Furanone Moiety as Potential Optical Brightening ... - ACS Publications

Sep 7, 2017 - ABSTRACT: A series of novel heterocyclic polymers with fluorescent brightening properties are synthesized via Click polymerization. Fast...
2 downloads 14 Views 3MB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article http://pubs.acs.org/journal/acsodf

Preparation and Characterization of Poly-1,2,3-triazole with Chiral 2(5H)‑Furanone Moiety as Potential Optical Brightening Agents Jingpei Huo,*,†,‡,∥ Zhudong Hu,†,∥ Dongchu Chen,† Shihe Luo,§ Zhaoyang Wang,*,§ Yonghui Gao,† Min Zhang,† and Hong Chen† †

College of Materials Science and Energy Engineering, Foshan University, Foshan 528000, P. R. China School of Chemistry and Chemical Engineering, South China University of Technology, Guangzhou 510641, P. R. China § School of Chemistry and Environment, South China Normal University, Guangzhou 510006, P. R. China ‡

S Supporting Information *

ABSTRACT: A series of novel heterocyclic polymers with fluorescent brightening properties are synthesized via Click polymerization. Fast synthesis of poly-1,2,3-triazoles (Mn ≥ 9.31 kDa) is described herein, with a high yield of up to 95%. The Click polymerization approach has a number of advantages, including facile operation and outstanding isolation yield. The resultant polymers have a high thermal stability, excellent UV resistance, as well as acid and light fastness. On embedding with optical brightening agents, the polymers display strong fluorescent brightening properties in the tetrahydrofuran solution. Moreover, these products have a strong solution emission intensity and extraordinary photostability under UV light. be added onto the coating surface more easily.24,25 On the basis of our previous research,15,26 novel polymers, 2, possessing multifunctional moieties are synthesized by the Click reaction (Scheme 1), and their structures and properties are determined by IR spectra, 1H NMR, gel permeation chromatography (GPC), UV, ultraviolet protection factor (UPF), and the waterlaundering test.

1. INTRODUCTION Poly-1,2,3-triazole exhibits outstanding chemical stability in hostile environments, even at high temperatures. Meanwhile, it also retains its aromatic character, as well as the formation of hydrogen bonds, medicinal properties, and so forth.1,2 Owing to these features, materials based on the 1,2,3-triazole polymer have been receiving a surge of interest for a variety industrial and medicinal applications.3,4 The Click reaction, with obvious development, has been extensively applied for different polymers with front design, including telechelic star polymers, segmented copolymer, spirodendrimers, and so on.5,6 It is crucial to expand the application of poly-1,2,3-triazole and take full advantage of 1,2,3-triazole.7,8 In contrast, most optical brightening agents (OBAs) are used in textile materials, liquid detergents, and paper making.9,10 As a kind of unique OBA, stilbene derivatives possess a carbon− carbon double bond and lose their luminescent properties under light irradiation.11,12 Therefore, it is essential to protect OBAs against direct light to make them more stable for storage.13,14 Prior to the preparation of bis-1,2,3-triazole,15 poly-1,2,3triazole shows excellent thermal stability, high solubility, and luminescent intensity.16,17 Moreover, the molecules in poly1,2,3-triazole may form a longer aromatic ring conjugated system, in addition to promoting whiteness and complementary effect.18,19 In addition, incorporation of chiral 2(5H)-furanone in the polymers is an effective approach to improve their antifungal activities,20,21 chemical durabilities, and thermostabilities.22,23 Also, with the help of chiral 2(5H)-furanone, the polymers may © 2017 American Chemical Society

2. RESULTS AND DISCUSSION To the best of our knowledge, an important design principle was to control the ratio between dialkynyl and diazido, aiming at improving the photostability of the OBAs (Scheme 1). The incorporation of 2(5H)-furanone gave rise to the larger steric effect, thereby avoiding photoisomerization and, in turn, effectively improving the light stability of the OBAs, 2, as well as increasing their UPF values.12 The following discussions on the properties show these fluorescent brightening effects, which are indeed as expected. 2.1. Characterization. The structures of the acquired polymers, 2, were confirmed by 1H NMR and IR spectra. According to the 1H NMR spectra (Figure S1), the singlet peak in the region of 5.30−5.31 ppm belonged to the characteristic peak of 5H in the 2(5H)-furanone unit. The disappearance of the signal corresponding to −CC−H (δ = 2.65 and 2.68 ppm) Received: February 19, 2017 Accepted: May 25, 2017 Published: September 7, 2017 5557

DOI: 10.1021/acsomega.7b00196 ACS Omega 2017, 2, 5557−5564

ACS Omega

Article

Scheme 1. Synthetic Route of Polymeric Products 2

showed that the Click polymerization between 1 and DASS was realized successfully.33 Furthermore, the diagnostic 1,2,3-triazole proton resonance for samples 2 was observed as a singlet at 7.39 ppm (Figure S1).12 As shown in Figure S2, the stretching vibration bands of C− H (3273 cm−1) and N3 (2091 cm−1) from the monomers became much weaker after polymerization. This indicates that most of the acetylene and azide functional groups from monomers 1 and DASS were expended via the Click polymerization and changed into triazole rings in polymer 2.34 At the same time, the N−H groups exhibited stretching absorption in the range of 3410− 3432 cm−1, whereas the unsaturated groups (C−H) from these aromatic and 1,2,3-triazole rings displayed stretching absorption in the 3074−3153 cm−1 range.35 The intensive stretching band (CO) was located in the range of 1747−1754 cm−1, with the most intensive stretching band (CC) appearing in the range of 1639−1654 cm−1.36 In addition, the skeletal stretching vibrations of aromatic rings were absorbed in the region of 1490−1561 cm−1. Also, the intensive stretching band (SO) was located in range 1182−1197 cm−1,37 and the rocking vibration from the unsaturated C−H bond in the trans-olefin moiety was located in the range 947−968 cm−1.31 In addition, the Mn (number average molecular weight) of polymers 2a−e appeared in the region of 1.21−9.31 kDa and their polydispersity indexes (PDIs) varied between 1.18 and 1.25 (Figure 1), reflecting the formation of a conjugated molecular skeleton with a high molecular weight. The corresponding data are summarized in Table 1. Indeed, the best polymerization result of an Mn of 9.31 kDa and a 98.7% yield (Table 1, polymer 2c) was achieved when the molar ratio of monomer 1 and DASS was 1.1:1.2. An increase in the concentration of either monomer

Table 1. Different Ratios, Molecular Weights, and Solubilities of Polymers 2a−e polymer

1/DASS molar ratio (nominal)

1/DASS molar ratio (XPS)

Mna (kDa)

PDIa

yield (%)

solubilityb

2a 2b 2c 2d 2e

5:1 2:1 1:1 1:2 1:5

5.1:0.9 2.2:1.3 1.1:1.2 1.2:2.3 1.1:5.2

1.21 5.52 9.31 7.62 3.47

1.25 1.19 1.16 1.18 1.21

95.1 97.2 98.7 96.4 95.6

T, S, F T, S, F T, S, F T, S, F T, S, F

a

Determined by GPC (THF eluent, calibrated using polystyrene standards). bSolubility tested in THF, DMSO, and DMF; T = THF, S = DMSO, F = DMF.

1 or DASS lowered the Mn, as higher monomer concentration could the hinder intermolecular monomer collisions, thereby lowering the polymerization results.38 All of the polymers exhibited good solubilities in organic solvents, including tetrahydrofuran (THF), dimethyl sulfoxide (DMSO), and dimethylformamide (DMF), as listed in Table 1, possibly due to the incorporation of alkyl chains from menthol and sodium salt.39,40 2.2. Thermal Properties. Generally, polymers containing conjugated or cross-linked structures tend to be thermally stable and are also good candidates for textile material.41 Thermal properties of polymeric products 2a−e were determined by thermogravimetric analysis and differential scanning calorimetry (DSC), and the data are summarized in Table 2. Figure 2 displays Table 2. Summary of Thermal and Spectroscopic Properties polymer

T5% (°C)a

Tg (°C)b

λex (nm)

λem (nm)

Φ (%)c

2a 2b 2c 2d 2e

207 288 315 267 242

126 171 235 154 137

360 370 397 392 378

450 475 493 468 457

45.1 80.9 92.5 72.4 66.8

a

Temperature losing 5% weight were studied. bGlass-transition temperature was evaluated via DSC. cΦ (solid fluorescence quantum yields) values were measured through the calibrated optimum.

the thermal degradation of these polymers when these were gradually heated from 50 to 700 °C. The degradation temperatures (Td) at the point of 5% weight loss were located in the region of 207−315 °C. Noticeably, polymers 2a−e retained more than 50% of their weights at temperatures over 700 °C. It was implied that they have a much better thermal stability than many commercial polymeric materials.42 As revealed by the high glass-transition temperatures (Tg, 126− 235 °C), all of the samples showed high morphological stability (Figure 3).31 Among them, 2c displayed the maximum stability

Figure 1. GPC traces of the obtained main chain poly-1,2,3-triazoles 2a−e. 5558

DOI: 10.1021/acsomega.7b00196 ACS Omega 2017, 2, 5557−5564

ACS Omega

Article

Figure 2. Thermogravimetric analysis of polymers 2a−e collected under a N2 atmosphere.

Figure 4. Normalized UV−vis (a) and PL (b) spectra of samples 2a−e.

pH values, and then, the pH was regulated with HCl or NaOH.45 Figure S3 shows that their emission intensity decreased with a decrease in the pH value because protonation of the 1,2,3-triazole moiety and sulfonic groups by a strong acid mainly narrowed the quenching process and resulted in enhanced overall intensity.46 As shown in Figure S3, in acidic solutions of all specimens, obvious redshifts in the emitting maximum in the range of 478− 512 nm occurred at pH 1. These redshifts in the acid system were attributed to the changeable highest occupied molecular orbital (HOMO)−lowest unoccupied molecular orbital (LUMO) gaps, which disrupt the ordered molecular packing.47 In addition, a larger gap between HOMO and LUMO ascribed to aromatic hydrocarbon and aliphatic amine interactions also influenced the electron density in the OBA system.48 In contrast, all of them exhibited mildly reduced emission intensities in alkaline solutions than in acidic solutions (Figure S3). The emitting bands of polymers were blue-shifted in the range from 428 to 466 nm. The shift may be arising from the disruption between the 2(5H)-furanone moiety and the amino acid ester in the system.15 Moreover, these hydrolyses will not only induce the trans−cis conversion (Scheme 2) more easily but also cause cleavage of the polymer structure, even influencing the fluorescence property and leading to blueshifts.48 2.5. XPS Analysis. Information on the electronic structure and chemical composition of polymer 2c was obtained by XPS analysis,49 as given in Table 1. A survey of the spectra of 2c shows that peaks are located at 929.4, 573.8, 532.3, 400.5, 289.6, and 163.7 eV, which are (Br 2s), (O 1s), (N 1s), (C 1s), and (S 2p), respectively, implying the high purity of the prepared samples.50

Figure 3. DSC curves of polymers 2a−e recorded under a N2 atmosphere.

(235 °C), as shown in Table 2. Consequently, the Tg decreased while the monomer ratio changed from 1.1:1.2. This may imply that interactions between side chains or the side chain and main chain reduced the Tg values in cases in which the speed of the Click polymerization slowed due to steric retardation with the azide moiety. Moreover, CuI-catalyzed Glaser homocoupling of the acetylene moiety to generate diacetylene could turn into a side reaction.43,44 2.3. UV−Visible (Vis) and Fluorescence Spectra. The UV−vis and luminance spectra of polymeric products 2a−e in aqueous solution are shown in Figure 4. Their UV−vis absorption spectra displayed two bands at about 263−271 and 360−397 nm (Figure 4a). Upon light irradiation (λ = 360−397 nm), the emission maxima were shown at about 388−440 nm (Figure 4b) for those specimens separately. A similar trend was observed for polymers 2a−e, whose quantum yields (Φ) are given in Table 2. 2.4. Influences of pH Values. For the purpose of investigating the acid-fastness of polymers 2a−e, the fluorescence response of these samples were evaluated at five different 5559

DOI: 10.1021/acsomega.7b00196 ACS Omega 2017, 2, 5557−5564

ACS Omega

Article

2.7. Evaluation of UPF. Aimed at studying the anti-UV ability of untreated cotton clothes and those deposited with 2a− e, UPFs were determined and are given in Table 3.55 More UV rays could pass through the original fabric, leading to a relatively high transmittance, with a corresponding UPF value of only 6.032. After treatment with polymers 2a−e, the transmittance of the samples decreased, with an obvious enhancement of UPF. Apparently, as the 2a−e concentrations increases, the UPF of the cotton fabric increased. When the concentration increased to 0.2% (o.w.f.), the UPF was over 20. It was shown that these polymeric products are extraordinarily suitable for protecting against harmful UV rays.56 Among them, 2c offered the maximum protection, with a UPF value that far exceeded that of pure fabric. Furthermore, when 0.5% (o.w.f.) 2c was added to the cotton fabric, the UPF increased to more than 51 (Table 3). Significantly, the UPF value in this case was larger than that in the previous study (UPF = 42.358),15 which accounted for OBAs’ main chain with multichiral units, thus restraining the trans-to-cis conversion in 2c.55 Afterward, the UPF value reduced with an excessive amount of FBAs. The reason for this phenomenon is that the light filtered by the excess FBAs, the high agglomeration of FBAs may block the active sites on the surface of the polymers.30 These results are also in agreement with those of previous photoisomerization tests. 2.8. Water-Laundering Test. After demonstrating the outstanding UV protection ability of 2c, we also measured the water durability of fabrics deposited with polymer 2c.57 The additional water-laundering durability test was carried out for 10 min without any recovery treatment. As illustrated in Figure 6, the fabric showed significant stability and the UPF value did not decrease significantly even after eight cycles of water laundering. Subsequently, the UPF of specimen 2c reduced from 51.553 to 50.492 after 10 cycles, and the UV (UV-A and UV-B) transmittance increased slightly by nearly 1%. However, the water-laundering durability of previous bis-1,2,3-triazole15 maintained for just three cycles (Figure S5) and decreased quickly afterward. Eventually, the UPF value of the small molecule dropped to 2.925 after 10 cycles of water laundering. Obviously, the water-laundering durability of polymer 2c is far higher than that of the previous bis-1,2,3-triazole molecule. Therefore, it can be concluded that the introduction of poly1,2,3-triazole can not only produce polymer 2c-consolidated cotton fabrics with ultrastrong UV-defending ability but also provide extraordinary water-laundering durability.58,59

Scheme 2. Photoisomerization of Polymer 2

Figure S4 shows that the C (1s) spectra were consistent with carbon−nitrogen (285.5 eV) and carbon−sulfur (286.7 eV) environments. Meanwhile, the peak binding energies of the S (2p3/2) and S (2p1/2) components were consistent with C− SO3H (sulfonate) environments and the signal at 398.2 eV corresponding to NN−C sp2-hybridized N atoms in the triazole ring (Figure S4)51 demonstrated a successful Click polymerization between monomer 1 and DASS. These results matched with other characterization information, strongly suggesting the model polymeric products. 2.6. Photoisomerization. The polymeric products having stilbene-type chromophores have cis and trans isomers. Generally, the trans state of the stilbene molecule is more stable than the cis state.3 As shown in Figure 5, the UV−vis spectra of multiblock copolymers 2a−e were measured in THF solution. According to Figure 5, their most intense absorption bands were below 280 nm in the ultraviolet region, arising from the π−π* transition localized on the 2(5H)-furanone ring within each polymer. The next weaker broad band, which appeared at approximately 350 nm, can be ascribed to the trans-stilbene sulfonate moiety.4 The UV−vis spectrum of these samples at UV 365 nm light irradiation exhibited a decrease in the high-intensity band at 350 nm along with a slight bathochromic shift to 340 nm owing to the trans-to-cis photoisomerization.52 Importantly, although a balance can be reached between cis and trans isomers of polymers 2a−e, the time of achieving their respective photostationary state is distinctive.53 Among these, polymer 2c reached a photostationary state in 240 min with irradiation at 365 nm. Notably, the time far exceeded that for prebis-1,2,3-triazole (80 min)15 owing to OBA’s main chain with multichiral units restraining the trans-to-cis conversion for 2c.53 In other words, the trans-to-cis photoisomerization rate of 2c was at least 10 times higher than that of other samples, even that of stilbene sulfate itself. The main reason was that both monomer 1 and DASS in the 2c system had undergone complete Click polymerization. However, further increase in DASS content led to a significant decrease in photoability, which may be attributed to an increase in the absorbance and scattering of photons due to excess DASS in the polymer system.54

3. CONCLUSIONS A simple and effective method was successfully developed to prepare poly-1,2,3-triazoles by one-pot Click polymerization of diyne and disulfonyl azide. These polymerizations utilize the advantages of Click reactions, such as outstanding efficiency, mild reaction, inexpensive catalyst, and good atom economy. The obtained polymers possess high UPF values as well as good thermal stability. In particular, when the diynes/disulfonyl azide ratio was 1.1/1.2, the highest Mn of copolymer was 9.31 kDa. With FBA moieties, the polymers display fluorescence-brightening characteristics. Moreover, their strong photostationary state features make them promising materials in textile, pulp, and paper making. 4. EXPERIMENTAL SECTION 4.1. Materials and Equipment. Among the solvent and reagents (analytical grade), sodium 6,6′-(ethene-1,2-diyl)bis(35560

DOI: 10.1021/acsomega.7b00196 ACS Omega 2017, 2, 5557−5564

ACS Omega

Article

Figure 5. UV−vis spectral changes of polymers 2a−e in THF upon UV light irradiation as a function of time: (a) irradiation time = 60 min; (b) irradiation time = 180 min; (c) irradiation time = 300 min; (d) irradiation time = 120 min; (e) irradiation time = 90 min.

Table 3. UPF Values of Fabric Samples 2a−e and Untreated Cotton Fabric amount applied; o.w.f. (%) sample 2a 2b 2c 2d 2e untreated cotton fabric

0.1

0.2

0.3

0.4

0.5

15.098 17.191 20.115 19.667 18.032 6.03215

22.127 25.183 29.224 27.015 26.143

35.154 37.368 40.147 38.021 36.258

40.289 42.104 47.096 45.668 44.099

45.319 48.681 51.553 49.006 46.421

azidobenzenesulfonate) (DASS, ≥98.0%) was acquired from TCI, whereas intermediate 1, containing 2(5H)-furanone moiety, was synthesized according to the literature.15,27 All of the fusing points were measured by Yuhua X4 fusing point equipment (Gongyi, China) and were uncorrected. IR spectra were recorded on an FT-IR spectrometer (Bruker Tensor 27 Fourier transform infrared). 1H NMR spectra were recorded

Figure 6. Water-laundering test of sample 2c-deposited fabric (10 cycles).

in DMSO-d6 on a Bruker DRX-400 MHz spectrometer, and tetramethylsilane served as an internal standard. 5561

DOI: 10.1021/acsomega.7b00196 ACS Omega 2017, 2, 5557−5564

ACS Omega

Article

Author Contributions

Both Mn and the corresponding molecular weight distribution for polymeric products 2a−e were measured by GPC with THF as the solvent under the testing conditions given below. Polystyrene worked as the reference material at 35 °C, with a flow velocity (1 mL min−1). DSC was determined by thermal analyzer (Perkin-Elmer DSC7, Norwalk, CT). The thermal stability of polymers 2a−e were determined on a thermogravimetric analyzer using NETZSCH STA 409 PC to ensure that the samples were stable at high temperatures of up to 700 °C.28 The chemical compositions of these specimens were analyzed by XPS (Axis Ultra DLD, Kratos) equipment. The UV absorption peaks were measured using an ultraviolet absorption detector (Shimazu UV-2550). The luminance spectrum was obtained using a Hitachi F-2500 spectrophotometer. UPF values was determined using a UV−vis spectrophotometer (Shimazu UV-2550) equipped with an optimum and a fabric-supporting device.15,29 The water-laundering durability of the fabric was measured by the AATCC Test Method 61-2006. The size of the fabric was 50−150 mm, and the measuring environments were as follows: the specimens were laundered in a closed canister (200 mL) with WOB (AATCC Standard, 0.37% w/w) for 45 min at a heating temperature of 40 ± 5 °C. Subsequently, the UPF of the fabric specimen was measured again under various washing times, simulating from 10 to 100 wash processes, that is, between 1st and 10th cycle.30 UV−vis spectra were recorded on an absorption spectrophotometer (Hitachi U-3010). Luminance spectra were recorded on a fluorescence spectrophotometer (Hitachi F-4500). In addition, the solid luminance quantum yields (ϕPL) were obtain using calibrated integrating sphere equipment.29 4.2. Synthesis. All Click polymerizations were preceded by a standard Schlenk technique under nitrogen. First, monomer 1 and DASS were blended in an anticipative molar feed ratio (Table 1) in a Schlenk tube (10 mL) equipped with a magnetic stirrer. Subsequently, CuI (0.02 mmol), DMF (0.5 mL), and Et3N (0.04 mmol) were added. After stirring at ambient temperature for 1 h, the copper salts were laundered and the crude product was obtained through precipitation in methanol. The mixture was further purified in a Soxhlet extractor with methanol overnight, followed by drying in vacuum overnight.31,32





Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS



REFERENCES

We are very thankful to the Natural Science Program of China (No. 21671038), the science and technology innovation special fund project of Foshan (2014AG10009), and high-level talent start-up research project of Foshan University (Gg040947, Gg040919, Gg040918).

(1) Osberger, T. J.; Rogness, D. C.; Kohrt, J. T.; Stepan, A. F.; White, M. C. Oxidative diversification of amino acids and peptides by smallmolecule iron catalysis. Nature 2016, 537, 214−219. (2) Gibson, B. A.; Zhang, Y. J.; Jiang, H.; Hussey, K. M.; Shrimp, J. H.; Lin, H. N.; Schwede, F.; Yu, Y. H.; Kraus, W. L. Chemical genetic discovery of PARP targets reveals a role for PARP-1 in transcription elongation. Science 2016, 353, 45−50. (3) Guerra, S.; Nguyen, T. L. A.; Furrer, J.; Nierengarten, J.-F.; Barberá, J.; Deschenaux, R. Liquid-crystalline dendrimers designed by click chemistry. Macromolecules 2016, 49, 3222−3231. (4) Wang, J.-T.; Saito, K.; Wu, H.-C.; Sun, H.-S.; Hung, C.-C.; Chen, Y.; Isono, T.; Kakuchi, T.; Satoh, T.; Chen, W.-C. High-performance stretchable resistive memories using donor-acceptor block copolymers with fluorene rods and pendent isoindigo coils. NPG Asia Mater. 2016, 8, No. e298. (5) Diehl, K. L.; Kolesnichenko, I. V.; Robotham, S. A.; Bachman, J. L.; Zhong, Y.; Brodbelt, J. S.; Anslyn, E. V. Click and chemically triggered declick reactions through reversible amine and thiol coupling via a conjugate acceptor. Nat. Chem. 2016, 8, 968−973. (6) Silva, L. N.; Zimmer, K. R.; Macedo, A. J.; Trentin, D. S. Plant natural products targeting bacterial virulence factors. Chem. Rev. 2016, 116, 9162−9236. (7) Xu, H.; Tao, S. S.; Jiang, D. L. Proton conduction in crystalline and porous covalent organic frameworks. Nat. Mater. 2016, 15, 722−726. (8) Taskin, O. S.; Yilmaz, G.; Yagci, Y. Fullerene-attached polymeric homogeneous/heterogeneous photoactivators for visible-light-induced CuAAC click reactions. ACS Macro Lett. 2016, 5, 103−107. (9) Zuo, X. L.; Morlet-Savary, F.; Graff, B.; Blanchard, N.; Goddard, J.P.; Lalevée, J. Fluorescent brighteners as visible LED-light sensitive photoinitiators for free radical photopolymerizations. Macromol. Rapid Commun. 2016, 37, 840−844. (10) Ventura, B.; Poronik, Y. M.; Deperasińska, I.; Gryko, D. T. How a small structural difference can turn optical properties of π-extended coumarins upside down: The role of non-innocent saturated rings. Chem. Eur. J. 2016, 22, 15380−15388. (11) Li, M. W.; Tian, R.; Yan, D. P.; Liang, R. Z.; Wei, M.; Evans, D. G.; et al. A luminescent ultrathin film with reversible sensing toward pressure. Chem. Commun. 2016, 52, 4663−4666. (12) Yilmaz, M. D. Conformation-dependent O-glycosylation of threonine-containing dipeptides. Carbohydr. Polym. 2016, 146, 174− 180. (13) Jaunet-Lahary, T.; Goupille, A.; Jacquemin, D.; Fleury, F.; Graton, J.; Laurent, A. D. A joint theoretical and experimental study of the behavior of the DIDS inhibitor and its derivatives. ChemPhysChem 2016, 17, 2434−2445. (14) Pathak, S. K.; Pradhan, B.; Gupta, M.; Pal, S. K.; Sudhakar, A. A. Liquid-crystalline star-shaped supergelator exhibiting aggregationinduced blue light emission. Langmuir 2016, 32, 9301−9312. (15) Huo, J.-P.; Luo, J.-C.; Wu, W.; Xiong, J.-F.; Mo, G.-Z.; Wang, Z.-Y. Synthesis and characterization of fluorescent brightening agents with chiral 2(5H)-furanone and bis-1,2,3-triazole structure. Ind. Eng. Chem. Res. 2013, 52, 11850−11857.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.7b00196. 1 H NMR spectra of 1, DASS, and 2; IR spectra of 1, DASS, and 2; effect of pH on the fluorescence properties of samples 2a−e; high-resolution XPS spectrum of polymer 2c; water-laundering test of the previously reported bis1,2,3-triazole (10 cycles); references (PDF)



J.H. and Z.H. contributed equally to this work.

AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected]. Tel: 86-757-82966588 (J.H.). *E-mail: [email protected]. Tel: 86-20-39310258 (Z.W.). ORCID

Jingpei Huo: 0000-0002-1875-4375 5562

DOI: 10.1021/acsomega.7b00196 ACS Omega 2017, 2, 5557−5564

ACS Omega

Article

(16) Shi, Y.; Cao, X. S.; Gao, H. F. The use of azide-alkyne click chemistry in recent syntheses and applications of polytriazole-based nanostructured polymers. Nanoscale 2016, 8, 4864−4881. (17) Singh, A.; Banerjee, S.; Komber, H.; Voit, B. Synthesis and characterization of highly fluorinated sulfonated polytriazoles for proton exchange membrane application. RSC Adv. 2016, 6, 13478−13489. (18) Dong, W. H.; Wu, H. Q.; Chen, M.; Shi, Y.; Sun, J. Z.; Qin, A. J.; Tang, B. Z. Anionic conjugated polytriazole: direct preparation, aggregation-enhanced emission, and highly efficient Al3+ sensing. Polym. Chem. 2016, 7, 5835−5839. (19) Warde, U.; Sekar, N. NLOphoric mono-azo dyes with negative solvatochromism and in-built ESIPT unit from ethyl 1,3-dihydroxy-2naphthoate: Estimation of excited state dipole moment and pH study. Dyes Pigm. 2017, 137, 384−394. (20) Freitas, J.; Perestrelo, R.; Cassaca, R.; Castillo, M.; Santos, M.; Pereira, J.; et al. A fast and environment-friendly MEPSPEP/UHPLCPDA methodology to assess 3-hydroxy-4,5-dimethyl-2(5H)-furanone in fortified wines. Food Chem. 2017, 214, 686−693. (21) Simeonov, S. P.; Nunes, J. P. M.; Guerra, K.; Kurteva, V. B.; Afonso, C. A. M. Synthesis of chiral cyclopentenones. Chem. Rev. 2016, 116, 5744−5893. (22) Jacques, R.; Pal, R.; Parker, N. A.; Sear, C. E.; Smith, P. W.; Ribaucourt, A.; Hodgson, D. M. Recent applications in natural product synthesis of dihydrofuran and -pyran formation by ring-closing alkene metathesis. Org. Biomol. Chem. 2016, 14, 5875−5893. (23) Li, X. D.; Lan, X. C.; Wang, T. F. Highly selective catalytic conversion of furfural to γ-butyrolactone. Green Chem. 2016, 18, 638− 642. (24) Huang, S.-Z.; Kong, F.-D.; Ma, Q.-Y.; Guo, Z.-K.; Zhou, L.-M.; Wang, Q.; Dai, H.-F.; Zhao, Y.-X. Nematicidal stemona alkaloids from stemona parviflora. J. Nat. Prod. 2016, 79, 2599−2605. (25) Ma, J.; Yuan, Z.-Z.; Kong, X.-W.; Wang, H.; Li, Y.-M.; Xiao, H.; Zhao, G. Reagent-controlled tandem reactions of vinyl epoxides: Access to functionalized γ-butenolides. Org. Lett. 2016, 18, 1450−1453. (26) Fang, Y.; Liu, W.; Teat, S. J.; Dey, G.; Shen, Z. Q.; An, L. T.; Yu, D. C.; Wang, L.; O’Carroll, D. M.; Li, J. A systematic approach to achieving high performance hybrid lighting phosphors with excellent thermal- and photostability. Adv. Funct. Mater. 2017, 27, No. 1603444. (27) Huo, J. P.; Deng, G. H.; Wu, W.; Xiong, J. F.; Zhong, M. L.; Wang, Z. Y. Electrophoretic deposition polymerization of diacetylenes with tunable structure. Macromol. Rapid Commun. 2013, 34, 1779−1784. (28) Han, G. D.; Park, S. S.; Liu, Y.; Zhitomirsky, D.; Cho, E.; Dincă, M.; Grossman, J. C. Photon energy storage materials with high energy densities based on diacetylene-azobenzene derivatives. J. Mater. Chem. A 2016, 4, 16157−16165. (29) Qian, Y.; Qiu, X. Q.; Zhu, S. P. Sunscreen performance of lignin from different technical resources and their general synergistic effect with synthetic sunscreens. ACS Sustainable Chem. Eng. 2016, 4, 4029− 4035. (30) Yu, M.; Wang, Z. Q.; Lv, M.; Hao, R. Z.; Zhao, R. T.; Qi, L. H.; Liu, S. M.; Yu, C. H.; Zhang, B. W.; Fan, C. H.; Li, J. Y. Antisuperbug cotton fabric with excellent laundering durability. ACS Appl. Mater. Interfaces 2016, 8, 19866−19871. (31) Deng, H. Q.; Han, T.; Zhao, E. G.; Kwok, R. T. K.; Lam, J. W. Y.; Tang, B. Z. Multicomponent click polymerization: A facile strategy toward fused heterocyclic polymers. Macromolecules 2016, 49, 5475− 5483. (32) Wang, D.; Guo, Q. S.; Gao, H.; Yang, Z.; Xing, Y.; Cao, H.; He, W. L.; Wang, H. H.; Gu, J. M.; Hu, H. Y. The application of double click to synthesize a third-order nonlinear polymer containing donor-acceptor chromophores. Polym. Chem. 2016, 7, 3714−3721. (33) Hu, J. Q.; Peng, K. M.; Guo, J. S.; Shan, D. Y.; Kim, G. B.; Li, Q. Y.; Gerhard, E.; Zhu, L.; Tu, W. P.; Lv, W. Z.; Hickner, M. A.; Yang, J. Click cross-linking-improved waterborne polymers for environment-friendly coatings and adhesives. ACS Appl. Mater. Interfaces 2016, 8, 17499− 17510. (34) Reshmi, S.; Hemanth, H.; Gayathria, S.; Naira, C. P. R. Polyether triazoles: An effective binder for ‘green’ gas generator solid propellants. Polymer 2016, 92, 201−209.

(35) Yang, H.; van de Ven, T. G. M. A bottom-up route to a chemically end-to-end assembly of nanocellulose fibers. Biomacromolecules 2016, 17, 2240−2247. (36) Brooks, K.; Yatvin, J.; McNitt, C. D.; Reese, R. A.; Jung, C.; Popik, V. V.; Locklin, J. Multifunctional surface manipulation using orthogonal click chemistry. Langmuir 2016, 32, 6600−6605. (37) Li, H. Q.; Liu, X. J.; Xu, J. M.; Xu, D.; Ni, H. Z.; Wang, S.; Wang, Z. Enhanced proton conductivity of sulfonated poly(arylene ether ketone sulfone) for fuel cells by grafting triazole groups onto polymer chains. J. Membrane Sci. 2016, 509, 173−181. (38) Shi, Y.; Cao, X. S.; Luo, S. J.; Wang, X. F.; Graff, R. W.; Hu, D. Q.; Guo, R. L.; Gao, H. F. Investigate the glass transition temperature of hyperbranched copolymers with segmented monomer sequence. Macromolecules 2016, 49, 4416−4422. (39) Roose, J.; Leung, A. C. S.; Wang, J.; Peng, Q.; Sung, H. H.-Y.; Williams, I. D.; Tang, B. Z. A colour-tunable chiral AIEgen: reversible coordination, enantiomer discrimination and morphology visualization. Chem. Sci. 2016, 7, 6106−6114. (40) Takahashi, Y.; Kishimoto, M.; Kondo, Y. Photoinduced formation of threadlike micelles from mixtures of a cationic surfactant and a stilbene amphiphile. J. Colloid Interface Sci. 2016, 470, 250−256. (41) Bonet-Aracil, M. A.; Díaz-García, P.; Bou-Belda, E.; Sebasti, N.; Montoro, A.; Rodrigo, R. UV protection from cotton fabrics dyed with different tea extracts. Dyes Pigm. 2016, 134, 448−452. (42) Song, S.; Ko, Y.-G.; Lee, H.; Wi, D.; Ree, B. J.; Li, Y. R.; Michinobu, T.; Ree, M. High-performance triazole-containing brush polymers via azide-alkyne click chemistry: a new functional polymer platform for electrical memory devices. NPG Asia Mater. 2015, 7, No. e228. (43) Castro, V.; Rodríguez, H.; Albericio, F. CuAAC: An efficient click chemistry reaction on solid phase. ACS Comb. Sci. 2016, 18, 1−14. (44) Macdonell, A.; Johnson, N. A. B.; Surman, A. J.; Cronin, L. Configurable nanosized metal oxide oligomers via precise “Click” coupling control of hybrid polyoxometalates. J. Am. Chem. Soc. 2015, 137, 5662−5665. (45) Thivaios, I.; Kakogianni, S.; Bokias, G. A library of quinolinelabeled water-soluble copolymers with pH-tunable fluorescence response in the acidic pH region. Macromolecules 2016, 49, 3526−3534. (46) Wu, L. L.; Li, X. L.; Huang, C. S.; Jia, N. Q. Dual-modal colorimetric/fluorescence molecular probe for ratiometric sensing of pH and its application. Anal. Chem. 2016, 88, 8332−8338. (47) Domínguez, S. E.; Meriläinen, M.; Ä ar̈ italo, T.; Damlina, P.; Kvarnström, C. Effect of alkoxy-spacer length and solvent on diluted solutions of cationic isothiouronium polythiophenes. RSC Adv. 2017, 7, 7648−7657. (48) Xie, T.; Jing, C.; Li, M.; Ma, W.; Ding, Z.; Long, Y.-T. pHresponse mechanism of a redox reaction between silver ions and hydroquinone. J. Phys. Chem. C 2016, 120, 23104−23110. (49) Obadia, M. M.; Mudraboyina, B. P.; Serghei, A.; Montarnal, D.; Drockenmuller, E. Reprocessing and recycling of highly cross-linked ion-conducting networks through transalkylation exchanges of C-N bonds. J. Am. Chem. Soc. 2015, 137, 6078−6083. (50) Fraser, R. C.; Carletto, A.; Wilson, M.; Badyal, J. P. S. Plasmachemical double click thiol-ene reactions for wet electrical barrier. ACS Appl. Mater. Interfaces 2016, 8, 21832−21838. (51) Dontsova, D.; Fettkenhauer, C.; Papaefthimiou, V.; Schmidt, J.; Antonietti, M. 1,2,4-Triazole-based approach to noble-metal-free visible-light driven water splitting over carbon nitrides. Chem. Mater. 2016, 28, 772−778. (52) Zhou, J. W.; Guo, X.; Katz, H. E.; Bragg, A. E. Molecular switching via multiplicity-exclusive E/Z photoisomerization pathways. J. Am. Chem. Soc. 2015, 137, 10841−10850. (53) Lin, L.-R.; Wang, X.; Wei, G.-N.; Tang, H.-H.; Zhang, H.; Ma, L.H. Azobenzene-derived tris-β-diketonate lanthanide complexes: reversible trans-to-cis photoisomerization in solution and solid state. Dalton Trans. 2016, 45, 14954−14964. (54) Hu, X.-S.; Shen, Y.; Xu, L.-H.; Wang, L.-M.; Xing, Y.-J. Preparation of flower-like CuS by solvothermal method and its 5563

DOI: 10.1021/acsomega.7b00196 ACS Omega 2017, 2, 5557−5564

ACS Omega

Article

photodegradation and UV protection. J. Alloys Compd. 2016, 674, 289− 294. (55) Coughlan, N. J. A.; Wallace, C. M.; Adamson, B. D.; Bieske, E. J. Photoisomerization of β-ionone protonated schiff base in the gas phase. J. Phys. Chem. A 2016, 120, 6557−6562. (56) Abdou, L. A. W.; El-Moll, M. M.; Hakeim, O. A.; El-Gammal, M. S.; Shamey, R. Synthesis of nanoscale binders through mini emulsion polymerization for textile pigment applications. Ind. Eng. Chem. Res. 2013, 52, 2195−2200. (57) Tian, M. W.; Hu, X. L.; Qu, L. J.; Du, M. Z.; Zhu, S. F.; Sun, Y. F.; Han, G. T. Ultraviolet protection cotton fabric achieved via layer-bylayer self-assembly of graphene oxide and chitosan. Appl. Surf. Sci. 2016, 377, 141−148. (58) Babaahmadi, V.; Montazer, M. Reduced graphene oxide/SnO2 nanocomposite on PET surface: Synthesis, characterization and application as an electro-conductive and ultraviolet blocking textile. Colloids Surf., A 2016, 506, 507−513. (59) Huo, J. P.; Zeng, H. P. Silver nanoparticles-sensitized cobalt complex for highly-efficient photocatalytic activity. Appl. Catal., B 2016, 199, 342−349 2016.

5564

DOI: 10.1021/acsomega.7b00196 ACS Omega 2017, 2, 5557−5564