Poly(ethylene terephthalate) Fibers with a Thin Layer of...

7
www.advmatinterfaces.de COMMUNICATION 1800628 (1 of 7) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Poly(ethylene terephthalate) Fibers with a Thin Layer of Click-Based Microporous Organic Network: Enhanced Capture Performance toward PM 2.5 Chang Wan Kang, Yoon-Joo Ko, Sang Moon Lee, Hae Jin Kim, and Seung Uk Son* DOI: 10.1002/admi.201800628 For example, porous polymeric monoliths have been prepared for the removal of PMs. [3] Recently, various membranes con- sisting of fibers have been engineered. [4–7] Especially, electrospinning is a powerful method used to fabricate polymer fibers for porous membranes. [8] For example, poly(ethylene terephthalate) (PET) fibers have been extensively engineered. [9] How- ever, the application of PET membrane for the capture of PMs has rarely been reported, possibly due to its poor adsorp- tion behavior. [4–14] To capture the PM 2.5 efficiently, electrospun fibers have been engineered to small diameters of sub µm, although they can suffer from a permeability issue. [4–7,10–14] As another chemical strategy, the adsorption performance of fibrous membranes can be enhanced through the introduction of additive mate- rials. [10–14] For example, porous materials such as metal–organic frameworks (MOFs) have been incorporated in the engineering of fibers by electrospinning. [10,13] However, the amounts of additive MOF materials in efficient fibrous membranes were quite high up to 50–100 wt%. [10,13] Because the PM 2.5 capture is related to the surface properties of fibrous membranes, thin coating approach may be atom-economically efficient. However, reports on efficient thin coating chemistry for fibrous mem- branes are insufficient. [14–17] Recently, microporous organic networks (MONs) have been synthesized by the coupling of organic building blocks. [18–26] For example, the click reaction of multiazidoarenes with multiethy- nylarenes resulted in MONs through the formation of triazole rings, showing high surface areas and microporosities. [27–33] Although the click-based MONs (C-MONs) have been tested as gas adsorbents, as far as we are aware, there were no reports on the engineering of C-MONs on fibers. We considered that the coating of PET fibers with C-MONs can enhance adsorp- tion performance. In this work, we report the engineering of C-MONs on the surface of PET fibers (PET@C-MON) and their enhanced adsorption performance toward PMs. Figure 1 shows a synthetic scheme for PET@C-MON fibrous membranes. First, metallic copper with a thickness of 200–300 nm was introduced to the surface of PET fibers with 13 µm diameter through electroless deposition. [34] The white PET membrane turned reddish brown through Cu deposi- tion (Figure 2a,b). It is well known that various Cu(I) species including Cu 2 O can catalyze the click reaction of azide–alkyne cycloaddition. [35] Thus, we partially oxidized metallic copper on This work shows that the adsorptive performance of fibrous membrane can be enhanced by microporous organic network (MON) chemistry. Copper species are introduced into the surface of poly(ethylene terephthalate) (PET) fibers through electroless deposition. Through the Cu 2 O-catalyzed click reactions of tetra(4-ethynylphenyl)methane with 1,4-diazidobenzene, click- based MON (C-MON) layers are formed on the PET@Cu@Cu 2 O. Etching of copper species results in the formation of PET@C-MON membranes that show enhanced removal performance by up to 13.3 times for particulate matters with smaller sizes than 2.5 µm (PM 2.5 ), compared with pristine PET fibers. C. W. Kang, Prof. S. U. Son Department of Chemistry Sungkyunkwan University Suwon 16419, Korea E-mail: [email protected] Dr. Y.-J. Ko Laboratory of Nuclear Magnetic Resonance The National Center for Inter-University Research Facilities (NCIRF) Seoul National University Seoul 08826, Korea Dr. S. M. Lee, Dr. H. J. Kim Korea Basic Science Institute Daejeon 34133, Korea The ORCID identification number(s) for the author(s) of this article can be found under https://doi.org/10.1002/admi.201800628. Microporous Network As the population and scale of industry have increased with the development of society, particulate matters (PMs) have caused serious problems to public health. [1,2] PMs can be generated not only on a large scale from factories and transportation but also on a small scale from cigarette smoking. Depending on their size, PMs can be classified into two groups: PM 10 with sizes of 2.5–10 µm and PM 2.5 with sizes smaller than 2.5 µm. [1] Due to their significantly small size, PMs can be suspended in air and adsorbed into the lungs through inhalation, resulting in serious diseases. Although both PM 10 and PM 2.5 are dangerous, PM 2.5 is raising more serious problems due to its relatively long life-time of suspension in air. While there have been con- tinuous accidents related to PMs over the world, today PMs are becoming an increasingly more serious problem in East Asia. [1,2] To address this problem, new chemical strategies need to be explored to develop more efficient adsorbent materials. Adv. Mater. Interfaces 2018, 1800628

Transcript of Poly(ethylene terephthalate) Fibers with a Thin Layer of...

Page 1: Poly(ethylene terephthalate) Fibers with a Thin Layer of ...home.skku.ac.kr/~sson/homepage/board/upload/sub3_1/... · (Figure 3c). While the analysis of the N 2 sorption isotherm

www.advmatinterfaces.de

COMMUNICATION

1800628 (1 of 7) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Poly(ethylene terephthalate) Fibers with a Thin Layer of Click-Based Microporous Organic Network: Enhanced Capture Performance toward PM2.5

Chang Wan Kang, Yoon-Joo Ko, Sang Moon Lee, Hae Jin Kim, and Seung Uk Son*

DOI: 10.1002/admi.201800628

For example, porous polymeric monoliths have been prepared for the removal of PMs.[3]

Recently, various membranes con-sisting of fibers have been engineered.[4–7] Especially, electrospinning is a powerful method used to fabricate polymer fibers for porous membranes.[8] For example, poly(ethylene terephthalate) (PET) fibers have been extensively engineered.[9] How-ever, the application of PET membrane for the capture of PMs has rarely been reported, possibly due to its poor adsorp-tion behavior.[4–14] To capture the PM2.5 efficiently, electrospun fibers have been

engineered to small diameters of sub µm, although they can suffer from a permeability issue.[4–7,10–14] As another chemical strategy, the adsorption performance of fibrous membranes can be enhanced through the introduction of additive mate-rials.[10–14] For example, porous materials such as metal–organic frameworks (MOFs) have been incorporated in the engineering of fibers by electrospinning.[10,13] However, the amounts of additive MOF materials in efficient fibrous membranes were quite high up to ≈50–100 wt%.[10,13] Because the PM2.5 capture is related to the surface properties of fibrous membranes, thin coating approach may be atom-economically efficient. However, reports on efficient thin coating chemistry for fibrous mem-branes are insufficient.[14–17]

Recently, microporous organic networks (MONs) have been synthesized by the coupling of organic building blocks.[18–26] For example, the click reaction of multiazidoarenes with multiethy-nylarenes resulted in MONs through the formation of triazole rings, showing high surface areas and microporosities.[27–33] Although the click-based MONs (C-MONs) have been tested as gas adsorbents, as far as we are aware, there were no reports on the engineering of C-MONs on fibers. We considered that the coating of PET fibers with C-MONs can enhance adsorp-tion performance. In this work, we report the engineering of C-MONs on the surface of PET fibers (PET@C-MON) and their enhanced adsorption performance toward PMs.

Figure 1 shows a synthetic scheme for PET@C-MON fibrous membranes. First, metallic copper with a thickness of 200–300 nm was introduced to the surface of PET fibers with ≈13 µm diameter through electroless deposition.[34] The white PET membrane turned reddish brown through Cu deposi-tion (Figure 2a,b). It is well known that various Cu(I) species including Cu2O can catalyze the click reaction of azide–alkyne cycloaddition.[35] Thus, we partially oxidized metallic copper on

This work shows that the adsorptive performance of fibrous membrane can be enhanced by microporous organic network (MON) chemistry. Copper species are introduced into the surface of poly(ethylene terephthalate) (PET) fibers through electroless deposition. Through the Cu2O-catalyzed click reactions of tetra(4-ethynylphenyl)methane with 1,4-diazidobenzene, click-based MON (C-MON) layers are formed on the PET@Cu@Cu2O. Etching of copper species results in the formation of PET@C-MON membranes that show enhanced removal performance by up to 13.3 times for particulate matters with smaller sizes than 2.5 µm (PM2.5), compared with pristine PET fibers.

C. W. Kang, Prof. S. U. SonDepartment of ChemistrySungkyunkwan UniversitySuwon 16419, KoreaE-mail: [email protected]. Y.-J. KoLaboratory of Nuclear Magnetic ResonanceThe National Center for Inter-University Research Facilities (NCIRF)Seoul National UniversitySeoul 08826, KoreaDr. S. M. Lee, Dr. H. J. KimKorea Basic Science InstituteDaejeon 34133, Korea

The ORCID identification number(s) for the author(s) of this article can be found under https://doi.org/10.1002/admi.201800628.

Microporous Network

As the population and scale of industry have increased with the development of society, particulate matters (PMs) have caused serious problems to public health.[1,2] PMs can be generated not only on a large scale from factories and transportation but also on a small scale from cigarette smoking. Depending on their size, PMs can be classified into two groups: PM10 with sizes of 2.5–10 µm and PM2.5 with sizes smaller than 2.5 µm.[1] Due to their significantly small size, PMs can be suspended in air and adsorbed into the lungs through inhalation, resulting in serious diseases. Although both PM10 and PM2.5 are dangerous, PM2.5 is raising more serious problems due to its relatively long life-time of suspension in air. While there have been con-tinuous accidents related to PMs over the world, today PMs are becoming an increasingly more serious problem in East Asia.[1,2] To address this problem, new chemical strategies need to be explored to develop more efficient adsorbent materials.

Adv. Mater. Interfaces 2018, 1800628

Page 2: Poly(ethylene terephthalate) Fibers with a Thin Layer of ...home.skku.ac.kr/~sson/homepage/board/upload/sub3_1/... · (Figure 3c). While the analysis of the N 2 sorption isotherm

www.advancedsciencenews.com

© 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim1800628 (2 of 7)

www.advmatinterfaces.de

the PET fibers to Cu2O by treating PET@Cu with air at 50 °C for one week. The original red brown color of PET@Cu turned dark brown, indicating the formation of Cu2O on the surface (Figure 2c).

Scanning electron microscopy (SEM) indicated that the smooth surface of PET fibers (Figure 2g–i) was uniformly coated with granular copper species of ≈300 nm thickness in PET@Cu@Cu2O (Figure 2j–l and Figure S1 in the Sup-porting Information). Through the click reaction of tetra(4-ethynylphenyl)methane[36] with 2 eq. 1,4-diazidobenzene[37] by the catalytic Cu2O in PET@Cu@Cu2O, C-MON layers were formed uniformly on the surface of Cu2O layers (Figure 2m–o and Figure S2 in the Supporting Information). We scanned the amounts of C-MON in PET@Cu@Cu2O@C-MON by increasing the amount of building blocks to prepare three materials: PET@Cu@Cu2O@C-MON-1, PET@Cu@Cu2O@C-MON-2, and PET@Cu@Cu2O@C-MON-3 prepared with 9.0, 36, and 144 µmol of tetra(4-ethynylphenyl)methane, respectively. After coating of the C-MON, the colors of PET@Cu@Cu2O@C-MON membranes became much darker (Figure 2d–f).

Cu@Cu2O was etched from the PET@Cu@Cu2O@C-MON to form PET@C-MON. As the amount of C-MON increased, the yellow color of PET@C-MON membranes became denser (Figure 2p–r). While the surfaces of PET@C-MON fibers were relatively rough (Figure 2t–v), compared with those of PET fibers (Figure 2g-i), existence of C-MON layers could not be confirmed. Thus, we etched inner PET fibers from PET@C-MON through base-catalyzed hydrolysis to form C-MON tubes (Figure 2s,w–y).[38] In the case of C-MON-1 tubes obtained through the PET etching from PET@C-MON-1, the tubes

mostly became flat due to their thin thickness of ≈30 nm. As the thickness of C-MON tubes increased from ≈70 nm (C-MON-2) to ≈140 nm (C-MON-3), the shapes of tubes became less flat (Figure 2w–y and Figure S3 in the Supporting Information).

The fibers and tube materials were further characterized by various analysis techniques. Powder X-ray diffraction studies (PXRD) of PET fibers showed three broad diffraction peaks at 2θ values of 17.4, 22.5, and 25.7°, matching well with the results in the literature (Figure 3a).[9]

The PXRD pattern of PET@Cu showed the appearance of diffraction peaks at the 2θ values of 43.2, 50.4, and 74.1°, corre-sponding to the (111), (200), and (220) peaks of metallic copper, respectively. The partial oxidation of Cu to Cu2O in PET@Cu@Cu2O was confirmed by the peak at the 2θ value of 36.2°, corresponding to the (111) peak of Cu2O. The C-MON tubes and C-MON materials in PET@C-MON were amorphous, matching well with the conventional properties of MON mate-rials prepared by the click reaction in the literature (Figure 3a and Figure S4 in the Supporting Information).[27–33] Infrared absorption spectroscopy (IR) of PET fibers showed main vibra-tion peaks at 1717, 1241, 1096, and 722 cm−1, corresponding to CO, CCO, OCC, and CH, respectively (Figure 3b).[39] In comparison, the IR spectrum of C-MON-3 tubes was com-pletely different from that of PET fibers and showed the NN vibration peak of triazole rings at 1608 cm−1,[29] matching well with those of the click-based polymers.[37] More-over, the CH or N3 vibration peaks of the terminal alkyne and azide groups in building blocks were absent or very small at ≈3300 and 2100 cm−1, respectively, indicating that the C-MON materials were formed through click-based networking of the building blocks (Figure 3b and Figure S5 in the Supporting

Adv. Mater. Interfaces 2018, 1800628

Figure 1. Synthetic scheme for PET@C-MON.

Page 3: Poly(ethylene terephthalate) Fibers with a Thin Layer of ...home.skku.ac.kr/~sson/homepage/board/upload/sub3_1/... · (Figure 3c). While the analysis of the N 2 sorption isotherm

www.advancedsciencenews.com

© 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim1800628 (3 of 7)

www.advmatinterfaces.de

Adv. Mater. Interfaces 2018, 1800628

Figure 2. Photographs of a) PET, b) PET@Cu, c) PET@Cu@Cu2O, d) PET@Cu@Cu2O@C-MON-1, e) PET@Cu@Cu2O@C-MON-2, and f) PET@Cu@Cu2O@C-MON-3. SEM images of g–i) PET, j–l) PET@Cu@Cu2O, and m–o) PET@Cu@Cu2O@C-MON-2. Photographs of p) PET@C-MON-1, q) PET@C-MON-2, and r) PET@C-MON-3. s) Synthetic scheme of C-MON tubes from PET@C-MON through PET etching. SEM images of t) PET@C-MON-1, u) PET@C-MON-2, v) PET@C-MON-3, w) C-MON-1 tubes, x) C-MON-2 tubes, and y) C-MON-3 tubes.

Page 4: Poly(ethylene terephthalate) Fibers with a Thin Layer of ...home.skku.ac.kr/~sson/homepage/board/upload/sub3_1/... · (Figure 3c). While the analysis of the N 2 sorption isotherm

www.advancedsciencenews.com

© 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim1800628 (4 of 7)

www.advmatinterfaces.de

Information).[29,37] While the IR peaks of C-MON materials in the PET@C-MON were too small to characterize, minor peaks could be observed at 1522 cm−1, corresponding to the CC vibration peak of aromatic groups. This result indicated that the C-MON materials in PET@C-MON are minor in quantity. Through measuring the weights of C-MON tubes, the contents of C-MON were calculated as 1.8, 4.2, and 6.1 wt% for PET@C-MON-1, PET@C-MON-2, and PET@C-MON-3, respectively. Thermogravimetric analysis (TGA) showed that the C-MON-3 tubes and PET are stable up to 320 and 330 °C, respectively (Figure 3c).

While the analysis of the N2 sorption isotherm curves of PET fibers indicated a poor surface area of ≈0.16 m2 g−1, the C-MON-3 tubes showed a high surface area of 926 m2 g−1 and microporosity (pore sizes <2 nm, micropore volume of 0.31 cm3 g−1) (Figure 3d). Solid state 13C nuclear magnetic reso-nance spectroscopy of C-MON-3 tubes showed 13C peaks of benzyl carbon at 64 ppm and aromatic groups at 120, 128, 136, and 147 ppm, matching well with the results in the literature (Figure 3e).[27]

Considering the high surface area and porosity of the C-MON coating in PET@C-MON membranes, we studied

Adv. Mater. Interfaces 2018, 1800628

Figure 3. a) PXRD patterns of PET, PET@Cu, PET@Cu@Cu2O, PET@Cu@Cu2O@C-MON-3, PET@C-MON-3, and C-MON-3 tubes. b) IR absorption spectra and c) TGA curves of PET, PET@C-MON-3, and C-MON-3. d) N2 isotherm curves obtained at 77 K, pore size distribution diagram (based on the DFT method) and e) solid-state 13C NMR spectrum of C-MON-3 tubes.

Page 5: Poly(ethylene terephthalate) Fibers with a Thin Layer of ...home.skku.ac.kr/~sson/homepage/board/upload/sub3_1/... · (Figure 3c). While the analysis of the N 2 sorption isotherm

www.advancedsciencenews.com

© 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim1800628 (5 of 7)

www.advmatinterfaces.de

their adsorptive removal performance toward PM2.5, compared with that of PET fiber mem-brane. (Figure 4).

Through a literature survey, we found that reproducible PM2.5 can be convincingly gen-erated through burning cigarettes (refer to Experimental Section and Figure S6 in the Supporting Information for details).[13,39,40] The PM2.5 was passed through PET@C-MON and PET membranes (Figure 4a). The con-centration changes of PM2.5 after adsorption tests were measured using a PM2.5 counter. According to dynamic light scattering (DLS) studies, the PM2.5 materials had sizes in the range of 0.1–2 µm (Figure 4b).[40] Under our experimental conditions (initial concentra-tion of PM2.5: 464 ± 16 µg m−3, air flow rate: 7.5 L min−1), while the PET membrane showed a removal efficiency (RE) of 6 ± 3%, the RE gradually increased from 27 ± 7% to 68 ± 5 and 80 ± 5% for PET@C-MON-1, PET@C-MON-2, and PET@C-MON-3, respectively (Figure 4c). The PET@C-MON-3 membrane could be reused through removing the adsorbed PM2.5 by simple washing and showed the RE values of 83 ± 2%, 78 ± 4%, 81 ± 4%, and 82 ± 4% in the second, third, fourth, and fifth successive runs, respectively (Figure 4d). When we measured the weight changes of PET and PET@C-MON membranes after adsorp-tion tests, the adsorption trends observed in the PM2.5 counting experiments could be confirmed (Figure S7 in the Supporting Information). The PET and PET@MON-3 membranes showed a nearly same air pres-sure drop of 8 ± 2 and 9 ± 2 Pa, respectively, at a flow rate of 1 L min−1, indicating the thin and porous properties of membranes and the thin coating of MON on the surface of PET fibers.

As shown in the photographs of Figure 4e–l, the increased amount of adsorbed PM2.5 could be observed in the order of PET < PET@C-MON-1 < PET@C-MON-2 < PET@C-MON-3. The SEM image in Figure 4m shows the PM2.5 generated in this work, confirming the size range of 0.1–2 µm. The PET fibers obtained after the adsorp-tion test for PM2.5 showed nearly no change in the SEM image, compared with that of original one (Figure 2g,h and Figure 4n). In comparison, the adsorbed PM2.5 could be observed in the SEM images of the recovered PET@C-MON membranes and increased from PET@C-MON-1 to PET@C-MON-2 and PET@C-MON-3 (Figure 4o–r). The SEM image of PET@C-MON-3 obtained after

Adv. Mater. Interfaces 2018, 1800628

Figure 4. a) Illustration of PM2.5 removal by PET and PET@C-MON membranes. b) DLS spectrum of PM2.5 used in this work. c) Removal performance (average values of five inde-pendent tests at 24 °C and RH 23%) of PM2.5 (464 µg m−3) by PET, PET@C-MON membranes. d) Recyclability of PET@C-MON-3 membrane. Photographs of e,i) PET, f,j) PET@C-MON-1, g,k) PET@C-MON-2, and h,l) PET@C-MON-3 membranes before and after removal of PM2.5. SEM images of m) PM2.5 used in this work, n) PET, o) PET@C-MON-1, p) PET@C-MON-2, and q,r) PET@C-MON-3 obtained after the adsorption tests for PM2.5.

Page 6: Poly(ethylene terephthalate) Fibers with a Thin Layer of ...home.skku.ac.kr/~sson/homepage/board/upload/sub3_1/... · (Figure 3c). While the analysis of the N 2 sorption isotherm

www.advancedsciencenews.com

© 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim1800628 (6 of 7)

www.advmatinterfaces.de

removing the adsorbed PM2.5 by washing was nearly same as the original PET@C-MON-3, indicating chemical stability of PET@C-MON membranes (Figure S8 in the Supporting Infor-mation). These results show that the C-MON thin coating on the PET fibers is critically beneficial in the adsorptive removal of PM2.5. We think that the PMs in this work have various sizes. The relatively smaller sized materials can be easily adsorbed in the C-MON coating and then, larger PMs can coagulate with adsorbed PMs.

Very recently, there have been reports on the enhanced adsorption performance of fibrous adsorbents by adding porous materials such as MOFs.[10–14] Although direct comparison of the adsorption performance of PET@C-MON with those in the literature is limited due to the different experimental condi-tions, it is noteworthy that the contents of MOFs in the litera-ture were in the range of 50–100 wt% for removal efficiencies of 73–89% (PM2.5 concentration of 350–410 µg m−3).[10,13] In contrast, the content of C-MON-3 in PET@C-MON-3 is rela-tively low (6.1 wt%) for removal efficiencies of 80–83% (PM2.5 concentration of 464 µg m−3), indicating the synthetic efficiency of this work for the functionalization of fibers. The removal efficiency of PET@C-MON-3 membrane was enhanced by up to 13.3 times, compared with bare PET membranes.[41] More-over, the unique chemical stability of the C-MON coating could be an advantage in the recycling of PET@C-MON membranes.

In conclusion, this work shows that the C-MON materials can be applied for the thin coating of PET fibers, resulting in changes of the chemical and physical properties of sur-faces. Due to the high surface area and microporosity of C-MON coating, the PET@C-MON membranes showed much enhanced adsorptive removal of PM2.5, compared with the orig-inal PET membrane. We believe that the engineering strategy of this work can be applied for the introduction of C-MON materials to various fibers to enhance their functionalities.

Experimental SectionCharacterization: SEM studies were conducted using a field emission

SEM (FE-SEM) (JSM7100F). PXRD patterns were obtained using a Rigaku MAX-2200. The surface area was measured through the analysis of N2 adsorption–desorption isotherm curves which were obtained at 77 K using a Micromeritics ASAP2020. A pore size distribution diagram was obtained based on the density functional theory method. IR absorption studies were performed using a Bruker VERTEX 70 Fourier transform infrared (FT-IR) spectrometer. Solid state 13C-NMR studies were conducted at CP-TOSS mode using a 500 MHz Bruker ADVANCE II NMR spectrometer at the NCIRF of Seoul National University. A 4 mm magic angle spinning probe was used. The weights of adsorbed PMs were measured using a high-resolution balance of Sartorious CPA225D semimicrobalance (readability of 0.01 mg). DLS studies were conducted using a zetasizer (Malvern Zetasizer Nano ZS90, Malvern Instruments, Worcestershire, UK) with a He–Ne laser (633 nm, 4 mW). Concentration of PM2.5 was measured using a PM analyzer (TES-5321, TES Electrical Electronic Corp.). Pressure drop was measured using a manometer (KIMO Instruments, MP50). TGA curves were obtained under N2 using a Seiko Exstar 7300.

Synthesis of PET@C-MONs: PET fibrous membrane (nonwoven, 18 g m−2, thickness of 0.1 mm, Toray Advanced Materials Korea Inc.) was cut into a square piece with an area of 2.5 cm × 6 cm, washed in acetone (100 mL) with sonication for 10 min, and dried under air. The PET piece was dipped into 1.2 m HCl solution (500 mL), sonicated for 1 h, washed with distilled water three times, washed with acetone once, and

dried at 50 °C in an oven for 3 h. For electroless Cu deposition, three solutions of SnCl2, PdCl2, and CuSO4 were prepared as follows. For the preparation of SnCl2 solution, SnCl2 (3.4 g, 18 mmol) was dissolved in a mixture of distilled water (880 mL) and HCl solution (37%, 20 mL) by sonication. Distilled water was added to make a total volume of 1 L. For preparation of PdCl2 solution, PdCl2 (0.10 g, 0.56 mmol) was dissolved in a mixture of distilled water (880 mL) and HCl solution (37%, 20 mL) by sonication. Distilled water was added to make a total volume of 1 L. For the preparation of CuSO4 solution, CuSO4·5H2O (5.0 g, 20 mmol), potassium sodium tartrate tetrahydrate (25 g, 89 mmol), and NaOH (7.0 g, 0.17 mol) were dissolved in distilled water to make a total volume of 1 L.

For electroless Cu deposition, the PET piece was dipped into SnCl2 solution (50 mL) in a 70 mL vial and sonicated for 10 min. Then, the PET piece was extracted and dipped into PdCl2 solution (50 mL in a 70 mL vial) for 20 min (without sonication). After HCHO solution (36%, 1 mL) was added to CuSO4 solution (70 mL) at 40 °C, the mixture was stirred with a glass rod. The PET piece treated with SnCl2 and PdCl2 solution was dipped into the mixture of CuSO4 and HCHO for 5 min. While we scanned the deposition time of 5, 10, 20, and 30 min, the homogeneity of Cu coating was best with the deposition time of 5 min. The Cu deposited PET membrane was washed with distilled water (200 mL) three times and acetone (100 mL) once and dried under air. For partial oxidation of Cu to Cu2O, the PET@Cu plate was heated at 50 °C under air in an oven for one week to form PET@Cu@Cu2O.

For coating of C-MON, tetra(4-ethynylphenyl)methane (3.8 mg, 9.5 µm)[36] and 1,4-diazidobenzene (3.0 mg, 19 µm)[37] were dissolved in a mixture of dimethyl sulfoxide (DMSO) (45 mL) and distilled water (5 mL). The PET@Cu@Cu2O plate was dipped in the mixture and heated at 85 °C without stirring overnight. After being cooled to room temperature, PET@Cu@Cu2O@C-MON-1 plate was extracted, washed with acetone (100 mL) two times and ethanol (100 mL) once, and dried at 50 °C in an oven for 3 h. For the synthesis of PET@Cu@Cu2O@C-MON-2, the same synthetic method was applied as that applied for PET@Cu@Cu2O@C-MON-1 except using tetra(4-ethynylphenyl)methane (15 mg, 36 µm) and 1,4-diazidobenzene (12 mg, 75 µm). For the synthesis of PET@Cu@Cu2O@C-MON-3, the same synthetic method was applied as that applied for PET@Cu@Cu2O@C-MON-1 except using tetra(4-ethynylphenyl)methane (62 mg, 149 µm) and 1,4-diazidobenzene (48 mg, 300 µm).

For etching Cu/Cu2O, a PET@Cu@Cu2O@C-MON plate was dipped in Cu etchant solution (50 mL, Aldrich Co., Catalogue no. 667528) at 40 °C for 6 h, taken out, washed with acetone (100 mL) two times and ethanol (100 mL) three times, and dried under air to form PET@C-MON.

Synthesis of PET@C-MON Tubes: For etching PET,[38] NaOH (0.30 g, 7.5 mmol) and tetramethylammonium bromide (0.010 g, 0.065 mmol) were dissolved in a mixture of methanol (40 mL) and 1,4-dioxane (20 mL) in a 250 mL round-bottomed flask. The PET@C-MON plate was cut into small pieces (≈0.5 cm × 0.5 cm) and heated at 60 °C in a mixture of NaOH and tetramethylammonium bromide solution for 6 h. The resultant C-MON tubes were retrieved through filtration using a syringe filter, washed with methylene chloride (200 mL), methanol (200 mL), distilled water (200 mL), and acetone (200 mL) three times each, and dried under vacuum at 80 °C overnight.

Procedure of Adsorption Tests: A three-way cork was connected to a 250 mL flask (refer to Figure S6 in the Supporting Information). One out-put was connected to vacuum line. The other out-put was connected to a PM2.5 source through a filter set. The filter set was prepared by holding PET or PET@C-MON membranes (circular shape with a diameter of 1.4 cm) with a syringe filter. From a literature survey,[39,40] it was found that the convincing and reproducible PM2.5 can be generated by burning cigarettes. After the filter part of a commercial cigarette (Marlboro) was removed, the cigarette was used as a PM2.5 source. After evacuating air in a flask for 10 s by a vacuum line, the PM2.5 generated by burning a cigarette was passed through a filter set for 2 s. After additional 10 s, the concentration of PM2.5 was measured. The average air flow rate was 7.5 L min−1. Initial concentration of PM2.5 was 464 ± 16 µg m−3. The removal efficiencies in Figures 4c,d were obtained by

Adv. Mater. Interfaces 2018, 1800628

Page 7: Poly(ethylene terephthalate) Fibers with a Thin Layer of ...home.skku.ac.kr/~sson/homepage/board/upload/sub3_1/... · (Figure 3c). While the analysis of the N 2 sorption isotherm

www.advancedsciencenews.com

© 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim1800628 (7 of 7)

www.advmatinterfaces.de

measuring the concentration of PM2.5 after adsorption. The adsorption tests were conducted at 24 °C and RH 23%. For each membrane, five independent membranes were tested for the adsorption and the results were treated statistically. For the recyclability tests, five PET@MON-3 membranes were tested independently for the successive five recycle tests. After each run, the membrane was washed with acetone (500 mL) at 60 °C, dried, and used for the next run. In addition, the weight changes of membranes were measured using a high-resolution balance (Sartorious CPA225D semimicrobalance with readability of 0.01 mg) (Figure S7 in the Supporting Information).

Supporting InformationSupporting Information is available from the Wiley Online Library or from the author.

AcknowledgementsThis work was supported by the Basic Science Research Program (Grant No. 2016R1E1A1A01941074) through the National Research Foundation of Korea (NRF) funded by the Ministry of Science, ICT, and Future Planning and the grants CAP-15-02-KBSI (R&D Convergence Program) of National Research Council of Science & Technology (NST) of Korea.

Conflict of InterestThe authors declare no conflict of interest.

Keywordsadsorbent, click reaction, fibers, microporous organic polymers, particulate matter

Received: April 24, 2018Revised: May 10, 2018

Published online:

[1] R. Zhang, G. Wang, S. Guo, M. L. Zamora, Q. Ying, Y. Lin, W. Wang, M. Hu, Y. Wang, Chem. Rev. 2015, 115, 3803.

[2] A. Laskin, J. Laskin, S. A. Nizkorodov, Chem. Rev. 2015, 115, 4335.[3] J. Liu, M. Li, P. Wang, K. Liu, Y. Fang, Chem. Eng. J. 2018, 339, 14.[4] A. Lee, J. W. Elam, S. B. Darling, Environ. Sci.: Water Res. Technol.

2016, 2, 17.[5] R. Zhang, C. Liu, P.-C. Hsu, C. Zhang, N. Liu, J. Zhang, H. R. Lee,

Y. Lu, Y. Qiu, S. Chu, Y. Cui, Nano Lett. 2016, 16, 3642.[6] H. Souzandeh, K. S. Johnson, Y. Wang, K. Bhamidipaty,

W.-H. Zhong, ACS Appl. Mater. Interfaces 2016, 8, 20023.[7] Q. Li, Y. Xu, H. Wei, X. Wang, RSC Adv. 2016, 6, 65275.[8] S. S. Ray, S.-S. Chen, C.-W. Li, N. C. Nguyen, H. T. A. Nguyen, RSC

Adv. 2016, 6, 85495.[9] C. Wang, M.-F. Lee, Y.-J. Wu, Macromolecules 2012, 45, 7939.

[10] H. Liang, X. Jiao, C. Li, D. Chen, J. Mater. Chem. A 2018, 6, 334.[11] H. Gao, Y. Yang, O. Akampumuza, J. Hou, H. Zhang, X. Qin,

Environ. Sci.: Nano 2017, 4, 864.

[12] S. Zhang, H. Liu, J. Yu, W. Luo, B. Ding, J. Mater. Chem. A 2016, 4, 6149.

[13] Y. Zhang, S. Yuan, X. Feng, H. Li, J. Zhou, B. Wang, J. Am. Chem. Soc. 2016, 138, 5785.

[14] J. G. Kim, T. J. Choi, J. Y. Chang, Chem. Eng. J. 2016, 306, 242.[15] N. Kang, J. H. Park, M. Jin, N. Park, S. M. Lee, H. J. Kim, J. M. Kim,

S. U. Son, J. Am. Chem. Soc. 2013, 135, 19115.[16] J. Chun, S. Kang, N. Park, E. J. Park, X. Jing, K.-D. Kim, H. O. Seo,

S. M. Lee, H. J. Kim, W. H. Kwon, Y.-K. Park, J. M. Kim, Y. D. Kim, S. U. Son, J. Am. Chem. Soc. 2014, 136, 6786.

[17] C. W. Kang, J. Cho, Y.-J. Ko, S. M. Lee, H. J. Kim, J. P. Kim, S. U. Son, ACS Appl. Mater. Interfaces 2017, 9, 36936.

[18] N. Chaoui, M. Trunk, R. Dawson, J. Schmidt, A. Thomas, Chem. Soc. Rev. 2017, 46, 3302.

[19] L. Tan, B. Tan, Chem. Soc. Rev. 2017, 46, 3322.[20] Y. Xu, S. Jin, H. Xu, A. Nagai, D. Jiang, Chem. Soc. Rev. 2013, 42,

8012.[21] R. Dawson, A. I. Cooper, D. J. Adams, Prog. Polym. Sci. 2012, 37,

530.[22] F. Vilela, K. Zhang, M. Antonietti, Energy Environ. Sci. 2012, 5, 7819.[23] D. Wu, F. Xu, B. Sun, R. Fu, H. He, K. Matyjaszewski, Chem. Rev.

2012, 112, 3959.[24] M. Mastalerz, Angew. Chem., Int. Ed. 2008, 47, 445.[25] C. Weder, Angew. Chem., Int. Ed. 2008, 47, 448.[26] N. B. McKeown, P. M. Budd, Chem. Soc. Rev. 2006, 35, 675.[27] J. Lee, J. Choi, D. Kang, Y. Myung, S. M. Lee, H. J. Kim, Y.-J. Ko,

S.-K. Kim, S. U. Son, ACS Sustainable Chem. Eng. 2018, 6, 3525.[28] D. Kang, J. H. Ko, J. Choi, K. Cho, S. M. Lee, H. J. Kim, Y.-J. Ko,

K. H. Park, S. U. Son, Chem. Commun. 2017, 53, 2598.[29] H. Zhong, C. Liu, Y. Wang, R. Wang, M. Hong, Chem. Sci. 2016, 7,

2188.[30] L. Li, H. Zhao, R. Wang, ACS Catal. 2015, 5, 948.[31] O. Plietzsch, C. I. Schilling, T. Grab, S. L. Grage, A. S. Ulrich,

A. Comotti, P. Sozzani, T. Muller, S. Bräse, New J. Chem. 2011, 35, 1577.

[32] P. Pandey, O. K. Farha, A. M. Spokoyny, C. A. Mirkin, M. G. Kanatzidis, J. T. Hupp, S. T. Nguyen, J. Mater. Chem. 2011, 21, 1700.

[33] J. R. Holst, E. Stöckel, D. J. Adams, A. I. Cooper, Macromolecules 2010, 43, 8531.

[34] Y.-C. Liao, Z.-K. Kao, ACS Appl. Mater. Interfaces 2012, 4, 5109.[35] Y.-H. Tsai, K. Chanda, Y.-T. Chu, C.-Y. Chiu, M. H. Huang, Nanoscale

2014, 6, 8704.[36] S. Yuan, S. Kirklin, B. Dorney, D.-J. Liu, L. Yu, Macromolecules 2009,

42, 1554.[37] Y. Wang, D. Wang, C. Xu, R. Wang, J. Han, S. Feng,

J. Organomet. Chem. 2011, 696, 3000.[38] V. A. Kosmidis, D. S. Achilias, G. P. Karayannidis,

Macromol. Mater. Eng. 2001, 286, 640.[39] I. N. Strain, Q. Wu, A. M. Pourrahimi, M. S. Hedenqvist,

R. T. Olsson, R. L. Andersson, J. Mater. Chem. A 2015, 3, 1632.[40] M. J. Kleeman, J. J. Schauer, G. R. Cass, Environ. Sci. Technol. 1999,

33, 3516.[41] In our experimental conditions (PM2.5: 464 ± 16 µg m−3, air flow

rate: 7.5 L min−1), the commercial car air filter (Fouring Co.) with a thickness of ≈0.34 mm showed a RE of 87%. The air pressure drop was measured as 40 ± 5 Pa at a flow rate of 1 L min−1. In contrast, the PET@MON-3 membrane with a thickness of 0.1 mm showed a RE of 80–83%. The air pressure drop was measured as 9 ± 2 Pa at a flow rate of 1 L min−1.

Adv. Mater. Interfaces 2018, 1800628