Water-Insoluble Hydrophilic Electrospun Fibrous Mat of Cyclodextrin

Nov 29, 2018 - Received 19 October 2018 ... Microsoft cofounder Paul G. Allen, who died from complications of non-Hodgkin's lymphoma in October,...
0 downloads 0 Views 1MB Size
Subscriber access provided by La Trobe University Library

Article

Water-Insoluble Hydrophilic Electrospun Fibrous Mat of Cyclodextrin-Epichlorohydrin Polymer as Highly Effective Sorbent Asli Celebioglu, Fuat Topuz, and Tamer Uyar ACS Appl. Polym. Mater., Just Accepted Manuscript • DOI: 10.1021/acsapm.8b00034 • Publication Date (Web): 29 Nov 2018 Downloaded from http://pubs.acs.org on December 2, 2018

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 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 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.

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 27 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 Polymer Materials

1

Water-Insoluble Hydrophilic Electrospun Fibrous Mat

2

of Cyclodextrin-Epichlorohydrin Polymer

3

as Highly Effective Sorbent

4

Asli Celebioglu, Fuat Topuz and Tamer Uyar*

5

Institute of Materials Science & Nanotechnology, UNAM-National Nanotechnology Research

6

Center, Bilkent University, 06800 Ankara, Turkey

7 8

*

To whom correspondence should be addressed: Prof. T. Uyar ([email protected]) Phone: +90 (312) 290 8987

9 10 11 12 13 14 15 16 17 18

1 ACS Paragon Plus Environment

ACS Applied Polymer Materials 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 27

19

ABSTRACT. The electrospinning of cyclodextrin (CD) functional fibers has emerged as a

20

promising strategy to develop high performance fibers with great promising applications in

21

many fields, including water treatment. Here, we report the fabrication of water-insoluble

22

hydrophilic poly(cyclodextrin-epichlorohydrin) (poly(CD-ECH)) fibers by the electrospinning

23

of an aqueous solution of hydroxypropyl--CD (HP--CD) and ECH, and their cross-linking

24

through heat treatment. The viscosity and time-dependent oscillatory deformation tests

25

revealed that the cross-linking reactions between CD and ECH are highly sensitive to

26

temperature, and main cross-linking takes place after the heat-treatment of the electrospun

27

mat, which was confirmed by FTIR and XPS analyses. The water contact-angle (WCA)

28

measurement showed the hydrophilicity of the poly(CD-ECH) mat, which could maintain its

29

fibrous structure in water. Furthermore, the poly(CD-ECH) fibers were stable in various

30

organic solvents, i.e., acetonitrile, ethanol, methanol, trichloromethane, dimethylformamide

31

and dimethyl sulfoxide. Thermal gravimetric analysis (TGA) showed the cross-linking

32

increased thermal stability of the poly(CD-ECH) fibers compared to the pristine CD, while

33

XRD analysis revealed the amorphous structure of the poly(CD-ECH) fibers. As a proof-of-

34

concept study, the fiber and granule forms of the poly(CD-ECH) were exploited for the

35

scavenging of an organic dye (i.e., phenolphthalein) and a polycyclic aromatic hydrocarbon

36

(PAH) (i.e., phenanthrene) from their aqueous solutions, and rapid removal was observed for

37

the poly(CD-ECH) fibers than its granule form. Given that the hydrophilic nature of the

38

poly(CD-ECH) electrospun mat, this high scavenging performance can be ascribed to the

39

presence of high active CD content in the poly(CD-ECH) fibers, along with their high specific

40

surface area.

41

Keywords.

42

Cyclodextrin,

electrospinning,

nanofibers,

inclusion-complexation,

phenolphthalein, phenanthrene, water treatment

43 2 ACS Paragon Plus Environment

Page 3 of 27 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 Polymer Materials

44

INTRODUCTION

45

Cyclodextrin (CD) is a cyclic oligomer of glucose with a unique torus-like shape with a

46

central hydrophobic cavity interior and a hydrophilic exterior.1 CDs are inherently nontoxic2

47

and edible molecules,3 and have been widely used as solubilization enhancers for poorly

48

soluble drug molecules,4 drug carriers,5-6 food3 and tissue engineering7, and textile industry8.

49

Likewise, materials derived from CDs have taken considerable interest in a wide spectrum of

50

applications, in which each CD molecule can involve in the removal of water

51

micropollutants.9-13 In one example, Alsbaiee et al. reported the synthesis of porous -CD

52

polymers for the removal of various organic micropollutants, i.e., 2, 4-dichlorophenol, 2-

53

naphthol, 1-napthylamine, metolachlor, bisphenol A, bisphenol S, ethinyl oestradiol,

54

propranolol hydrochloride, from water.14 The authors used tetrafluoroterephthalonitrile as a

55

linker for the production of porous CD polymers with a surface area as high as 263 m2 g-1,

56

which allowed rapid removal of organic pollutants. The same research group used

57

decafluorobiphenyl as a cross-linker of -CD moieties under nucleophilic aromatic

58

substitution conditions and exploited the CD networks for the scavenging of environmentally

59

persistent perfluorooctanoic acid (PFOA).15 The materials could sequester PFOA in a few

60

hours from aqueous solutions. Likewise, epichlorohydrin (ECH) cross-linked CD polymers

61

were exploited as sorbents for the removal of various types of pollutants.16-17 The sorption

62

performance of such materials can be enhanced with increasing their surface areas. In this

63

regard, the electrospinning of CD-functional nanofibers has been a focus of interest owing to

64

their high specific surface area (i.e., more active CD content) and highly porous structure,

65

which enable faster and efficient removal of organic micropollutants.

66

The embedment of CD molecules via blending with a polymeric carrier is a simple route to

67

produce CD-functional electrospun fibers. But, the possibility of leaching of CD molecules

68

from the fiber matrix restricts their reuse18 unless they are chemically attached to the fibrous 3 ACS Paragon Plus Environment

ACS Applied Polymer Materials 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 27

69

network19. Further, their embedment within the polymeric matrix can substantially block the

70

CD cavity to able to form inclusion-complexation. Many studies have been reported using this

71

strategy due to its simplicity. In this regard, Zhao et al. reported the preparation of CD-

72

functional fibers using the mixed solution of poly(acrylic acid), citric acid and -CD. The

73

electrospinning and thereafter, the heat-treatment of the resultant fibers led to CD-functional

74

fibers, which were exploited for the removal of methylene blue with a maximum adsorption

75

capacity of 826.45 mg/g.20 -CD was also loaded in the poly(ether sulfone) (PES) nanofibers

76

and implemented for water depollution from hormone estradiol and pesticide chlorpyrifos.21

77

The removal performance of the materials was compared with the CD-free PES fibers, and the

78

results revealed an enhanced sorption capacity with the incorporation of 10 wt.% CD. CD was

79

also used for the decoration of the fiber surface to enhance the sorption performance of fibers.

80

In this regard, Kayaci et al. reported the surface modification of polyester nanofibers with CD

81

motifs and exploited the fibers for the removal of phenanthrene.22 On the other hand, the use

82

of electrospun fibers made by only CDs would be a rational strategy with enhanced

83

performance for sorption applications.23-24 However, such fibers suffer from rapid dissolution

84

in water by reason of their uncross-linked structure.25-26 To address this problem, in our recent

85

study, we have reported the production of electrospun fibers of cross-linked CD polymers

86

using a multivalent carboxylic acid (i.e., 1,2,3,4-butanetetracarboxylic acid) functional linker

87

and their use in the efficient removal of methylene blue from water.27 Even though the

88

presence of their high potential applications, yet, there is not much strategy in the fabrication

89

of water-insoluble electrospun CD fibers. Therefore, there exists an increasing demand

90

towards the production of the insoluble CD fibers while maintaining their hydrophilic

91

structure for the snaring of pollutants from an aqueous environment through inclusion-

92

complexation.

4 ACS Paragon Plus Environment

Page 5 of 27 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 Polymer Materials

93

The proportion of active CD content in the fiber can be explored by means of inclusion-

94

complexation with phenolphthalein through UV-Vis measurements because of its color

95

change from pink to colorless with the formation of lactonoid dianion.28 Moreover, it can be

96

regarded as a water micropollutant due to its adverse health effects on animals. In this regard,

97

Dunnick et al. reported the continuous administration of phenolphthalein to mice over two

98

years using the concentrations of 0, 3 000, 6 000, and 12 000 ppm and observed multiple

99

carcinogenic effects associated with phenolphthalein uptake.29 Furthermore, Artymowicz et

100

al. reported phenolphthalein-induced toxic epidermal necrolysis in humans.30 Their findings

101

suggests that that phenolphthalein should be included in the list of drug-induced toxic

102

epidermal necrolysis (TEN). CD molecules can form ICs with phenolphthalein, and

103

particularly, -CD has shown good complexation with phenolphthalein with a rate constant of

104

3.7 x 10-7 M-1 s-1.31 Thus, the materials derived from -CD molecules would be rational

105

platforms for the removal of phenolphthalein from water. In this regard, poly(CD-ECH) fibers

106

can be used for the removal of phenanthrene, which is one of the most ubiquitous

107

environmental micropollutants. Unlike most polycyclic aromatic hydrocarbons (PAHs), the

108

water solubility of phenanthrene exceeds 1 ppm, and thus, it can be found in water at a larger

109

extent than other PAH molecules. Furthermore, the half-lives of phenanthrene in soil and

110

sediment were reported in the range of 16-126 days,32 and with time, phenanthrene can

111

contaminate water sources. Therefore, timely clean-up of PAH molecules, such as

112

phenanthrene from contaminated sites, particularly from water sources, is crucial. Like

113

phenolphthalein, CD molecules can also form IC with phenanthrene since CD presents a

114

suitable microcavity to able to host phenanthrene molecules.19, 22

115

In this study, poly(CD-ECH) fibers were produced by the electrospinning of CD molecules

116

in the presence of epichlorohydrin (ECH) as the cross-linker. Rheology was used to explore

117

the flow behavior of the electrospinning solution and their cross-linking at 25 and 50 oC over 5 ACS Paragon Plus Environment

ACS Applied Polymer Materials 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 27

118

time. Then, the electrospun fibrous mat was exposed to thermal treatment at 150 oC for the

119

cross-linking of hydroxyl moieties of CD molecules with ECH. The stability of the

120

electrospun poly(CD-ECH) fibers was explored in water and various organic solvents (i.e.,

121

trichloromethane (TCM), dimethylformamide (DMF), dimethyl sulfoxide (DMSO), ethanol

122

(EtOH), methanol (MeOH), and acetonitrile (ACN)). As a proof-of-concept study, the

123

poly(CD-ECH) fibrous mats were used for the removal of phenolphthalein and phenanthrene

124

from aqueous solutions, and their performance was compared with the granule form of the

125

poly(CD-ECH).

126

EXPERIMENTAL SECTION

127

Materials and Methods. Hydroxypropyl--cyclodextrin (HP--CD, Cavasol® W7 HP))

128

was obtained as a gift from Wacker Chemie GmbH, Germany. Acetonitrile (ACN, ≥99.5%),

129

methanol (MeOH, ≥99.8%), ethanol (EtOH, ≥99.8%), dimethylformamide (DMF, ≥99.8%),

130

dimethyl sulfoxide (DMSO, ≥99.9%), and trichloromethane (TCM, ≥99%) were purchased

131

from Sigma Aldrich and used as received. Epichlorohydrin (ECH, 99%), phenolphthalein

132

(ACS reagent) and phenanthrene (98%) were received from Sigma Aldrich. The distilled

133

water was produced by a Millipore Milli-Q Ultrapure System.

134

Electrospinning and Cross-linking of CD-ECH Fibers. HP--CD molecules were

135

dissolved in an aqueous solution of NaOH (25% (w/v, g/mL)), and ECH (1:10 molar ratio of

136

HP--CD/ECH, nECH/HP--CD=10) was slowly added. The solution was mixed at 50 oC for few

137

minutes to increase viscosity due to partial cross-linking and thereafter, cooled down to room

138

temperature. The solution was transferred into 3 mL syringes having sharped edged metallic

139

needles of 0.6 mm inner diameter. The syringes were horizontally placed on a syringe pump,

140

and the flow rate was set to 1 mL/h for the electrospinning. A high voltage power supply

141

(Matsusada, AU series) was used to apply 15 kV. The fibers were deposited on a grounded

6 ACS Paragon Plus Environment

Page 7 of 27 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 Polymer Materials

142

stationary rectangular metal collector, which was covered by a piece of an aluminum foil. The

143

distance between needle and collector set to 10 cm. The electrospinning was performed at ~24

144

oC

145

and kept at 150

146

electrospinning of aqueous solutions of HP--CD and HP--CD/ECH was performed. After

147

thermal treatment, poly-CD fiber webs were washed with water and ethanol, respectively to

148

remove the unreacted reagents.

in an enclosed Plexiglas chamber. After the electrospinning, the fibrous mat put in an oven oC

for 5 h for the cross-linking reactions. As control fibers, the

149

Synthesis of Poly(CD-ECH) Granules. As a control material, poly(CD-ECH) granules

150

were produced using identical concentrations of the precursors: HP--CD (120% (w/v,

151

g/mL)) was dissolved in an aqueous solution of NaOH (25% (w/v, g/mL)), and thereafter,

152

ECH (nECH/HP--CD=10) was gradually added to the solution. The solution kept stirring at 50 oC

153

for few minutes, and then, left in oven at 150 oC for 5 h for cross-linking reactions.

154

Characterization. The morphology of the poly(CD-ECH) fibers and granules was explored

155

by scanning electron microscopy (SEM, Quanta 200 FEG, FEI). Prior to SEM analysis, the

156

samples were coated with 5 nm Au/Pd using a Gatan PECS 682 etching and coating system.

157

The mean diameter of the fibers was calculated from the SEM images of the fibers with

158

ImageJ software (US National Institutes of Health, USA) by analyzing 100 fibers for each

159

sample. The rheological analysis of the samples was performed with an Anton Paar Physica

160

MCR 301 rheometer. A cone-and-plate geometry was used with a gap size of 104 m and a

161

diameter of 50 mm (angle of 1o) during the experiments. Flow tests were performed in the

162

shear rate range of 10-3-103 s-1 at 50 oC. Afterwards, the solution was subjected to a dynamic

163

oscillatory deformation test at 50 oC for 90 min at the constant shear of 0.1% and frequency of

164

1 Hz. Fourier transform infrared spectroscopy (FTIR) analysis was performed using a Bruker-

165

VERTEX 70 spectrometer system. The samples were mixed with KBr and pressed to obtain

7 ACS Paragon Plus Environment

ACS Applied Polymer Materials 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 27

166

transparent pellets. The measurements were performed between 4000 and 400 cm-1 with a

167

resolution of 4 cm-1. An X-ray photoelectron spectrometer (XPS) (Thermo Scientific) was

168

used to study the chemical composition of the fiber surface. XPS was used by means of a

169

flood gun charge neutralizer system equipped with an Al K-α X-ray source (hυ=1486.6 eV).

170

High-resolution spectra were recorded for the spectral regions of C1s at a pass energy of 50

171

eV. The thermal analysis of the samples was performed using a thermogravimetric analyzer

172

(TGA, Q500, TA Instruments). The measurements were performed between 25 and 500 oC

173

with a heating rate of 20 oC under N2 atmosphere.

174

Removal of Phenolphthalein from Aqueous Solutions. Removal tests were performed

175

with a Varian Cary 5000 spectrophotometer. The measurements were performed in the

176

wavelength range of 400-800 nm to monitor changes in the concentration of phenolphthalein

177

during the treatment with the poly(CD-ECH) fibers and granules. In this regard, the fiber

178

sample (0.1 g and 6 cm * 6 cm dimensions) and granules (0.1 g) were separately transferred

179

into the bottom of quartz cuvettes containing phenolphthalein solution (0.4 mM) prepared

180

using a buffer solution (pH= 11). The reaction was performed at room temperature (25±2 oC)

181

by continuous stirring, and the spectra were collected in a fixed time interval up to 10 h.

182

Removal of Phenanthrene from Aqueous Solutions. The removal tests of phenanthrene

183

were performed using a HPLC system (Agilent 1200 Series). The separation of phenanthrene

184

molecules was performed with a Zorbax Eclipse XDB-C18 column (150 mm × 4.6 mm, 5 µm

185

particle size), and the concentration was detected at the wavelength of 254 nm. Acetonitrile

186

(100%) was used as a mobile phase at the flow rate of 0.3 mL/min, and the injection volume

187

was 10 L. Phenanthrene was dissolved in acetonitrile and afterwards, diluted in water to

188

perform filtration measurements. The fiber sample (~0.1 g and 6 cm * 6 cm dimensions) and

189

granules were treated with phenanthrene solutions (c= 1.8 ppm) at room temperature (25±2

190

oC)

up to 27 h. The calibration curve of phenanthrene was prepared using the stock solutions 8 ACS Paragon Plus Environment

Page 9 of 27 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 Polymer Materials

191

with different concentrations: 1.8, 0.9, 0.45, and 0.23 μg/ml. The curve showed a linearity

192

with R2≥0.99. The results were adapted to this calibration curve in terms of the peak area

193

under curves. Moreover, the reusability of the poly(CD-ECH) fibrous mat was explored at

194

room temperature (25±2 oC). In this regard, the poly(CD-ECH) mat was washed with

195

acetonitrile for 10 min, which led to the removal of 80% of the adsorbed phenanthrene from

196

the fibrous mats. The samples were retreated with phenanthrene solutions at room temperature

197

(25±2 oC) for 27 h, and the sorption performance was determined as previously reported.

198

RESULTS and DISCUSSION

199

The cross-linking between HP--CD and ECH takes place over primary and secondary

200

hydroxyl groups of HP--CD. A highly alkaline condition leads to the formation of alkoxide

201

groups. Bifunctional ECH reacts with the formed alkoxide groups, giving rise to water-

202

soluble CD macromolecules through polycondensation in highly branched structures.33 The

203

reaction between the hydroxyl groups of CD molecules with the epoxide ring of ECH results

204

in glyceryl bridges between neighboring CD molecules. However, ECH molecules can also

205

react themselves under alkaline conditions, leading to the homopolymers of ECH. As ECH

206

was added to the solution prior to the electrospinning, it may lead to changes on the solution

207

properties with time. In this regard, the solution viscosity becomes a critical factor for the

208

electrospinning of CD molecules due to their partial cross-linking with ECH. Further, the

209

cross-linking reactions are highly sensitive to temperature and therefore, the cross-linking

210

significantly rises with an increase in temperature. After the addition of ECH, the reaction

211

was stirred for a short time at 50 oC, and thereafter, cooled down to room temperature. Figure

212

1a shows the flow curve of the HP--CD solution containing ECH molecules (nECH/HP--

213

CD=10).

214

shear rate ( &) range of 10-3-103 Pa·s, where the solution showed a strong decrease in the 

215

with increasing &. The solution did not reach to the infinite-shear viscosity (∞ range owing

The viscosity () of the electrospinning solution was explored by rheology in the

9 ACS Paragon Plus Environment

ACS Applied Polymer Materials 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 27

216

to the occurrence of the partial cross-linking reactions between precursors at the onset of the

217

measurement. While a dynamic oscillatory sweep test displays changes in shear moduli (i.e.,

218

elastic (G’) and loss (G’’)) over time at 50 oC because of cross-linking reactions (Figure 1b).

219

Initially, the loss modulus (G’’) was lower than the elastic modulus (G’), suggesting the sol-

220

state of the mixture. With time, G’ rapidly increased and could led to a densely cross-linked

221

network with a MPa level of G’, demonstrating the formation of a mechanically strong

222

network. The cross-over of G’ and G’’ was observed at 18 min, and thereafter, both moduli

223

drastically increased with time. This cross-over of both moduli can be considered as the

224

gelation point. Thus, the incubation time at 50 oC is highly critical and can lead to a cross-

225

linked network, which cannot be electrospun into fibers. The formation of a cross-linked

226

network can also be seen over the photos of the sample before and after heat treatment. Figure

227

1b (insets) shows the optical photos of the HP--CD/ECH solution before and after heat

228

treatment, where a color change to yellowish was observed. Further, the CD solution

229

transformed into a viscoelastic gel network at the inverted position. On the other hand, the

230

electrospinning solution did not lead to a gel network at 25 oC in (Figure S1). This is highly

231

critical since the electrospinning requires a steady-flow over time without any drastic changes

232

on the solution viscosity in a given time. With that, the electrospinnability of the system does

233

not significantly change over time throughout the electrospinning so that the proposed

234

concept is straightforward and reproducible.

235 10 ACS Paragon Plus Environment

Page 11 of 27 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 Polymer Materials

236

Figure 1. Rheological properties of an alkaline solution of HP--CD molecules in the

237

presence of ECH at 50 oC (nECH/HP--CD=10). (a) Viscosity () versus shear rate () and (b) the

238

respective shear elastic (G’) and loss (G’’) moduli of the solution as a function of time (t).

239

Inset photos show the CD solution before and after cross-linking at 50 oC for 90 min. cHP--CD=

240

120% (w/v, g/mL), NaOH (25% (w/v, g/mL)) and ECH ([ECH/HP--CD]=10).

241

When the viscosity of the electrospinning solution started to increase, the solution cooled

242

down to room temperature to prevent further cross-linking reactions and thereafter,

243

electrospun into fibers. The CD fibers were further kept in oven at 150 oC for 5 h for cross-

244

linking reactions, which led to the formation of the cross-linked poly(CD-ECH) fibers. The

245

optical photo of the poly(CD-ECH) fibers shows a free-standing electrospun mat whose SEM

246

analysis revealed the formation of bead-free fibers (Figure 2b-c). The cross-linking

247

mechanism was shown in Figure 2d, where neighboring CD molecules attached each other via

248

glyceryl bridges. The mats were stable in water due to the occurrence of highly efficient

249

cross-linking reactions between CD and ECH (Figure 2e), as revealed by rheological

250

characterization of the respective system. Further, water-contact angle (WCA) measurements

251

revealed the spreading of a water droplet on the fiber surface due to hydrophilic structure of

252

the fibrous mat owing to the hydrophilicity of both HP--CD and ECH and formed glyceryl

253

bridges (Figure 2f).

11 ACS Paragon Plus Environment

ACS Applied Polymer Materials 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 27

254 255

Figure 2. (a) An optical photo of the poly(CD-ECH) mat and inset photo (i) shows the

256

folded fiber mat. (b) Scanning electron micrograph of CD fibers before (b) and after (c) cross-

257

linking (cHP--CD= 120% (w/v, g/mL), NaOH (25% (w/v, g/mL)) and ECH ([ECH/HP--CD]=

258

10). (d) Cross-linking mechanism between CD and ECH, and (e) the poly(CD-ECH) mat

259

exposed to water, showing water-insoluble nature and structure stability of the poly(CD-ECH)

260

mat in water. (f) An optical photo shows the spreading of a water droplet on the poly(CD-

261

ECH) mat.

262

The electrospraying of the HP--CD molecules took place at the CD concentration of 50%

263

(w/v, g/mL) and [ECH/CD] of 8 and 10, while increasing the concentration of HP--CD to 75

264

and 100% (w/v, g/mL) led to beaded fibers (Figure S2, Supporting Information). On the other

265

hand, with increasing the CD concentration to 120% (w/v, g/mL), the electrospinning of the

266

HP--CD solution led to bead-free fibers with a mean diameter of 1.01 ±0.73 m, which is

267

higher than the ECH-free HP--CD fibers, demonstrating that the partial cross-linking of HP-

268

-CD molecules increased the viscosity and gave rise to the formation of larger fibers (Figure

269

3a). On the other hand, mean size of the fibers prior to the cross-linking was calculated as 12 ACS Paragon Plus Environment

Page 13 of 27 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 Polymer Materials

270

1.06 ±0.53 m, suggesting that cross-linking did not induce significant changes on the fiber

271

size (Figure 2b). The mean fiber diameter of the ECH-free CD system at the HP--CD

272

concentration of 160% (w/v, g/mL) was previously reported as 0.75 m25 while it increased to

273

1.01 ±0.73 m with the incorporation of ECH, suggesting that the occurrence of partial cross-

274

linking prior to the electrospinning process. This was also seen over the concentration of HP-

275

-CD needed to form bead-free fibers. Normally, the concentration of HP--CD should be

276

over 160% (w/v, g/mL) in water form to obtain bead-free fibers. While it decreased to 120%

277

(w/v, g/mL) when ECH was added to solution for the cross-linking. The partial cross-linking

278

of HP--CD molecules prior to the electrospinning led to higher viscosity, and therefore, the

279

lower concentration of HP--CDs was enough to produce bead-free fibers. After exposure to

280

water for 24 h, the fiber size remained nearly unchanged (Figure 3b).

281 282 283

Figure 3. Scanning electron micrographs of the cross-linked poly(CD-ECH) fibers before (a) and after (b) exposure to water for 24 h.

284

The stability of the poly(CD-ECH) fibers was also tested in various organic solvents (i.e.,

285

methanol, ethanol, acetonitrile, DMF, TCM and DMSO) for 24 h. The SEM photos of the

286

respective fibers showed different fiber morphologies depending on the solvent treated

13 ACS Paragon Plus Environment

ACS Applied Polymer Materials 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 27

287

(Figure 4). The poly(CD-ECH) fibers were stable in all solvents tested, and could preserve

288

their fibrous morphology. The thinnest fibers were observed after exposure to MeOH,

289

followed by EtOH, DMSO, DMF, and TCM, while the largest fibers were observed after

290

treatment with ACN. It is known that alcohols are bad solvents for CD systems and can lead

291

to the collapse of the fiber matrix. Thus, the smallest diameter in alcohols can be attributed to

292

their non-solvent character for CD molecules.

293 294

Figure 4. Scanning electron micrographs of the cross-linked poly(CD-ECH) fibers after

295

exposure to various solvents for 24 h. (a) TCM, (b) DMF, (c), DMSO, (d) EtOH, (e) MeOH

296

and (f) ACN.

297

The cross-linking of the HP--CD fibers with ECH was confirmed by XPS measurements

298

(Figure 5). The atomic composition of the HP--CD and poly(CD-ECH) fibers was shown in

299

Table S1 (Supporting Information). The XPS analysis of the poly(CD-ECH) fibers revealed

300

the presence of Na1s (4.27%) and Cl2p (3.28%) atoms at significant proportions. While no

14 ACS Paragon Plus Environment

Page 15 of 27 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 Polymer Materials

301

peaks related to these elements appeared for the pure HP--CD fibers since they were

302

electrospun from water in the absence of ECH. The deconvoluted C1s spectra show a broad

303

peak related to C atom, which can further be deconvoluted into three peaks as follows: the

304

peaks appeared at 284.8, 286, and 288 eV are respectively ascribed to C-C (or C-H), C-O and

305

O-C-O.34 The broader C peak centered at 286.4 eV was observed for the poly(CD-ECH)

306

fibers due to the formation of glyceryl bridges with ECH. On the other hand, the atomic

307

percentage of C-C drastically decreased from 10.16 to 3.6% with the incorporation of ECH,

308

which increased the ratio of C-O in overall composition from 45.6 to 50.45% by taking

309

account of C and O atoms (Table S1).

310 311

Figure 5. The deconvoluted C1s spectra of (a) HP--CD and (b) poly(CD-ECH) fibers.

312

The chemical analysis of the poly(CD-ECH) fibers was further explored by FT-IR analysis

313

(Figure 6a). The bands between 1035 and 1155 cm-1 are due to the stretching band of the

314

glycosidic links of CD molecules: the IR bands at 1155, 1082 and 1035 cm-1 can be attributed

315

to the stretching vibration of C-H and C-O bonds. Similar IR bands were also observed in

316

poly(CD-ECH) fibers. Here, the absorbance ratio (A1082/A1035) in this region of the IR

317

spectrum gives information about the degree of cross-linking and A1082/A1035 ratio of

318

poly(CD-ECH) fibers is 66% higher compared to HP--CD fibers. This ratio reveals the

319

partial cross-linking such that around 66% (approximately 66% of OH groups) of the primary 15 ACS Paragon Plus Environment

ACS Applied Polymer Materials 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 27

320

OH groups of HP--CD transformed into secondary OH forming the cross-link junction.27 For

321

the poly(CD-ECH) fibers, the peak intensity at 1456 cm-1 due to the bending of methylene

322

groups of the formed glyceryl links drastically increased, while the peaks at 948 and 1143 cm-

323

1

324

presence of unreacted epoxy groups. The thermal analysis of the fibers was carried out using

325

TGA, which revealed the enhanced stability of HP--CD fibers after cross-linking with ECH

326

(Figure 6b). The pyrolysis of HP--CD molecules starts over 300 oC, while it increased to 372

327

from 347 oC due to thermal stabilization of the CD fibers with the cross-linking by ECH. It is

328

known that the cross-linking of molecules generally rises the degradation temperature, and

329

thus, an increase in the degradation temperature can be attributed to the presence of a densely

330

cross-linked CD network, which stabilized the CD structure on the course of heating.

decreased for the poly(CD-ECH) fibers. The typical absorption band at 850 cm-1 shows the

331 332 333

Figure 6. (a) FTIR and (b) TGA of the pure HP--CD fibers, uncross-linked HP-CD/ECH fibers and poly(CD-ECH) fibers.

334

The structural properties of the pure HP--CD fibers and poly(CD-ECH) fibers were also

335

explored by wide-angle XRD measurements (Figure S3, Supporting Information). Unlike

336

native CD molecules (,  and ), hydroxypropyl-substituted CD molecules are amorphous

337

and give a broad peak centered at 19o. Likewise, the XRD patterns of the pure HP--CD 16 ACS Paragon Plus Environment

Page 17 of 27 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 Polymer Materials

338

fibers and poly(CD-ECH) fibers displayed amorphous diffraction peaks. Interestingly, the

339

peak positions the HP--CD and poly(CD-ECH) fibers shifted to higher 2 and peak

340

intensities decreased with respect to the powder form of HP--CD (Figure S3). This shift and

341

decrease in the intensity after cross-linking may be attributed to the loss of structural

342

arrangement (i.e., an increase in the amorphous structure) of the HP--CD with heat

343

treatment.

344

The presence of active CD contents in the cross-linked (poly(CD-ECH) fibers was explored

345

using phenolphthalein over host-guest inclusion-complexation of accessible CD molecules.

346

Phenolphthalein is a model guest molecule, which is widely used for the exploration of active

347

CD contents in the CD-based materials.35 Furthermore, it is a dye molecule with some

348

toxicity.29 Thus, the removal of phenolphthalein from aqueous media is desired for water

349

treatment. It forms IC with CD molecules, which can easily be monitored through UV-Vis

350

measurements due to the change of its pink color to colorless, driven by the constrained

351

conformation of phenolphthalein in the CD cavity.28, 35 The removal performance was further

352

compared with the granule form of the poly(CD-ECH). Figure 7 shows the UV-Vis spectra of

353

phenolphthalein solutions exposed to the poly(CD-ECH) fibers and granules over time. The

354

maximum absorption intensity at 553 nm decreased with time, and after 120 min no peak

355

related to phenolphthalein was observable, suggesting about 80% phenolphthalein formed

356

inclusion complex with accessible CD molecules. The complete removal of phenolphthalein

357

was accomplished in 2h while 50% removal was reached in 25 min (Fig 7a-b), suggesting the

358

rapid removal of phenolphthalein molecules from water. On the other hand, the identical

359

concentration of phenolphthalein could be removed in 10 h using poly(CD-ECH) granules

360

(Figure 7b-d). The removal of phenolphthalein by the poly(CD-ECH) fibers was also shown

361

over the optical photos of the solutions over time (Figure 7b inset) with the decolorization of

362

the phenolphthalein solution. The faster removal of phenolphthalein by the poly(CD-ECH) 17 ACS Paragon Plus Environment

ACS Applied Polymer Materials 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 27

363

fibers can be attributed to their smaller size and the presence of a highly porous structure. The

364

morphology of the poly(CD-ECH) fibers could be preserved after the sorption tests,

365

suggesting their high potential for water treatment applications from organic micropollutants

366

(Figure S4).

367 368

Figure 7. Filtration performance of the electrospun poly(CD-ECH) fibrous mat over

369

phenolphthalein. (a) UV-Vis spectra of phenolphthalein solution treated with the poly(CD-

370

ECH) fibers as a function of time and (b) decrease in the absorbance (λ=553 nm) of the

371

phenolphthalein solution with time. (c) UV-Vis spectra of a phenolphthalein solution treated

372

with the poly(CD-ECH) granules with time and (d) decrease in the absorbance intensity

373

(λ=553 nm) over time.

374

The electrospun poly(CD-ECH) fibrous mat was also used for the removal of phenanthrene

375

from aqueous solutions with time (Figure 8). The concentration of phenanthrene was

376

measured over time by HPLC. The stock solutions of phenanthrene was prepared at the 18 ACS Paragon Plus Environment

Page 19 of 27 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 Polymer Materials

377

maximum solubility limits (i.e., 1.8 g/mL) in water/ACN mixtures (59 mL/10 L). Within

378

the first 40 min, 50% removal of the phenanthrene was accomplished, while after 350 min the

379

concentration of phenanthrene decreased to 0.2 g/mL. The removal performance of

380

poly(CD-ECH) fibrous mat was very high and could scavenge 88% of initial phenanthrene

381

content. On the other hand, poly(CD-ECH) granules could reach to an equilibrium sorption

382

level at ~900 min. Both fibrous mat and granules exhibited similar sorption levels, but in

383

different rate. The use of poly(CD-ECH) fibrous mat fastened the sorption rate by nearly 3-

384

fold, which can be attributed to the smaller size of the fibers and their highly porous structure.

385

Despite showing a similar sorption capacity, there are drastic differences in the sorption rate. This can

386

be attributed to the enhanced surface area for the nanofiber form of the poly(CD-ECH), which boosts

387

the sorption rate. For many applications, rapid removal of such pollutants is highly desired.

388

Furthermore, unlike granule form of the poly(CD-ECH), nanofibers are easy to use and do not require

389

any specific container during their use. Furthermore, after washing with ACN to remove

390

phenanthrene molecules from the poly(CD-ECH) mat, the poly(CD-ECH) mat was retreated

391

with phenanthrene solution containing an identical phenanthrene concentration of the first

392

solution. The poly(CD-ECH) mat could remove 70% of phenanthrene after second use

393

(Figure S5). The performance loss in removal capacity can be attributed to the presence of

394

remained phenanthrene molecules in CD cavities, which could not be washed away easily

395

after the first use. The SEM analysis of the poly(CD-ECH) fibers and poly(CD-ECH)

396

granules after sorption tests revealed no significant changes in their morphology (Figure S6).

397

The sorption capacity for phenanthrene was found to be 944 g per gram poly(CD-ECH)

398

fibers. This is higher than many sorbent materials used for phenanthrene removal from water

399

(Table S1, Supporting Information). Although the presence of some sorbent systems with

400

better sorption values, they used alcohols as co-solvent to increase the solubility of

401

phenanthrene. However, among water-based systems, poly(CD-ECH) fibers showed a good

19 ACS Paragon Plus Environment

ACS Applied Polymer Materials 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 27

402

sorption capacity, which is better than DNA-based nanogels (sorption capacity = 720 g/g)36

403

and CD-grafted cellulose acetate fibers (sorption capacity = 540 g/g)19 (Table S2).

404 405 406

Figure 8. Time-dependent removal of phenanthrene from aqueous solutions using poly(CDECH) fibers and granules.

407

The poly(CD-ECH) electrospun fibers are a novel type of materials with great promising

408

applications for water treatment. As we have shown here with the removal of phenolphthalein

409

and phenanthrene, they can be exploited as high-performance sorbent materials due to their

410

high specific surface area, along with a high active CD content. Given that their high stability

411

in aqueous and various organic solvent systems owing to their highly cross-linked network by

412

epichlorohydrin-based chemistry, these highly stable fibrous materials may find more

413

application area as highly effective sorbents in coming years. Unlike other methods used for

414

the cross-linking for CD molecules, epichlorohydrin-based cross-linking offers robust CD

415

networks, which are stable even in polar aprotic solvents, such as DMF. This allows their

416

reuse for sorption-based applications after removal of the adsorbed molecules with intense

417

washing with aprotic solvents.

20 ACS Paragon Plus Environment

Page 21 of 27 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 Polymer Materials

418 419

CONCLUSIONS

420

Insoluble poly(CD-ECH) fibers were produced by the electrospinning of the aqueous

421

solutions of HP--CD and ECH and their cross-linking by heat-treatment. The flow and

422

dynamic oscillatory deformation tests revealed partial cross-linking of CD molecules prior to

423

the electrospinning, however main cross-linking takes place after the heat-treatment of the

424

fibers at 150 oC. Due to highly efficient cross-linking reactions, poly(CD-ECH) fibers were

425

stable in water and various organic solvents (i.e., ACN, TCM, DMSO, DMF, MeOH and

426

EtOH) and could maintained their fibrous structure. The presence of high active CD content

427

in the fibers was revealed by removal tests of phenolphthalein and phenanthrene through

428

inclusion-complexation. The removal performance of poly(CD-ECH) fibers was compared

429

with poly(CD-ECH) granules, and the results showed that rapid removal of phenolphthalein

430

and phenanthrene by the fibers compared to the granule form of poly(CD-ECH) due to their

431

smaller size and high porous structure. The poly(CD-ECH) fibrous membrane with a high

432

specific surface area and structural stability in water and organic solvents are promising

433

materials in water treatment applications.

434

ASSOCIATED CONTENT

435

Supporting Information

436

The following files are available free of charge.

437

Dynamic oscillatory time sweep profile of an aqueous mixture of HP--CD and ECH. The

438

SEM images of the poly(CD-ECH) fibers and granules synthesized at various concentrations

439

and after their use in removal of phenolphthalein and phenanthrene. The atomic composition

21 ACS Paragon Plus Environment

ACS Applied Polymer Materials 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 27

440

of the fibers obtained by XPS analysis. The reuse of poly(CD-ECH) fibers in the removal of

441

phenanthrene.

442

AUTHOR INFORMATION

443

Corresponding Author

444

* E-mail: [email protected] (T.U)

445

ORCID ID

446

F. T.: 0000-0002-9011-4495

447

T. U.: 0000-0002-3989-4481

448 449

Funding:

450

The Scientific and Technological Research Council of Turkey (TUBITAK, project number:

451

113Y348) is acknowledged for funding the research.

452

Author contributions:

453

A.C., F.T and T.U. conceived, built and carried out the experiment and analyzed the data. All

454

authors contributed to the preparation of the manuscript.

455

Notes:

456

The authors declare no competing financial interest.

457

REFERENCES:

458

(1)

10975.

459 460

(2)

Irie, T.; Uekama, K., Pharmaceutical Applications of Cyclodextrins. III. Toxicological Issues and Safety Evaluation. J. Pharm. Sci. 1997, 86 (2), 147-162.

461 462

Crini, G., Review: A History of Cyclodextrins. Chem. Rev. 2014, 114 (21), 10940-

(3)

Astray, G.; Gonzalez-Barreiro, C.; Mejuto, J. C.; Rial-Otero, R.; Simal-Gándara, J., A

463

review on the use of cyclodextrins in foods. Food Hydrocolloids 2009, 23 (7), 1631-

464

1640. 22 ACS Paragon Plus Environment

Page 23 of 27 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

465

ACS Applied Polymer Materials

(4)

solubilization and stabilization. J. Pharm. Sci. 1996, 85 (10), 1017-1025.

466 467

Loftsson, T.; Brewster, M. E., Pharmaceutical applications of cyclodextrins. 1. Drug

(5)

Li, J. J.; Zhao, F.; Li, J., Supramolecular Polymers Based on Cyclodextrins for Drug

468

and Gene Delivery. In Biofunctionalization of Polymers and their Applications,

469

Nyanhongo, G. S.; Steiner, W.; Gübitz, G., Eds. Springer Berlin Heidelberg: Berlin,

470

Heidelberg, 2011; pp 207-249.

471

(6)

Expert Opinion on Drug Delivery 2005, 2 (2), 335-351.

472 473

Loftsson, T.; Jarho, P.; Másson, M.; Järvinen, T., Cyclodextrins in drug delivery.

(7)

Prabaharan, M.; Jayakumar, R., Chitosan-graft-β-cyclodextrin scaffolds with

474

controlled drug release capability for tissue engineering applications. Int. J. Biol.

475

Macromol. 2009, 44 (4), 320-325.

476

(8)

Szejtli, J., Cyclodextrins in the Textile Industry. Starch - Stärke 2003, 55 (5), 191-196.

477

(9)

Topuz, F.; Uyar, T., Poly-cyclodextrin cryogels with aligned porous structure for

478

removal of polycyclic aromatic hydrocarbons (PAHs) from water. J. Hazard. Mater.

479

2017, 335, 108-116.

480

(10)

Topuz, F.; Uyar, T., Cyclodextrin-functionalized mesostructured silica nanoparticles

481

for removal of polycyclic aromatic hydrocarbons. J. Colloid Interface Sci. 2017, 497,

482

233-241.

483

(11)

Mhlanga, S. D.; Mamba, B. B.; Krause, R. W.; Malefetse, T. J., Removal of organic

484

contaminants from water using nanosponge cyclodextrin polyurethanes. J. Chem.

485

Technol. Biotechnol. 2007, 82 (4), 382-388.

486

(12)

Uyar, T.; Havelund, R.; Hacaloglu, J.; Besenbacher, F.; Kingshott, P., Functional

487

Electrospun Polystyrene Nanofibers Incorporating α-, β-, and γ-Cyclodextrins:

488

Comparison of Molecular Filter Performance. ACS Nano 2010, 4 (9), 5121-5130.

489

(13)

Zhao, F.; Repo, E.; Yin, D.; Chen, L.; Kalliola, S.; Tang, J.; Iakovleva, E.; Tam, K. C.;

490

Sillanpää, M., One-pot synthesis of trifunctional chitosan-EDTA-β-cyclodextrin

491

polymer for simultaneous removal of metals and organic micropollutants. Sci. Rep.

492

2017, 7 (1), 15811. 23 ACS Paragon Plus Environment

ACS Applied Polymer Materials 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

493

(14)

Page 24 of 27

Alsbaiee, A.; Smith, B. J.; Xiao, L.; Ling, Y.; Helbling, D. E.; Dichtel, W. R., Rapid

494

removal of organic micropollutants from water by a porous β-cyclodextrin polymer.

495

Nature 2015, 529, 190.

496

(15)

Xiao, L.; Ling, Y.; Alsbaiee, A.; Li, C.; Helbling, D. E.; Dichtel, W. R., β-

497

Cyclodextrin Polymer Network Sequesters Perfluorooctanoic Acid at Environmentally

498

Relevant Concentrations. J. Am. Chem. Soc. 2017, 139 (23), 7689-7692.

499

(16)

cyclodextrin–epichlorohydrin polymers. Prog. Polym. Sci. 2013, 38 (2), 344-368.

500 501

Morin-Crini, N.; Crini, G., Environmental applications of water-insoluble β-

(17)

Morin-Crini, N.; Winterton, P.; Fourmentin, S.; Wilson, L. D.; Fenyvesi, É.; Crini, G.,

502

Water-insoluble β-cyclodextrin–epichlorohydrin polymers for removal of pollutants

503

from aqueous solutions by sorption processes using batch studies: A review of

504

inclusion mechanisms. Prog. Polym. Sci. 2018, 78, 1-23.

505

(18)

Uyar, T.; Havelund, R.; Nur, Y.; Hacaloglu, J.; Besenbacher, F.; Kingshott, P.,

506

Molecular filters based on cyclodextrin functionalized electrospun fibers. J. Membr.

507

Sci. 2009, 332 (1), 129-137.

508

(19)

Celebioglu, A.; Demirci, S.; Uyar, T., Cyclodextrin-grafted electrospun cellulose

509

acetate nanofibers via “Click” reaction for removal of phenanthrene. Appl. Surf. Sci.

510

2014, 305, 581-588.

511

(20)

Zhao, R.; Wang, Y.; Li, X.; Sun, B.; Wang, C., Synthesis of β-Cyclodextrin-Based

512

Electrospun Nanofiber Membranes for Highly Efficient Adsorption and Separation of

513

Methylene Blue. ACS Applied Materials & Interfaces 2015, 7 (48), 26649-26657.

514

(21)

Schäfer, A. I.; Stelzl, K.; Faghih, M.; Sen Gupta, S.; Krishnadas, K. R.; Heißler, S.;

515

Pradeep, T., Poly(ether sulfone) Nanofibers Impregnated with β-Cyclodextrin for

516

Increased Micropollutant Removal from Water. ACS Sustainable Chemistry &

517

Engineering 2018, 6 (3), 2942-2953.

518

(22)

Kayaci, F.; Aytac, Z.; Uyar, T., Surface modification of electrospun polyester

519

nanofibers with cyclodextrin polymer for the removal of phenanthrene from aqueous

520

solution. J. Hazard. Mater. 2013, 261, 286-294.

24 ACS Paragon Plus Environment

Page 25 of 27 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

521

ACS Applied Polymer Materials

(23)

Zhang, W.; Chen, M.; Zha, B.; Diao, G., Correlation of polymer-like solution

522

behaviors with electrospun fiber formation of hydroxypropyl-β-cyclodextrin and the

523

adsorption study on the fiber. PCCP 2012, 14 (27), 9729-9737.

524

(24)

Celebioglu, A.; Sen, H. S.; Durgun, E.; Uyar, T., Molecular entrapment of volatile

525

organic compounds (VOCs) by electrospun cyclodextrin nanofibers. Chemosphere

526

2016, 144, 736-744.

527

(25)

Celebioglu, A.; Uyar, T., Electrospinning of nanofibers from non-polymeric systems:

528

polymer-free nanofibers from cyclodextrin derivatives. Nanoscale 2012, 4 (2), 621-

529

631.

530

(26)

2010, 46 (37), 6903-6905.

531 532

Celebioglu, A.; Uyar, T., Cyclodextrin nanofibers by electrospinning. Chem. Commun.

(27)

Celebioglu, A.; Yildiz, Z. I.; Uyar, T., Electrospun crosslinked poly-cyclodextrin

533

nanofibers:

534

complexation. Sci. Rep. 2017, 7 (1), 7369.

535

(28)

Highly

efficient

molecular

filtration

thru

host-guest

inclusion

Kuwabara, T.; Takamura, M.; Matsushita, A.; Ikeda, H.; Nakamura, A.; Ueno, A.;

536

Toda, F., Phenolphthalein-Modified β-Cyclodextrin as a Molecule-Responsive

537

Colorless-to-Color Change Indicator. The Journal of Organic Chemistry 1998, 63

538

(24), 8729-8735.

539

(29)

Effects in Experimental Model Systems. Cancer Res. 1996, 56 (21), 4922-4926.

540 541

(30)

Artymowicz, R. J.; Childs, A. L.; Paolini, L., Phenolphthalein-Induced Toxic Epidermal Necrolysis. Ann. Pharmacother. 1997, 31 (10), 1157-1159.

542 543

Dunnick, J. K.; Hailey, J. R., Phenolphthalein Exposure Causes Multiple Carcinogenic

(31)

Taguchi, K., Transient binding of phenolphthalein-.beta.-cyclodextrin complex: an

544

example of induced geometrical distortion. J. Am. Chem. Soc. 1986, 108 (10), 2705-

545

2709.

546 547

(32)

Shuttleworth, K. L.; Cerniglia, E., Environmental aspects of PAH biodegradation. Appl. Biochem. Biotechnol. 1995, 54 (1), 291-302.

25 ACS Paragon Plus Environment

ACS Applied Polymer Materials 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

548

(33)

Page 26 of 27

Renard, E.; Deratani, A.; Volet, G.; Sebille, B., Preparation and characterization of

549

water soluble high molecular weight β-cyclodextrin-epichlorohydrin polymers. Eur.

550

Polym. J. 1997, 33 (1), 49-57.

551

(34)

Transport and Shape-Selective Permeability. Adv. Mater. 2017, 29 (26), 1606641.

552 553

Villalobos, L. F.; Huang, T.; Peinemann, K.-V., Cyclodextrin Films with Fast Solvent

(35)

Kettel, M. J.; Hildebrandt, H.; Schaefer, K.; Moeller, M.; Groll, J., Tenside-free

554

Preparation of Nanogels with High Functional β-Cyclodextrin Content. ACS Nano

555

2012, 6 (9), 8087-8093.

556

(36)

Topuz, F.; Singh, S.; Albrecht, K.; Möller, M.; Groll, J., DNA Nanogels To Snare

557

Carcinogens: A Bioinspired Generic Approach with High Efficiency. Angew. Chem.

558

Int. Ed. 2016, 55 (40), 12210-12213.

559 560 561 562 563 564 565 566 567 568 569 570 571

26 ACS Paragon Plus Environment

Page 27 of 27 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 Polymer Materials

572 573

Table of Contents (ToC) Graphic

574

27 ACS Paragon Plus Environment