Fiber Alignment and Liquid Crystal Orientation of Cellulose

Sep 19, 2017 - Sulfate cellulose nanocrystal (CNC) dispersions always present specific self-assembled cholesteric mesophases which is easily affected ...
0 downloads 14 Views 2MB Size
Subscriber access provided by University of Sussex Library

Article

Fiber Alignment and Liquid Crystal Orientation of Cellulose Nanocrystals in the Electrospun Nanofibrous Mats Weiguang Song, Dagang Liu, Nana Prempeh, and Renjie Song Biomacromolecules, Just Accepted Manuscript • DOI: 10.1021/acs.biomac.7b00927 • Publication Date (Web): 19 Sep 2017 Downloaded from http://pubs.acs.org on September 20, 2017

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

Biomacromolecules 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 25

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

Biomacromolecules

1

Fiber Alignment and Liquid Crystal Orientation of

2

Cellulose Nanocrystals in the Electrospun Nanofibrous

3

Mats

4

Weiguang Song,a,b Dagang Liu,*a,b,c Nana Prempeh,b Renjie Song a

5

6

7

a

Department of Chemistry, Nanjing University of Information Science and Technology, Nanjing 210044, China

8

9

b

Collaborative Innovation Center of Atmospheric Environment and Equipment

10

Technology, Nanjing University of Information Science and Technology, Nanjing

11

210044, China

12

c

Department of Physics, University of Colorado, Boulder, CO 80309, USA

13

ACS Paragon Plus Environment

Biomacromolecules

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

14

ABSTRACT:

15

Sulfate cellulose nanocrystal (CNC) dispersions always present specific self-assembled

16

cholesteric mesophases which is easily affected by the inherent properties of particle size,

17

surface charge and repulsion or affinity interaction, and external field force generated from ionic

18

potential of added electrolytes, magnetic or electric field, and/or mechanical shearing or

19

stretching. Aiming at understanding the liquid crystal orientation and fiber alignment under high-

20

voltage electric field, randomly distributed, uniform-aligned, or core-sheath nanofibrous mats

21

involving charged CNCs and PVA were electrospun; and amongst them, specific straight arrayed

22

fine nanofibers with average diameter of 270 nm were manufactured by using a simple and

23

versatile gap collector. Moreover, arrayed composite nanofibers regularly aligned along the

24

vertical direction of gap plates and selectively reflected frequent and continuous birefringence

25

which was regarded as nematic phases of CNCs induced by the uniaxial stretching under high-

26

voltage electric field. As a synergic effect of rigidness of nanocrystals and stretching orientation

27

of nematic phases, the aligned nanofibrous arrays exhibited a higher tensile strength and strain

28

than the randomly-oriented or core-sheath nanofibrous mats at the same loading of CNCs. By

29

contrast, mesophase transition of CNCs from cholesteric to nematic occurred in the coaxially

30

spun core-sheath nanofibers at a loss of long-ranged chiral twist. Hence, the structure-effect

31

relationship between liquid crystal orientation of charged nanorods in polymer-based fine

32

nanofibers and the flexibility or mechanical integrity of the aligned fiber array will be favorable

33

for strategic development of functional liquid crystal fabrics.

34

Keywords: Cellulose nanocrystals, PVA, Electrospinning, Nanofibrous mats, Alignment

ACS Paragon Plus Environment

Page 2 of 25

Page 3 of 25

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

35

Biomacromolecules

INTRODUCTION

36

Up until now, electrospinning has been accepted as a simple and efficient method for

37

developing polymer nanofibers with width range from 3 nm to 5 µm,1,2 depending on the

38

spinning parameters. Under a high-voltage electric field the electrospinning process itself

39

involves polymer droplets stretching and deformation into conical cylindrical threads with a

40

rapid evaporation of the solvent.3 Electrospun nanofibers have several remarkable characteristics

41

such as porosity,4 high aspect ratio,5 flexibility in surface functionalities, and special mechanical

42

and biological properties which makes them excellent candidates for applications such as

43

nanofibrous membranes or filters,6 mats in tissue engineering constructs and wound dressings,7,8

44

military wears with chemical and biological toxin-resistance,9 electronic sensors,10 and so on.

45

Generally, the flow rate, voltage, and tip-to-collector distance are thought as very important

46

processing parameters that directly affects the morphology, size distribution, and mechanical

47

performance of nanofiber,11,12 whereas, instability of polymer jet streams has often led to poorly

48

aligned fibers with varying morphology and diameters during the electrospinning process.

49

Moreover, the disorderliness in fiber formation is problematic as much more applications

50

requiring the use of fibrous mats relies heavily on well-aligned and highly ordered

51

architectures.13 Up to now, several approaches such as modification of the collection devices by

52

the use of rectangular metal frame,14 central point electrode and peripheral ring electrode,15 a pair

53

of parallel electrodes or addition of electrode,16 circular copper wire drum collector, 17 or an

54

additive electric field,18 have been considered for the development of aligned and continuous

55

nanofibers.

56

Sulfuric acid hydrolysis of native cellulose microfibrils produces suspensions of cellulose

57

nanocrystals (CNCs), which are highly stable in aqueous suspensions because of the repulsion

ACS Paragon Plus Environment

Biomacromolecules

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

58

among negatively charged sulfate ester bearings on the surface of CNCs.19,20 While the

59

homogeneous suspension of sulfate CNCs is solidified after a slow evaporation induced self-

60

assembly process, cholesteric liquid crystalline films with chiral reflective iridescence are

61

formed.21,22 Namely, the solidified films generated from liquid crystalline suspensions inherits a

62

short-range position arrangement and long-range chiral nematic orientation of 3-D photonic

63

CNC colloids.23 This fascinating liquid crystal properties of CNCs have received an increasing

64

attention for the fabrication of novel functional chiroptical materials.24,25 At hand, the scalization

65

and industrialization of CNC based materials of film, coating, fiber, etc., are confined by the

66

stiffness and fragile nature of CNCs. Several researches involving neutral polymers of PEG,26

67

PEO,27 and PVA,28 with excellent compatibility and affinity with/to CNCs was performed to

68

improve the flexibility of generated composites. More interestingly, both PEG,26 and PVA,28

69

presented severe impact on the structure-color and chiral nematic mesophases of CNCs due to

70

microphase separation, thereby leading to the loss of the long-range orientation and chiroptical

71

reflectivity of bulk coatings, films or fibers exceeding a critical loading content of the compatible

72

polymers.

73

More recently, nanofibrous composite mats of PVA was electrospun by employing CNCs

74

with loading dose of 0-15 wt% as effective reinforcement agents.29-31 As a matter of fact, it was

75

thought that overall tension in the spinning dope of CNCs with a high charge density on the

76

surface would self-repulse the excess charges on the jet, thus, as the charge density increase, the

77

continuity, stability, and alignment of the fibers meet great challenges in electrospinning.

78

Notwithstanding the foregoing, no research work on the alignment of nanofibrous CNC

79

composite mats as well as the orientation of CNCs in a liquid jet under high voltage electric field

80

during electrospinning is reported and discussed in detail. In this work, single-fluid of composite

ACS Paragon Plus Environment

Page 4 of 25

Page 5 of 25

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

Biomacromolecules

81

dispersions or dispersion mixture containing CNC and hydrophilic PVA was deployed as

82

spinning dope and a tailored gap between a couple electrode plates or a rotating roller was

83

applied as a collection device of electrospinning. Aiming at improving the alignment of

84

nanofibers and orientation of highly surface-charged CNCs, we attempted to set up a new

85

approach to fabricate nanofiber arrays by controlling the physicochemical properties of spinning

86

dope and processing parameters, and tended to understand the mesophase orientation and

87

alignment of fibers under electrical stretching or repulsive and attractive interaction, and

88

anticipated to build up the relationship between the tailored composing structure of random

89

distribution, uniformed alignment, or core-sheath and properties of as-spun fibrous mats.

90

MATERIALS AND METHODS

91

Materials. PVA (AH-26, Mw = 114,400) having a degree of polymerization (DP) of about

92

2600 and saponification of 98% and microcrystalline cellulose (MCC, 9004-34-6) of analytical

93

grade were purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China).

94

Concentrated sulfuric acid of analytical grade was purchased from Shanghai Lingfeng Chemical

95

Reagent Co. Ltd.

96

Preparation of spinning dope. Referring to our previous work,23 MCC was hydrolyzed by

97

sulfuric acid (64 wt%) at 50 oC in a water bath under vigorous mechanical stirring for 2.5 h. The

98

resultant suspension was diluted 5-fold by deionized water and then poured into cellulose

99

dialysis tube with MWCO of 8,000-14,000 to dialyze against deionized water for more than one

100

week to remove residual sulfuric acid and many small molecules. Subsequently, impurities and

101

unmodified coarse particles in the dialyzed suspension was removed by centrifugation at 12000

102

rpm for 10 min. The as-prepared CNCs after a mild sonication have a length of 150 ± 30 nm and

103

width of 35 ± 5 nm, and a negative surface charge density of 0.60 e/nm2.23 Alternatively, PVA

ACS Paragon Plus Environment

Biomacromolecules

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

104

was dissolved in deionized water at 90 oC under strong mechanical stirring for 1 h to produce a

105

homogeneous solution with a concentration of 10 wt%. Subsequently, CNC suspension and PVA

106

solution were mixed in specific proportions under continuous agitation for 4 days at 30 oC. The

107

individual resultant PVA/CNC mixture as a single fluidic spinning dope was then stored in the

108

refrigerator to allow a slow evaporation and absolute self-assembly at 10 oC until a desired

109

concentration of the spinnable dope was obtained.

110

Electrospinning of PVA/CNC nanofiber mats. As illustrated in Supplementary Figure

111

S1a, the electrospinning set-up of roller-collection comprised of a high voltage power supply (0-

112

50 DC kV), a micro-syringe pump (V = 20 mL) connected to a metal needle (di = 0.40 mm) or

113

coaxial spinneret (do = 2.00 mm; di = 0.40 mm), and a collector of grounded rotating cylindrical

114

roller (D = 4 cm; L = 15.5 cm) with a tip-to-collector distance of 0-20 cm. In this roller-

115

collection device, the spinneret acting as a positive electrode was connected to the high voltage

116

power supply to generate electric field. During the electrospinning process, the dope was

117

pumped through a polytetrafluoroethylene (PTFE) tube, and then injected from the metal

118

spinneret as a fluidic jet at a specific rate. After water evaporation, the solidified jet in the form

119

of nanofibers were collected on the metallic cylindrical roller. According to mass ratio of

120

CNC/PVA at 1:9, 1:3, and 1:1, the manufactured nanofibers mats with random orientation were

121

denoted as PVA/CNC10R, PVA/CNC25R, and PVA/CNC50R, respectively. Alternatively, in a

122

coaxial electrospinning process, CNC suspension and PVA solution were independently fed into

123

a coaxial spinneret at a speed of 0.2 mL/h and 0.6 mL/h, respectively, and spun in a core-shell

124

manner. The fine core-sheath nanofibers of PVA/CNC collected on the metallic cylindrical roller

125

were denoted as PVA@CNC10 and CNC10@PVA corresponding to mass ratio (1:9) of CNCs to

126

PVA in the fabricated core-sheath nanofibers.

ACS Paragon Plus Environment

Page 6 of 25

Page 7 of 25

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

Biomacromolecules

127

Specifically, we set up a tailored gap collector with a couple of parallel plates (Figure S1b)

128

involving a negative metal plate of 30-mm thickness and a ground polyoxymethylene (POM)

129

plate of 30-mm thickness. On the upper right side of the POM plate a small thin metal sheet with

130

thickness of 5 mm was covered. The tailored gap-collector was built up to collect the well-

131

aligned nanofibrous mats of PVA/CNC by adjusting the applied voltage, needle-to-collector

132

distance, working temperature, and the gap distance between two plates. In this case, additional

133

substrates could be placed within the gap to support or collect the nanofibers without any

134

significant influence on the quality of the resultant nanofibers. The as-manufactured fine

135

nanofibers with good alignment were denoted PVA/CNC10A, PVA/CNC25A, and

136

PVA/CNC50A corresponding to the weight ratio of CNC/PVA at 1:9, 1:3, and 1:1, respectively.

137

The as-manufactured nanofibrous mats were vacuum dried at 40 °C for 24 h to remove any

138

residual water, and then vacuum-stored for further analysis.

139

Characterization. Zeta potential of the spinning dope was measured using a dynamic light

140

scattering (DLS, ZS90, Malvern Instruments, U.K.) with a combined laser Doppler velocimetry

141

(LDV) and phase analysis light scattering (PALS). The viscosity of the spinning dope was tested

142

using a RS6000 rheometer (Thermo Scientific, Karlsruhe, Germany) with a parallel plate

143

geometry (25 mm diameter, 0.3 mm gap) at 25 °C. Nanofibrous mats were deposited onto a

144

rectangular glass slide for liquid crystal phase observation on a polarized optical microscope

145

(POM; LV100POL, Nikon, Japan). Morphology observations were carried out on a field

146

emission scanning electron microscope (FE-SEM, Hitachi S-4700 microscope) with an

147

accelerating voltage of 20 kV. The sample mats were fixed on conductive carbon tape and then

148

sputtered with gold for SEM observations. Size distribution of nanofibers was statistically

149

analyzed by using the ImageJ Tool. Fourier-transform infrared spectroscopy (FTIR) of

ACS Paragon Plus Environment

Biomacromolecules

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

150

nanofibrous mats was performed on a Bio-Rad FTS 6000 FTIR spectrometer equipped with an

151

IR microscope (UMA-500); the polarized infrared beam, with rays in parallel, goes through the

152

microtomed sample mats, while the transmittance was measured with the aid of a liquid nitrogen

153

cooled Mercury Cadmium Telluride (MCT) detector. The spectra were recorded at a resolution

154

of 8 cm-1 with a spectral range of 650-4000 cm-1. During the measurement, a home-built sample

155

holder was employed to keep the mats in position at a constant inclination angle θ while a set of

156

different polarizer positions φ was scanned. Tensile properties of the standard rectangular strips

157

(40 × 100 mm2) of nanofibrous mats were performed on a Shimadzu SLBL-500N tensile tester

158

(Shimadzu Inc., Japan) at a crosshead speed of 5 mm/min. The testing results were evaluated as

159

an average of at least 10 measurements.

160

RESULTS AND DISCUSSION

161

Homogeneous composite dispersions of PVA/CNC as spinning dopes were sprayed in a

162

single-fluid manner and collected on a rotating cylindrical roller as shown in Figure S1a.

163

Judging from the fineness, continuity, defect-free, and uniformity of nanofibers, the

164

optimized spinning parameters of nanofibers obtained by roller collector were 10-16 wt%, 12 kV,

165

35 °C, 51%, 0.3 mL/h, 13 cm, and 94.7 mm/min corresponding to dope concentration, voltage,

166

temperature, relative humidity, flow rate, needle-to-collector distance, and nozzle translational

167

velocity, respectively as demonstrated in Table S1 and Figure S2-5. The as-spun winding

168

nanofibers of PVA/CNC10R presents a randomly distributed diameter widely ranging from 300

169

to 600 nm (Figure 1a), indicating the spiral path and whipping of flexible macromolecular

170

chains of PVA.

171

Interestingly, while a tailored gap between double electrode plates was substituted for roller

172

collector (Figure S1b) under voltages of 12 kV, needle-to-collector distance of 11 cm, flow rate,

ACS Paragon Plus Environment

Page 8 of 25

Page 9 of 25

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

Biomacromolecules

173

0.3 mL/h; translational velocities of 28.8 mm/min (Figure S6-7), multiple jets were released and

174

the individual jets simultaneously traveled from one end of collecting electrode plane, and then

175

crossed the gap to the other end of electrode plane whilst undergoing a stretching and

176

disintegration process as shown in Video S1. The generated uniaxially aligned nanofibrous

177

arrays were finally deposited across the gap (7 cm) with the longitudinal fibrous axes

178

perpendicular to the gap plate, and the arrayed fine nanofibers of PVA/CNC10A display mean

179

diameters of 270 ± 25 nm (Figure 1b), much more uniform and smaller than PVA/CNC10R,

180

indicating a pretty strong stretching underwent between the gap plates.

181

Nonetheless, while concentrated CNC dispersion and hydrophilic PVA solution were ejected

182

in a coaxial manner, no arrayed nanofiber mats could be collected on either gap or rollers

183

because of the distinctly different fluidic properties and discordant fiber-formation behavior of

184

individual dope. Specifically, limited by the low viscosity, poor spinnability and inadequate

185

water evaporation of CNC dispersions, PVA@CNC10 with a diameter varied from 288 to 1007

186

nm displays flat and occasionally collapsed surface (Figure 1c). In the case of CNC10@PVA

187

with a wide diameter distribution (98 - 1359 nm), some droplets of CNCs caused by insufficient

188

water evaporation had a strong tendency to pool together with PVA nanothreads after a

189

continuous deposition, thereby continuously aggregating into crosslinked porous fibrous network

190

(Figure 1d). Therefore, morphology and size of four kind of nanofibers could be controlled by

191

the tailored spinning approaches and optimized processing parameters.

192

In general, due to the structure anisotropy of the CNCs,32 its aqueous dispersions at a high

193

concentration always display self-assembled tactoids with fantastic fingerprint textures as shown

194

in Figure 2a. As one component in the coaxial spinning dope, CNC fluid was discontinuously

195

ejected and elongated, thus leading to intermittent shiny birefringent segments evenly distributed

ACS Paragon Plus Environment

Biomacromolecules

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 25

196

inside the long fiber of PVA@CNC10 (Figure 2b), 33 otherwise, reflective iridescent CNC beads

197

(diameter of 0.7-3 µm) were dotted along the long PVA threads (average diameter of 279 nm) to

198

form into regular pearl-chain structure in CNC10@PVA (Figure 2c). When a compensator of

199

530 nm retardation plate was inserted between the polarizer and analyzer at a fixed angle of 45°,

200

nematic rather than cholesteric mesophases of CNCs were randomly distributed in the core of

201

PVA@CNC10 or at the outer surface of CNC10@PVA (Figure S8), indicating a mesophase

202

transition from the chiral twisting lamella to nematic phases of CNCs.

203

In the single-fluidic spinning dope, no mesophase of CNCs was recognized due to the

204

homogeneous and isotropic dispersion of nanorods (Figure 2d). However, under polarized light,

205

shining overlaid birefringence could be visualized in the randomly-distributed nanofibers of

206

PVA/CNC10R (Figure 2e). Moreover, frequent and continuous birefringence with selective

207

reflection was obviously visualized and regarded as nematic CNCs in the PVA/CNC10A (Figure

208

2f). Micrographs of nanofibrous mats in the presence of a 530-nm retardation plate between

209

crossed polarizers are shown in Figure 3. By changing the rotation angle of the sample,

210

continuous transition of alternating blue and orange bands were visualized in the arrayed

211

nanofiberous mats of PVA/CNC10A (Figure 3e,f) and PVA/CNC50A (Figure 3h,i).

212

Accordingly, the fuchsia-colored image regions corresponded to uniaxially fibrous orientation (n)

213

roughly parallel to the slow axis (γ) of the retardation plate whereas the orange-red hue

214

corresponded to n⊥γ, indicating high refractive index of the axis (γ) or the uniaxial fiber axis due

215

to positive birefringence anisotropy (ne > no) of CNCs.34 Specifically, PVA/CNC50A exhibited a

216

relatively strong coloration difference between longitudinal and transverse fiber axis, suggesting

217

more uniform and strong orientation along uniaxial fiber axis as a result of the intra-particle

218

repulsion of surface-charged nanorods and stretching-induced alignment under electric field

ACS Paragon Plus Environment

Page 11 of 25

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

Biomacromolecules

219

(Figure 3h,i).35 In contrast, as no accordance of the γ axis to the fibrous axis, randomly-

220

distributed PVA/CNC10R did not present overall arrangement and mesophase orientation of

221

CNCs but a tilt with respect to the longitudinal axis (Figure 3a,b).

222

Polarized FTIR spectra was deployed to verify the particle orientation in the nanofibers

223

(Figure S9). The characteristic band centered at 1160 cm-1 is associated with stretching and

224

asymmetric pyranose ring breathing vibrations of cellulose; the peak at around 1100 cm-1 is

225

assigned to C-O stretching vibrations of PVA;36 and the band at around 1060 cm-1 is attributed to

226

stretching vibration of C-OH of cellulose. Polar plots of absorbance of ν

227

(1100 cm-1), ν

228

three bands of PVA/CNC10R have very few response to the rotational angle of incidence

229

polarized light from 0° to 360° and present a circular shape, indicating a total isotropic character

230

under polarized infrared light (Figure 4a). However, as evidenced by the ellipsoid or dumbbell

231

shape, polar plots of IR bands in the aligned nanofibrous arrays exhibits a significant response to

232

the polarization angle with the maximum difference occurring between absorbance of ν C-OH or ν

233

C-O-C at

234

indicator of the degree of uniaxial orientation of CNCs in the mats, dichroic ratio (R) was

235

calculated according to the following equation.36

236

R=

237

where A// and A⊥ are IR absorbance at parallel and perpendicular mode, respectively. R of the

238

randomly-distributed nanofiber of PVA/CNC10R, as listed in Table 1, is nearly equal to 1 due to

239

the isotropic character. In the case of a perfectly uniaxial orientation along the fiber axis, R

240

should be theoretically equal to infinity.37 In the arrayed fibers, R of ν

241

was high up to 1.35-1.58 or 1.52-2.09, respectively, suggesting that the orientation of CNCs

A //

C-O-C

C-OH

(1060 cm-1), ν

C-O

(1160 cm-1) of spun nanofibers are shown in Figure 4. The peak value of

parallel mode (A//) and perpendicular mode (A⊥) to the fiber axes, (Figure 4b,c,d). As an

(1)

A⊥

ACS Paragon Plus Environment

C-OH

and ν

C-O-C of

CNCs

Biomacromolecules

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

242

rather than PVA occurred along the fiber axis.38 Therefore, in the gap region, the stretching and

243

the Coulombic force possibly synergistically induced the orientation of charged nanorods and

244

initiated the nematic phase formation in the jet.39,40

Page 12 of 25

245

Typical stress-strain profiles of 4 nanofiber mats are shown in Figure 5, and the resultant

246

averaged stress, stain, and elastic modulus are listed in Table 1. Reinforcement effects of a

247

certain amount of CNCs (10 wt%) are distinguished, e.g., mean tensile strength of

248

PVA/CNC10R and PVA/CNC10A were 8.27, and 10.25 MPa, respectively, much higher than

249

that of PVA (7.50 MPa); meanwhile, the modulus of PVA/CNC10R, PVA/CNC10A,

250

CNC10@PVA, and PVA@CNC10 were all much higher than that of PVA since the rigidity of

251

CNCs played a crucial role in the enhancement of tensile strength. However, fibrous mats with

252

high loading doses of CNCs (≥ 25 wt%) exhibit a sharply decreasing stress and strain in

253

comparison to PVA/CNC10. Meanwhile, tensile strength and elongation at break of

254

CNC10@PVA are higher than that of PVA@CNC10 due to the crosslinking network effects of

255

CNCs as depicted in Figure 1d. That is to say, interface contact and evaporation rate of dope

256

components are dependent factors that influence the structure, morphology and mechanical

257

properties of coaxial nanofibrous mats. Interestingly, the aligned nanofibers of PVA/CNC at the

258

same loading doses of CNCs exhibit a higher tensile stress than that of randomly distributed

259

nanofibers along the fibrous axis because both good alignment of nanofibers and continuous

260

nematic segments of CNCs in aligned nanofibers provided a substantially strong resistance to

261

external stretching force. However, the increment in tensile strength of the aligned nanofibers of

262

PVA/CNC didn’t occur at the expense of tensile strain but exhibit a higher elongation at break

263

than that of randomly distributed nanofibers at the same loading doses of CNCs. This is ascribed

264

to the fact that under uniaxial drawing along fiber axis the nematic mesophases of CNCs were re-

ACS Paragon Plus Environment

Page 13 of 25

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

Biomacromolecules

265

orientated, thus leading to an extended deformation and improved flexibility.28 The aligned

266

nanofibrous mats is somewhat analogy to liquid crystal elastomers with macroscopic elastic

267

response arising from the phase separation under large deformations.

268

CONCLUSIONS

269

In this work, CNC dispersion and aqueous PVA solution as a mixed spinning dope or

270

separated dopes were sprayed in a manner of single-fluid or coaxial electrospinning and

271

collected on a rotating cylindrical roller or across a gap. Without any doubt, fine nanofibers

272

generated from single fluidic dope was continuous, smooth, and uniform due to homogenous

273

dispersion of CNC into PVA, whereas coaxial nanofiber had a larger average diameter, broader

274

size distribution and rough surface. The single-fluidic jet deposited across gap plates were well

275

aligned along the direction vertical to plates because the gap was favorable for the electric-field-

276

induced-alignment of CNCs by the uniaxial stretching as well as the movement of CNCs

277

unrestricted by the length of nanofibers as far as possible. The anisotropic and birefringent

278

character of fiber arrays oriented along the uniaxial fiber axis were certified by polarized IR and

279

POM with a retardation plate. The alignment of nanofibers was highly dependent on the

280

processing condition, and the orientation of CNCs with positive refractive index anisotropy and

281

surface charges was related to the high voltage electric field. Being different with single-fluidic

282

jet, coaxially-electrospun core-sheath nanofibers also showed birefringence under polarized light

283

because nematic mesophases of CNCs in the concentrated dispersions was switched. The

284

uniaxial, continuous orientation of CNCs played an important role in enhancement of mechanical

285

properties of the nanofibrous composite mats, e.g., average tensile strength (10.25 MPa) and

286

elongation at break (87.91%) of PVA/CNC10A were much higher than that of other spun

287

nanofibrous mats. In this work, we have provided a strategy for encapsulating homogeneously

ACS Paragon Plus Environment

Biomacromolecules

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

288

nematic nanorods into nanoscaled fibrous mats which is anticipated to be explored as novel

289

photonic and biological tissues, biosensors, and environmental-friendly filters.

290

Supporting Information Available. Diagram of electrospinning set-up, properties of

291

spinning dope, optimization for randomly-distributed and aligned electrospun nanofibers, POM

292

of CNCs in the core-sheath nanofibers, polarized FTIR spectra of nanofibers. This material is

293

available free of charge via the Internet at http://pubs.acs.org.

294

AUTHOR INFORMATION:

295

Corresponding author

296



297

ORCID

298

Dagang Liu: 0000-0002-1320-7030

299

Notes

300

The authors declare no competing financial interest.

301

ACKNOWLEDGEMENTS

302

The authors are grateful to National Natural Science Foundation of China (Nos. 51473077 and

303

21277073), CSC scholarship (201608320064), and Six Talents Summit Program and 333 High-

304

Level Talent Cultivation Program of Jiangsu Province for financial support.

305

REFERENCES

E-mail: [email protected] (D. Liu)

306

(1) Antaya, H.; Richardlacroix, M.; Pellerin, C. Macromolecules 2010, 43, 4986-4990.

307

(2) Yao, L.; Haas, T. W.; Guiseppi-Elie, A.; Bowlin, G. L.; Simpson, D. G.; Wnek, G. E. Chem.

308 309

Mater. 2003, 15, 1860-1864. (3) Li, D.; Babel, A.; Jenekhe, S.; Xia, Y. Adv. Mater. 2004, 16, 2062–2066.

ACS Paragon Plus Environment

Page 14 of 25

Page 15 of 25

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

Biomacromolecules

310

(4) Ma, Z.; Ramakrishna, S. J. Membrane Sci. 2008, 319, 23-28.

311

(5) Zhang, H. T.; Nie, H. L.; Yu, D. G.; Wu, C. Y.; Zhang, Y. L.; White, C. J. B.; Zhu, L.

312

Desalination 2010, 256, 141-147.

313

(6) Makaremi, M.; Silva, R. T. D.; Pasbakhsh, P. J. Phys. Chem. C 2015, 119, 7949-7958.

314

(7) Stephens, J. S.; Chase, D. B.; Rabolt, J. F. Macromolecules 2004, 37, 877-881.

315

(8) Luo, Y.; Shen, H.; Fang, Y.; Cao, Y.; Huang, J.; Zhang, M.; Dai, J.; Shi, X.; Zhang, Z. J. ACS

316 317

Appl. Mater. Interface 2015, 7, 6331. (9) Unnithan, A. R.; Barakat, N. A.; Pichiah, P. B.; Gnanasekaran, G.; Nirmala, R.; Cha, Y. S.;

318

Jung, C. H.; EI-Newehy, M.; Kim, H. Y. Carbohydr. Polym. 2012, 90, 1786-93.

319

(10) Xu, L.; Dong, B.; Wang, Y.; Bai, X.; Chen, J. S.; Liu, Q.; Song, H. W. J. Phys. Chem. C

320

2010, 114, 9089-9095.

321

(11) Zong, X. H.; Kim, K.; Fang, D. F.; Ran, S. F.; Hsiao, B. S.; Chu, B. Polymer 2002, 43,

322

4403-4412.

323

(12) C. Wang, H. S. Chien, K. W. Yan, C. L. Hung, K. L. Hung, S. J. Tsai, H. J. Jhang, Polymer

324

2009, 50, 6100-6010.

325

(13) Gu, S. Y.; Ren, J.; Vancso, G. J. Eur. Polym. J. 2005, 41, 2559-2568.

326

(14) Fan, Z.; Ho, J. C.; Jacobson, Z. A.; Yerushalmi, R.; Alley, R. L.; Razavi, H.; Javey, A. Nano

327

Lett. 2008, 8, 20-25.

328

(15) Dersch, R.; Liu, T. Q.; Schaper, A. K.; Greiner, A.; Wendorff, J. H.; J. Polym. Sci. A. Polym.

329

Chem. 2003, 41, 545–553.

330

(16) Xie, J.W.; Macewan, M. R.; Ray, W. Z.; Liu, W. Y.; Siewe, D. Y.; Xia, Y. N. ACS Nano

331

2010, 4, 5027.

ACS Paragon Plus Environment

Biomacromolecules

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 25

332

(17) Sundaray, B.; Subramanian, V.; Natarajan, T. S.; Xiang, R. Z.; Chang, C. C.; Fann, W. S.

333

Appl. Phys. Lett. 2004, 84, 1222-1224.

334

(18) Katta, P.; Alessandro, M.; Ramsier, R. D.; Chase, G. G. Nano Lett. 2004, 4, 2215-2218.

335

(19) Marchessault, R. H.; Morehead, F. F.; Koch, M. J. J. Colloid. Sci. 1961, 16, 327-344.

336

(20) Beck-Candanedo, S.; Roman, M.; Gray, D.G. Biomacromolecules 2005, 6, 1048-1054.

337

(21) Revol, J. F.; Godbout, L.; Dong, X. M.; Gray, D. G.; Chanzy, H.; Maret, G. Liq. Cryst. 1994,

338

16, 127-134.

339

(22) Liu, D.; Chen, X.; Yue, Y.; Chen, M.; Wu, Q. Carbohydr. Polym. 2011, 84, 316-322.

340

(23) Liu, D.; Wang, S.; Ma, Z.; Tian, D.; Gu, M.; Lin, F. RSC Adv. 2014, 4, 39322-39331.

341

(24) Shopsowitz, K. E.; Qi, H.; Hamad, W. Y.; MacLachlan, M. J. Nature 2010, 468, 422-428.

342

(25) Shopsowitz, K. E.; Hamad, W. Y.; MacLachlan, M. J. J. Am. Chem. Soc. 2012, 134,

343

867−870.

344

(26) Gu, M.; Jiang, C.; Liu, D.; Prempeh, N.; Smalyukh, I. I. ACS Appl. Mater. Interface 2016, 8,

345

32565-32573.

346

(27) Zhou, C.; Chu, R.; Wu, R.; Wu, Q. Biomacromolecules 2011, 12, 2617–2625.

347

(28) Liu, D.; Li, J.; Sun, F.; Xiao, R.; Guo, Y.; Song, J. RSC Adv. 2014, 4, 30784-30789.

348

(29) Lee, J.; Deng, Y. Macromol. Res. 2012, 20, 76-83.

349

(30) Peresin, M. S.; Habibi, Y.; Zoppe, J. O.; Pawlak, J. J.; Rojas, O. J. Biomacromolecules 2010,

350

11, 674–681.

351

(31) Huan, S. Q.; Bai, L.; Cheng, W. L.; Han, G. P. Polymer 2016, 92, 25-35.

352

(32) Csoka, L.; Hoeger, I. C.; Peralta, P.; Peszlen, I.; Rojas, O. J. J. Colloid Interface Sci. 2011,

353

363, 206–212.

354

(33) Enz, E.; Lagerwall, J. J. Mater. Chem. 2010, 20, 6866–6872.

ACS Paragon Plus Environment

Page 17 of 25

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

Biomacromolecules

355

(34) Liu, Q.; Campbell, M. G.; Evans, J. S.; Smalyukh, I. I. Adv. Mater. 2014, 26, 7178-7184.

356

(35) Llanos, G. R.; Sefton, M. V. Macromolecules 1991, 24, 6065-6072.

357

(36) Pedicini, A.; Farris, R. J. Polymer 2003, 44, 6857-6862.

358

(37) Haridas, M.; Smalyukh, I. I. Small 2015, 11, 5572–5580.

359

(38) Kongkhlang, T.; Tashiro, K.; Kotaki, M. Chirachanchai, S. J. Am. Chem. Soc. 2008, 130,

360

15460–15466.

361

(39) Nakashima, K.; Tsuboi, K.; Matsumoto, H.; Ishige, R.; Tokita, M.; Watanabe, J.; Tanioka, A.

362

Macromol. Rapid Comm. 2010, 31, 1641–1645.

363

(40) Herrera, N. V.; Mathew, A. P.; Wang, L. Y.; Oksman, K. Plastics, Rubber and Composites.

364

2011, 40, 57-64.

365

ACS Paragon Plus Environment

Biomacromolecules

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

366

Page 18 of 25

Table 1. Dichroic ratio (R) and mechanical properties of as-spun nanofibrous mats. Nanofibrous

R

R

mats

(ν C-OH)

PVA

-

-

PVA/CNC10R

0.94

PVA/CNC10A

R

Tensile

Elongation at

Elastic

stress (Pa)

break (%)

modulus (MPa)

-

7.50±0.43

77.00±7.22

38.01±2.32

1.03

0.96

8.27±0.25

20.17±4.23

853.15±14.13

1.40

1.06

1.52

10.25±0.34

87.91±7.34

185.33±4.17

PVA/CNC25R

-

-

-

4.03±0.24

3.13±0.81

430.79±7.15

PVA/CNC25A

1.58

1.08

1.86

6.48±0.13

29.51±3.33

342.72±6.8

PVA/CNC50R

-

-

-

0.40±0.032

2.86±0.31

16.96±2.02

PVA/CNC50A

1.35

1.11

2.09

2.79±0.02

8.12±0.92

236.39±5.32

CNC10@PVA

-

-

-

8.82±0.26 109.19±10.73 115.29±1.71

PVA@CNC10

-

-

-

6.84±0.32

(ν C-O) (ν C-O-C)

37.53±6.28

367

ACS Paragon Plus Environment

837.12±13.76

Page 19 of 25

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

Biomacromolecules

368

Figure Captions:

369

Figure 1. SEM micrographs and histograms of width distributions for randomly orientated

370

nanofibers of PVA/CNC10R (a), well-aligned nanofibers of PVA/CNC10A (b), and core-sheath

371

nanofibers of PVA@CNC10 (c), CNC10@PVA (d), respectively.

372

Figure 2. Polarized optical micrographs of spinning dope of CNC (a) and PVA/CNC10 (d), and

373

electrospun nanofibers of PVA@CNC10 (b), CNC10@PVA (c), PVA/CNC10R (e),

374

PVA/CNC10A (f), respectively.

375

Figure 3. POM images of nanofibrous mats of PVA/CNC10R (a), PVA/CNC10A (d),

376

PVA/CNC50A (g), and POM images of nanofibrous mats of PVA/CNC10R (b,c),

377

PVA/CNC10A (e,f), PVA/CNC50A (h,i) with a full-wavelength (530 nm) retardation plate with

378

a slow axis (γ) marked by a yellow arrow.

379

Figure 4. Polar plots of the normalized absorbance as a function of rotational angle (α) for ν C-OH

380

(1060 cm-1), ν

381

PVA/CNC25A, (d) PVA/CNC50A.

382

Figure 5. Typical stress-strain profiles of PVA/CNC10R, PVA/CNC25R, PVA/CNC50R (a);

383

PVA/CNC10A,

384

PVA@CNC10 and CNC10@PVA fiber mats (c). The tensile test was performed along the fiber

385

axis.

C-O (1100

cm-1), ν

C-O-C (1160

PVA/CNC25A,

cm-1) of (a) PVA/CNC10R, (b) PVA/CNC10A, (c)

PVA/CNC50A (b);

and

ACS Paragon Plus Environment

coaxial

nanofiber mats

of

Biomacromolecules

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 25

386

387 388

Figure 1. SEM micrographs and histograms of width distributions for randomly orientated

389

nanofibers of PVA/CNC10R (a), well-aligned nanofibers of PVA/CNC10A (b), and core-sheath

390

nanofibers of PVA@CNC10 (c), CNC10@PVA (d), respectively.

ACS Paragon Plus Environment

Page 21 of 25

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

Biomacromolecules

391 392

Figure 2. Polarized optical micrographs of spinning dope of CNC (a) and PVA/CNC10 (d), and

393

electrospun nanofibers of PVA@CNC10 (b), CNC10@PVA (c), PVA/CNC10R (e),

394

PVA/CNC10A (f), respectively.

ACS Paragon Plus Environment

Biomacromolecules

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 25

395 396

Figure 3. POM images of nanofibrous mats of PVA/CNC10R (a), PVA/CNC10A (d),

397

PVA/CNC50A (g), and POM images of nanofibrous mats of PVA/CNC10R (b,c),

398

PVA/CNC10A (e,f), PVA/CNC50A (h,i) with a full-wavelength (530 nm) retardation plate with

399

a slow axis (γ) marked by a yellow arrow.

ACS Paragon Plus Environment

Page 23 of 25

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

Biomacromolecules

400 401

Figure 4. Polar plots of the normalized absorbance as a function of rotational angle (α) for ν C-OH

402

(1060 cm-1), ν

403

PVA/CNC25A, (d) PVA/CNC50A.

C-O (1100

cm-1), ν

C-O-C (1160

cm-1) of (a) PVA/CNC10R, (b) PVA/CNC10A, (c)

ACS Paragon Plus Environment

Biomacromolecules

a 12

b12

PVA/CNC10R PVA/CNC25R PVA/CNC50R

10

8

8

8

6

6

6

4

4

4

2

2

2

0.06 0.04 0.02 0.00

0.06 0.04 0.02 0.00

0.06 0.04 0.02 0.00

0

5

10

15

20

25

Strain (%)

404

c 12

PVA/CNC10A PVA/CNC25A PVA/CNC50A

10

Stress (MPa)

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

Page 24 of 25

0

20

40

60

80

PVA@CNC10 CNC10@PVA

10

100

0

Strain (%)

20

40

60

80 100 120

Strain (%)

405

Figure 5. Typical stress-strain profiles of PVA/CNC10R, PVA/CNC25R, PVA/CNC50R (a);

406

PVA/CNC10A,

407

PVA@CNC10 and CNC10@PVA fiber mats (c). The tensile test was performed along the fiber

408

axis.

PVA/CNC25A,

PVA/CNC50A (b);

and

409

ACS Paragon Plus Environment

coaxial

nanofiber mats

of

Page 25 of 25

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

Biomacromolecules

410

For Table of Contents Use Only

411

Graphical Abstract:

412 413

414

415

Fiber Alignment and Liquid Crystal Orientation of Cellulose Nanocrystals in the

416

Electrospun Nanofibrous Mats

417

Weiguang Song,a,b Dagang Liu,*a,b,c Nana Prempeh,b Renjie Song a

ACS Paragon Plus Environment