Strong and tough chitin film from α-chitin nanofibers prepared by high

Dec 7, 2018 - Chitin film prepared from the chitin nanofiber suspension by vacuum filtration exhibited a true ... MJ/m3), much higher than those chiti...
1 downloads 0 Views 3MB Size
Subscriber access provided by TULANE UNIVERSITY

Article

Strong and tough chitin film from #-chitin nanofibers prepared by high pressure homogenization and chitosan addition NGESA EZEKIEL MUSHI, Takashi Nishino, Lars A. Berglund, and Qi Zhou ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.8b05452 • Publication Date (Web): 07 Dec 2018 Downloaded from http://pubs.acs.org on December 11, 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 24 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 Sustainable Chemistry & Engineering

1

Strong and tough chitin film from -chitin

2

nanofibers prepared by high pressure

3

homogenization and chitosan addition

4

Ngesa Ezekiel Mushi,a,b Takashi Nishino,c Lars A Berglund,a Qi Zhoua,d,*

5

a

6

Engineering Sciences in Chemistry, Biotechnology and Health, KTH Royal Institute of

7

Technology, SE-100 44 Stockholm, Sweden

8

b

9

Technology, University of Dar es Salaam, P.O. Box 35131, Dar es Salaam, Tanzania

Wallenberg Wood Science Center, Department of Fiber and Polymer Technology, School of

Department of Mechanical and Industrial Engineering, College of Engineering and

10

c

Department of Chemical Science and Engineering, Kobe University, Kobe 657-8501, Japan

11

d

Division of Glycoscience, Department of Chemistry, School of Engineering Sciences in

12

Chemistry, Biotechnology and Health, KTH Royal Institute of Technology, AlbaNova

13

University Centre, SE-106 91 Stockholm, Sweden

14

KEYWORDS. chitin nanofibers, chitin film, chitosan, mechanical, nanostructure.

15

ACS Paragon Plus Environment

1

ACS Sustainable Chemistry & Engineering 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 24

16

17

ABSTRACT.

18

Chitin nanofibers is an interesting biological nanomaterial for advanced applications, for

19

example, in medicine, electronics, packaging and water purification. The challenge is to

20

separate chitin nanofibers from protein in the exoskeleton structure of arthropods and avoid

21

nanofibril aggregation to realize the mechanical potential of chitin. In this work, we

22

developed a new method for the preparation of chitin nanofibers from lobster shell

23

exoskeleton using 10 wt.% chitosan as a sacrificial polymer. The addition of chitosan in the

24

raw chitin colloidal suspension during high pressure homogenization process at pH 3

25

significantly reduced the agglomeration of chitin nanofibers as revealed by dynamic light

26

scattering and transmission electron microscopy. Chitin film prepared from the chitin

27

nanofiber suspension by vacuum filtration exhibited a true nanofibrils network structure

28

without fibril aggregations as characterized by scanning electron microscopy. The presence

29

of chitosan not only improves the colloidal stability of chitin nanofibers suspension but also

30

facilitates the formation of chitin nanofiber network structure in the film as indicated by wide

31

angle X-ray diffraction analysis. The chitin nanofiber film with 4 ± 1 wt.% residual chitosan

32

showed high tensile strength (187.2  5.6 MPa) and high work of fracture (12.1  0.4

33

MJ/m3), much higher than those chitin and chitosan films reported previously in literature.

34

ACS Paragon Plus Environment

2

Page 3 of 24 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 Sustainable Chemistry & Engineering

35

Introduction

36

Chitin is an important building block in load-bearing exoskeleton structures in arthropods and

37

the second most abundant natural biopolymer with vast mechanical potential for structural

38

applications in nanotechnology. Chitin nanofibers have been prepared from -chitin

39

resources such as the exoskeletons of crustaceans (e.g. crabs, lobsters, and prawns) and

40

fungal cell walls (e.g. mushroom) by chemical pretreatment combined with mechanical

41

homogenization.1–4 The preparation of chitin nanofibers of high aspect ratio (length to width

42

ratio) from the less abundant -chitin resources such as squid pen and tubeworm has been

43

achieved by simple mechanical treatment under acid conditions, which is easier compared to

44

-chitin.5–6 This is possibly because there is no inter-sheet hydrogen bonds in -chitin and

45

polar solvents such as water and alcohol are able to penetrate into -chitin.7 In addition, -

46

chitin forms strong complexes with proteins through histidyl and aspartyl residues in nature,8

47

which makes it more difficult to extract pristine -chitin nanofibers. Moreover, the nature of

48

chitin agglomeration in water suspensions after extraction is a major obstacle for enhancing

49

colloidal stability of -chitin nanofibers. Agglomeration may compromise mechanical

50

performance of -chitin nanofiber network structures.

51

In our previous work, crude chitin powder from crab shells was treated with 2 M HCl for 2

52

days at room temperature, then boiled in 8 wt% NaOH for 2 days, and finally refluxed in

53

ethanol for 6 h to remove minerals, protein, and pigments, respectively. -Chitin nanofibers

54

with 7.0 wt% residual proteins were prepared by mechanical homogenization.9 Films

55

prepared from such chitin-protein composite fibers showed a tensile strength of 77  5.6

56

MPa. In addition, we have developed a mild process by performing all the chemical

57

pretreatments at room temperature.10 -Chitin nanofibers with a very small diameter (3.6 –

58

3.9 nm) and lower residual proteins content (4.7 wt%) were successfully prepared. Hydrogel

ACS Paragon Plus Environment

3

ACS Sustainable Chemistry & Engineering 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 24

59

prepared from such nanofibers showed a storage modulus of 13 kPa and a compression

60

modulus of 309 kPa at 2 wt.% chitin content, the highest reported for chitin hydrogels.11

61

Films prepared from such nanofibers demonstrated much higher transparency and enhanced

62

mechanical properties (tensile strength of 153  10.6 MPa). However, aggregates of chitin

63

nanofibers were still observed in the film. The mechanical potential of chitin was still not

64

fully realized. The axial elastic modulus of -chitin crystals is estimated at 41 GPa as

65

measured by an X-ray diffractometer equipped with a stretching device and a load cell.12 The

66

strength of single -chitin nanofibrils is 1.6 GPa as estimated via sonication-induced

67

fragmentation.6 A key challenge is to improve the individualization of chitin nanofibers and

68

reduce the presence of chitin/protein aggregates, and at the same time preserve the native

69

structure of chitin.

70

Chitosan is a deacetylated chitin derivative and a cationic polyelectrolyte with good colloidal

71

stability in water. The cationic polyelectrolyte properties, the biocompatibility, the linear

72

chain structure and solubility in aqueous system have made chitosan a very attractive

73

polymer. Fan et al.13 and Ifuku et al.14 performed surface deacetylation on chitin nanofibers

74

and produced chitin films by solution casting, reporting a tensile strength of 140 ± 48 MPa

75

and 156.5 ± 10.0 MPa, respectively. Previously, we discovered that nanocomposites of chitin

76

nanofibers and chitosan have a unique combination of modulus, strength and strain-to-failure

77

owing to reorientation, slippage and straightening of chitin nanofibers during deformation of

78

the ductile chitosan matrix.15 Nanocomposites with a chitosan content of 30 wt.% showed

79

considerable strength (140 MPa) and strain-to-failure (11%) due to good dispersion of chitin

80

nanofibers in the chitosan matrix. In the present work, chitosan was utilized as a sacrificial

81

polymer for improving the individualization of chitin nanofibers during mechanical

82

disintegration by increasing electrostatic repulsive forces between chitin nanofibers. The

83

advantage of this method is that it eliminates the need of direct chemical surface

ACS Paragon Plus Environment

4

Page 5 of 24 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 Sustainable Chemistry & Engineering

84

deacetylation of chitin nanofibers and avoids the associated hydrolysis of chitin. The

85

structure and colloidal stability of the chitin nanofibers were characterized by electron

86

microscopy and dynamic light scattering. Chitin films was prepared by vacuum filtration of

87

the water suspension of chitin nanofibers. The structure and mechanical properties of the

88

chitin films were studied and compared with chitin nanofibers prepared without the addition

89

of chitosan.

90 91

Experimental section

92

Materials. Chitin was extracted from lobster shell (Homarus Americanus of Northwest

93

Atlantic, Canada) by removing minerals, pigments and proteins according to our previous

94

work.10, 15 After the treatments of 2 M HCl and ethanol, the exoskeleton of lobster was treated

95

with 20 wt% NaOH for 2 weeks at room temperature. The degree of acetylation (DA) of the

96

extracted chitin was ca. 87% as measured by solid state

97

(NMR) spectroscopy. Chitosan powder from shrimp (high viscous, Sigma, Germany) with a

98

degree of acetylation of less than 15% was purchased from Sigma-Aldrich.

99

Preparation of chitin nanofibers. Chitosan was first dissolved in 4% acetic acid. The

100

extracted whitish chitin slurry was also suspended in 4% acetic acid with the addition of the

101

chitosan solution. The total solid content and the solid content of chitosan were 1 wt.% and

102

0.1 wt.% in the suspension, respectively. Thus, the initial content of chitosan was 10 wt.% in

103

the chitin/chitosan mixture. The suspension was mixed in a kitchen blender (Vita-Prep® 3,

104

Vita-Mix Corporation, USA) for 5 min and then passed through a Microfluidizer (M-110EH,

105

Microfluidics Ind., Newton, MA, USA) to obtain a chitin nanofiber hydrocolloidal

106

suspension. The suspension was first passed five times through the 400 and 200 μm

107

microfluidic chambers at a pressure of 900 bar and then five times through the 200 and 100

13C

ACS Paragon Plus Environment

Nuclear Magnetic Resonance

5

ACS Sustainable Chemistry & Engineering 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 24

108

μm chambers at a pressure of 1600 bar. This sample was coded as ChNF-c. The control

109

sample was prepared in the same manner without the addition of the chitosan solution and

110

coded as ChNF.

111

Preparation of chitin nanofiber film. The chitin nanofiber suspension was diluted to 0.1

112

wt.% using an Ultra Turrax Mixer (IKA, D125 Basic, USA) at a speed of 1200 rpm for 10

113

min to allow uniform dispersion and complete disentanglement of the nanofiber aggregates.

114

The suspension was then vacuum filtrated using a 0.65 μm pore size filter membrane (DVPP,

115

Millipore, USA) to form a wet cake with water content of ca. 90 %. The wet cake was dried

116

in the drying stage of a Rapid Köthen sheet former (Germany) at 70 mbar and 93 °C for 10

117

min to obtain a chitin nanofiber film. The film was further dried to constant weight at 105 oC.

118

The weight of the residue chitosan in the film was measured from the weight of the chitin

119

nanofiber/chitosan mixture before filtration to the weight of the film after filtration and

120

drying. The initial solid weight of the sample was based on the solid content of the colloidal

121

suspension (0.1 wt.%) and total weight of the suspension before filtration. The weight loss of

122

chitin nanofiber after filtration has not been observed for neat ChNF. Chitosan is completely

123

filtered through the filter membrane when neat chitosan solution was applied. Five films were

124

prepared for ChNC-c to measure the residual content of chitosan.

125

Characterization. Transmission electron microscopy (TEM) of chitin nanofibers was

126

performed using the Hitachi Model HT7700 transmission electron microscope operated in

127

high-contrast mode at 100 kV. Chitin nanofiber sample was deposited on a carbon coated

128

copper grid (Ultra-thin Carbon Type-A, Ted pella) and stained with 1% uranyl acetate.

129

The charge and size of the chitin nanofiber hydrocolloids were studied by dynamic light

130

scattering (DLS) using Zetasizer Nano (Model ZEN3600, Malvern Instruments Ltd, UK).

131

The light source was operated at a wavelength of 633 nm. Chitin nanofiber suspension was

ACS Paragon Plus Environment

6

Page 7 of 24 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 Sustainable Chemistry & Engineering

132

diluted to a concentration of 0.1 mg/mL at pH 3 and scanned three times in a Poly (methyl

133

methacrylate) cuvette at 21 °C. The size of chitin nanofiber aggregates was determined based

134

on Smoluchowski's approximations.

135

The topographical and cross-sectional morphology of the chitin nanofiber films was

136

characterized by using field emission scanning electron microscopy (FE-SEM, Hitachi S-

137

4800, Japan). The samples were conditioned in a desiccator overnight and then sputtered with

138

a thin layer of gold/palladium using Agar HR sputter coater (Cressington scientific

139

instruments ltd, UK). FE-SEM images were captured from secondary electrons at an

140

acceleration voltage of 1.0 kV.

141

To measure the bulk density of the chitin nanofiber films, rectangular samples with edge size

142

ranging between 1 and 2 cm were accurately weighed. Bulk volume was measured using

143

Mercury Intrusion Porosimeter (Micrometrics, USA). The samples were placed in a

144

penetrometer chamber. The air was evacuated and filled with mercury at atmospheric

145

pressure. The bulk volume was obtained from the difference between volume of the

146

penetrometer with and without the sample. Porosity was calculated from measured bulk

147

density of the sample using the following equation,

148

𝑃𝑜𝑟𝑜𝑠𝑖𝑡𝑦 = 1 ― 𝜌𝑐ℎ

149

where b corresponds to the bulk density of the film and ch corresponds to the density for

150

chitin, which is assumed to be 1.425 g/cm3.16

151

Tensile test of the chitin nanofiber films was performed using a universal tensile testing

152

machine (Model 5944, Instron, UK) equipped with a 500 N load cell at a strain rate of 4

153

mm/min. The specimens with a width of 5 mm and a length of 40 mm were preconditioned at

𝜌𝑏

(1)

ACS Paragon Plus Environment

7

ACS Sustainable Chemistry & Engineering 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 24

154

50% relative humidity and 23 °C overnight. Mechanical properties such as modulus, tensile

155

strength, and strain to failure were obtained based on conventional analysis of tensile stress-

156

strain curves. Toughness is defined as work to fracture and is calculated as the area under the

157

stress−strain curve.17

158

To study the orientation of the chitin nanofibers in the film before and after tensile stretching,

159

wide angle X-ray diffraction (WAXD) measurement was performed. The samples were

160

irradiated by Cu K radiation in the direction both perpendicular (in-the-plane) and parallel

161

(cross-section) to the film surface and the diffraction photographs were recorded. From

162

azimuthal intensity distribution graphs for the 110 equatorial reflection, the degree of

163

orientation () was calculated according to the following equation,

164

𝛱=

165

where FWHM is the full width at half-maximum.

180 ― 𝐹𝑊𝐻𝑀 180

(2)

166 167

Results and discussion

168

Structure of chitin nanofibers

169

The morphology of the chitin nanofibers (ChNFs) prepared using microfluidizer was directly

170

characterized by TEM analysis with negative staining using uranyl acetate. As shown Figure

171

1a, ChNFs with a width of ca. 4–6 nm can be individualized by mechanical homogenization

172

without the addition of chitosan. However, large aggregates of ChNFs or chitin/protein

173

bundles with a width of ca. 20–100 nm are also found in the sample. This is probably because

174

individualized ChNF tends to form aggregates together with the residual protein. It could also

175

be due to incomplete fibrillation of chitin-protein fibril bundles during the mechanical

ACS Paragon Plus Environment

8

Page 9 of 24 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 Sustainable Chemistry & Engineering

176

treatment. In our previous work,10 such aggregates were further disintegrated by

177

ultrasonication in very dilute solution (0.005 wt.%) before structural analysis. A fiber width

178

of ca. 3.6 nm was measured by AFM. However, nanofiber aggregates formed again after the

179

suspension was stored overnight. Interestingly, with the addition of chitosan to the extracted

180

lobster chitin, ChNFs with a uniform width of ca. 4–6 nm was completely individualized after

181

homogenization, see Figure 1b. This suggests that the addition of chitosan during

182

homogenization facilitates the fibrillation of the extracted lobster chitin fibrils, and prevent

183

the aggregation of individualized ChNFs. Such colloidal nanofiber suspension is stable upon

184

storage for a few weeks.

185

ACS Paragon Plus Environment

9

ACS Sustainable Chemistry & Engineering 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 24

186

Figure 1. Typical transmission electron microscopy (TEM) images of chitin nanofibers from

187

lobster shell (a) disintegrated without the addition of chitosan and (b) disintegrated in the

188

presence of chitosan in suspension.

189

190 191

Figure 2. Estimated size distributions of chitin nanofibers in water suspensions for neat chitin

192

(ChNF) and chitin with the addition of 10 wt.% chitosan (ChNF-c) in suspension during

193

mechanical homogenization, as obtained from dynamic light scattering (DLS) measurements.

194 195

Figure 2 presents the size (hydrodynamic radius) distribution of the ChNFs based on DLS

196

data, for neat ChNF and the ChNF-c which was prepared with the addition of chitosan. Two

197

peaks (a bimodal distribution) can be observed for neat ChNF, one with the particle size in

198

the range of 350–1300 nm and the other with particle size between 60–150 nm. For the

199

ChNF-c prepared with the addition of chitosan, there is only one peak (a unimodal

200

distribution) with lower particle size in the range of 400–700 nm and higher peak intensity

201

compared to the neat ChNF. This indicates that the size distribution of ChNF-c particles is

202

more uniform as compared to the neat ChNF particles. With the addition of chitosan during

203

mechanical homogenization, the colloidal chitin system appears more stable. Similar DLS

ACS Paragon Plus Environment

10

Page 11 of 24 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 Sustainable Chemistry & Engineering

204

results were observed with core-shell cellulose nanofibrils/xyloglucan (CNF/XG) system, in

205

which adsorbed XG provides steric stabilization of CNF nanofibrils.18 Quartz Crystal

206

Microbalance (QCM) analysis results from our previous work indicated that there is no

207

evidence of chitosan adsorption to chitin nanofiber at pH 3, due to strong electrostatic

208

repulsive forces between protonated amine groups from both chitin nanofibers and chitosan.15

209

Thus, chitosan is possibly strongly associated between ChNF nanofibirls, contributing to

210

electrostatic stabilization of the whole suspension. In our previous study, the colloidal

211

stability of ChNFs was significantly improved by adding 30 wt.% chitosan into the aqueous

212

suspension of ChNFs and the aggregation of ChNFs was completely reduced as revealed by

213

DLS analysis.15 However, such stabilization effect was not achieved for chitosan addition

214

lower than 30 wt.% to the aqueous suspension of ChNFs. Thus, it is essential to add chitosan

215

during the homogenization process of lobster chitin to prepare colloidal stable ChNFs.

216 217

Structure and mechanical properties of chitin films

218

Nanostructured chitin films were prepared from water suspension of chitin nanofibers by an

219

Ultra Turrax mixing step followed by vacuum filtration and drying with a Rapid Köthen sheet

220

former.17,19 In Figure 3a, aggregates and bundles of chitin nanofibers with a width of 20–100

221

nm are observed in the surface SEM micrograph of a of a neat ChNF film, similar to the

222

result by TEM analysis (Figure 1a). Chitin films prepared from ChNF-c shows much more

223

homogeneous nanofibril network structure with much less aggregates of ChNFs, as shown in

224

Figure 3b. Free chitosan was vacuum filtered during the film preparation process. 4 ± 1 wt.%

225

chitosan remained in the chitin film after filtration. It is evident from both SEM (Figure 3b)

226

and TEM (Figure 1b) analysis that the nanofiber length is several micrometers and nanofiber

227

ends are not apparent. The predominant chitin orientation in both chitin films appears to be

ACS Paragon Plus Environment

11

ACS Sustainable Chemistry & Engineering 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 24

228

random-in-the-plane. The layered structure in the cross section of fractured surfaces of both

229

chitin films is apparent, similar to cellulose nanopaper structures.17 The layered structure of

230

neat ChNF (Figure 3c) is rather distinct as compared to ChNF-c sample. This is possibly due

231

to facilitated floc formation of larger ChNF fibril aggregates in the high concentration region

232

just above the filter membrane. Individualized ChNFs associated with chitosan show

233

interfibril repulsion, are better dispersed and able to form a nanofibril network linking the

234

adjacent layers (Figure 3d). Such structure is further supported by WAXD analysis discussed

235

latter.

236

237 238

Figure 3. Scanning electron microscopy (SEM) micrographs of the chitin nanofibrillar

239

network on the surfaces of (a) neat ChNF and (b) ChNF-c films, and layered structure on the

240

cross section of fracture surfaces of (c) neat ChNF and (d) ChNF-c films.

ACS Paragon Plus Environment

12

Page 13 of 24 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 Sustainable Chemistry & Engineering

241 242

Figure 4. Typical tensile stress–strain curves of ChNF and ChNF-c chitin films under

243

uniaxial tensile deformation.

244

Table 1. Properties of the chitin films. (The values in parentheses are the sample standard

245

deviations.) Elastic

Yield

Tensile

Strain-

Work to

modulus

strength

strength

to-failure

fracture

(GPa)

(MPa)

(MPa)

(%)

(MJ/m3) 7.3 (1.2)

Porosity Density (%)

(g/cm3)

ChNF

16

1.19

6.5 (0.5)

57.4 (2.2)

153.0 (10.6)

7.6 (0.6)

ChNF-c

18

1.17

6.0 (0.6)

54.1 (1.2)

187.2 (5.6)

10.1(0.2) 12.1 (0.4)

246 247

Figure 4 presents the stress–strain curves of chitin film samples of neat ChNF and the ChNF

248

with added chitosan during mechanical homogenization (ChNF-c). Mechanical properties of

249

the chitin films are summarized in Table 1 together with porosity and density data. The

250

ChNF-c film has slightly higher porosity. The ChNF sample shows higher variations in

251

mechanical properties due to the presence of aggregates in the film. The stress–strain

252

behavior of both samples follows the same curve with an elastic deformation region at lower

253

strain (below 2%) followed by a linear and strong strain-hardening (plastic deformation)

ACS Paragon Plus Environment

13

ACS Sustainable Chemistry & Engineering 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 24

254

region. The modulus and yield strength of the nanostructured chitin ChNF-c film decreased

255

from 6.5 ± 0.5 to 6.0 ± 0.6 GPa and from 57.4 ± 2.2 to 54.1 ± 1.2 MPa, respectively, due to

256

the presence of a soft matrix from 4 ± 1 wt.% residual chitosan. However, the ChNF-c film

257

has higher strain-to-failure and ultimate tensile strength of 10.1 ± 0.2% and 187.2 ± 5.6 MPa,

258

respectively. This tensile strength value is higher than the neat ChNF sample (153.0 ± 10.6

259

MPa) and higher than for chitin films from chitin-protein composite fibers (77  5.6 MPa) in

260

our previous work.9 This value is also higher than the chitin films from surface deacetylated

261

chitin nanofibers from crab shell reported by Ifuku et al.14 (156.5 MPa) and Fan et al.13 (140

262

MPa), as well as the film from TEMPO-oxidized α-chitin nanowhiskers (110 MPa).13 The

263

work to fracture of the ChNF-c film is 12.1 ± 0.4 MJ/m3, 66% higher than that for the neat

264

ChNF film. Although the nanocomposite film of ChNF with 30 wt.% content of chitosan can

265

achieve a work to fracture of 12 MJ/m3, its tensile strength can only reach 141 MPa.15

266

In the strain-hardening region, the highly individualized nanoscale chitin nanofibers can slide

267

and reorient much easier in the presence of 4 ± 1 wt.% residual chitosan compared to the

268

larger aggregates and bundles of nanofibers in the neat chitin film, particularly for out-of-

269

plane nanofibers between the layers parallel to the surface plan. This and fewer aggregate

270

defects result in much higher strain-to-failure and ultimate tensile strength for the ChNF-c

271

film. The orientation of the chitin nanofibers in the film before and after the tensile test was

272

characterized by WAXD. The X-ray diffractograms perpendicular and parallel to the film

273

surface are shown in Figure 5. Before tensile stretching, the orientation of chitin in the plane

274

of the films is completely random. The plan parallel to the film surface is strongly ordered.

275

The degree of in-plane orientation () for the ChNF-c film is 74.2%, lower than that for neat

276

ChNF film (76.4%), indicating more out-of-plane nanofibers in the layered structure.

277

Interestingly this correlates with the lower modulus of ChNF-c compared with ChNF films,

278

see Table 1. These results confirm the observations by SEM (Figure 3). After tensile

ACS Paragon Plus Environment

14

Page 15 of 24 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 Sustainable Chemistry & Engineering

279

stretching, the orientation in the surface plane of the films is still completely random,

280

although there may be local orientation effects close to failure sites. However, the  value for

281

the ChNF-c film parallel to the film surface is increased by 4.8% after tensile testing, and this

282

is substantially higher than the 2.5% increase for the neat ChNF film. It means that the out-

283

of-plane nanofibers become more strongly oriented in the ChNF-c film during the tensile test.

284

285 286

Figure 5. X-ray diffractograms perpendicular (through) and parallel (edge) to the membrane

287

surfaces for the ChNF and ChNF-c samples before and after uniaxial tensile stretching. The

288

degree of orientation () was calculated according to equation (2) from azimuthal intensity

289

distribution graphs for the 110 equatorial reflection (Figure S1, Supporting Information).

290 291

Conclusions

292

The main problem of -chitin nanofibers extracted from crustaceans such as lobster is the

293

poor colloidal stability due to aggregate formation in the suspension. These aggregates form

ACS Paragon Plus Environment

15

ACS Sustainable Chemistry & Engineering 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 24

294

defects in the nanofiber network and lead to strong variations in mechanical properties. We

295

have demonstrated that addition of chitosan during homogenization of the chitin slurry

296

facilitates complete individualization of chitin nanofibers. Chitosan associates with chitin

297

nanofibers and increases the electrostatic repulsion forces and stability of the colloidal

298

suspension. Chitosan is used as a sacrificial polymer and is partially removed by vacuum

299

filtration during film forming process. The chitosan addition leads to increased tensile

300

strength from 156 MPa to 187 MPa and increased work to fracture (area under stress-strain

301

curve) from 7.3 to 12.1 MJ/m3 for the nanostructured chitin ChNF-c film. The improved

302

mechanical properties are attributed to the true nanoscale network structure from better

303

individualization of chitin nanofibers stabilized by the residual chitosan. The new method of

304

using sacrificial chitosan for preparation of colloidal stable chitin nanofibers from -chitin

305

resources is simple and more environmentally friendly than chemical methods for colloidal

306

stabilization such as surface deacetylation and TEMPO-mediated oxidation.

307 308

ASSOCIATED CONTENT

309

Supporting Information.

310

The following file is available free of charge.

311

Azimuthal intensity distribution graphs for the 110 reflection for the stretched and

312

unstretched ChNF and ChNF-c films (PDF)

313

AUTHOR INFORMATION

314

Corresponding Author

315

* To whom correspondence should be addressed, Tel: +46 8 790 96 25. E-mail: [email protected]

316

ACKNOWLEDGMENT

ACS Paragon Plus Environment

16

Page 17 of 24 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 Sustainable Chemistry & Engineering

317

The Wallenberg Wood Science Center (WWSC) is acknowledged for financial support for

318

this work. LB acknowledges funding from the Knut and Alice Wallenberg foundation

319

through a Wallenberg Scholar grant.

320 321

REFERENCES

322

(1)

323

Preparation of chitin nanofibers with a uniform width as -chitin from crab shells.

324

Biomacromolecules 2009, 10 (6), 1584–1588.

325

(2)

326

Simple preparation method of chitin nanofibers with a uniform width of 10–20 nm from

327

prawn shell under neutral conditions. Carbohyd. Polym. 2011, 84 (2), 762–764.

328

(3)

329

from mushrooms. Materials 2011, 4 (8), 1417–1425.

330

(4)

331

Nanoscale 2012, 4 (11), 3308–3318.

332

(5)

333

chitin by simple mechanical treatment under acid conditions. Biomacromolecules 2008, 9 (7),

334

1919–1923.

335

(6)

336

single chitin nanofibrils via sonication-induced fragmentation. Biomacromolecules 2017, 18

337

(12), 4405–4410.

338

(7)

339

Biomacromolecules 2012, 13 (1), 1–11.

Ifuku, S.; Nogi, M.; Abe, K.; Yoshioka, M.; Morimoto, M.; Saimoto, H.; Yano, H.

Ifuku, S.; Nogi, M.; Abe, K.; Yoshioka, M.; Morimoto, M.; Saimoto, H.; Yano, H.

Ifuku, S.; Nomura, R.; Morimoto, M.; Saimoto, H. Preparation of chitin nanofibers

Ifuku, S.; Saimoto, H. Chitin nanofibers: preparations, modifications, and applications.

Fan, Y. M.; Saito, T.; Isogai, A. Preparation of chitin nanofibers from squid pen -

Bamba, Y.; Ogawa, Y.; Saito, T.; Berglund, L. A.; Isogai, A. Estimating the strength of

Zeng, J. B.; He, Y. S.; Li, S. L.; Wang, Y. Z. Chitin whiskers: an overview.

ACS Paragon Plus Environment

17

ACS Sustainable Chemistry & Engineering 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 24

340

(8)

Hackman, R. Studies on chitin IV. The occurrence of complexes in which chitin and

341

protein are covalently linked. Aust. J. Biol. Sci. 1960, 13 (4), 568–577.

342

(9)

343

chitin-protein composite nanofibers-structure and mechanical properties. J. Appl. Polym. Sci.

344

2014, 131, 40121.

345

(10) Mushi, N. E.; Butchosa, N.; Salajkova, M.; Zhou, Q.; Berglund, L. A. Nanostructured

346

membranes based on native chitin nanofibers prepared by mild process. Carbohyd. Polym.

347

2014, 112, 255–263.

348

(11) Mushi, N. E.; Kochumalayil, J.; Cervin, N. T.; Zhou, Q.; Berglund, L. A.

349

Nanostructurally controlled hydrogel based on small-diameter native chitin nanofibers:

350

preparation, structure, and properties. Chemsuschem 2016, 9 (9), 989–995.

351

(12) Nishino, T.; Matsui, R.; Nakamae, K. Elastic modulus of the crystalline regions of

352

chitin and chitosan. J. Polym. Sci. B: Polym. Phys. 1999, 37 (11), 1191–1196.

353

(13) Fan, Y. M.; Fukuzumi, H.; Saito, T.; Isogai, A. Comparative characterization of

354

aqueous dispersions and cast films of different chitin nanowhiskers/nanofibers. Int. J. Biol.

355

Macromol. 2012, 50 (1), 69–76.

356

(14) Ifuku, S.; Ikuta, A.; Egusa, M.; Kaminaka, H.; Izawa, H.; Morimoto, M.; Saimoto, H.

357

Preparation of high-strength transparent chitosan film reinforced with surface-deacetylated

358

chitin nanofibers. Carbohyd. Polym. 2013, 98 (1), 1198–1202.

359

(15) Mushi, N. E.; Utsel, S.; Berglund, L. A. Nanostructured biocomposite films of high

360

toughness based on native chitin nanofibers and chitosan. Front. Chem. 2014, 2, 99.

361

(16) Carlström, D. The crystal structure of -chitin. J. Biophys. Biochem. Cytol. 1957, 3 (5),

362

669–683.

363

(17) Henriksson, M.; Berglund, L. A.; Isaksson, P.; Lindström, T.; Nishino, T. Cellulose

364

nanopaper structures of high toughness. Biomacromolecules 2008, 9 (6), 1579–1585.

Mushi, N. E.; Butchosa, N.; Zhou, Q.; Berglund, L. A. Nanopaper membranes from

ACS Paragon Plus Environment

18

Page 19 of 24 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 Sustainable Chemistry & Engineering

365

(18) Prakobna, K.; Terenzi, C.; Zhou, Q.; Furo, I.; Berglund, L. A. Core–shell cellulose

366

nanofibers for biocomposites – nanostructural effects in hydrated state. Carbohyd. Polym.

367

2015, 125, 92–102.

368

(19) Sehaqui, H.; Liu, A. D.; Zhou, Q.; Berglund, L. A. Fast preparation procedure for large,

369

flat cellulose and cellulose/inorganic nanopaper structures. Biomacromolecules 2010, 11 (9),

370

2195–2198.

371 372 373 374

BRIEFS. Strong and tough chitin film from highly individualized -chitin nanofibers that are

375

prepared by using chitosan as a sacrificial polymer during homogenization.

376 377

SYNOPSIS.

378 379

ACS Paragon Plus Environment

19

ACS Sustainable Chemistry & Engineering 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

Figure 1. Typical transmission electron microscopy (TEM) images of chitin nanofibers from lobster shell (a) disintegrated without the addition of chitosan and (b) disintegrated in the presence of chitosan in suspension. 70x142mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 20 of 24

Page 21 of 24 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 Sustainable Chemistry & Engineering

Figure 2. Estimated size distributions of chitin nanofibers in water suspensions for neat chitin (ChNF) and chitin with the addition of 10 wt.% chitosan (ChNF-c) in suspension during mechanical homogenization, as obtained from dynamic light scattering (DLS) measurements. 81x60mm (300 x 300 DPI)

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

Figure 3. Scanning electron microscopy (SEM) micrographs of the chitin nanofibrillar network on the surfaces of (a) neat ChNF and (b) ChNF-c films, and layered structure on the cross section of fracture surfaces of (c) neat ChNF and (d) ChNF-c films. 142x106mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 22 of 24

Page 23 of 24 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 Sustainable Chemistry & Engineering

Figure 4. Typical tensile stress–strain curves of ChNF and ChNF-c chitin films under uniaxial tensile deformation. 79x61mm (300 x 300 DPI)

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

Figure 5. X-ray diffractograms perpendicular (through) and parallel (edge) to the membrane surfaces for the ChNF and ChNF-c samples before and after uniaxial tensile stretching. The degree of orientation (Π) was calculated according to equation (2) from azimuthal intensity distribution graphs for the 110 equatorial reflection (Figure S1, Supporting Information). 165x82mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 24 of 24