gelatin hydrogels with enhanced mechanical strength

gelatin hydrogels with enhanced mechanical strength

Accepted Manuscript Title: Rapid Shape Memory TEMPO-Oxidized Cellulose Nanofibers/Polyacrylamide/Gelatin Hydrogels with Enhanced Mechanical Strength A...

2MB Sizes 108 Downloads 73 Views

Accepted Manuscript Title: Rapid Shape Memory TEMPO-Oxidized Cellulose Nanofibers/Polyacrylamide/Gelatin Hydrogels with Enhanced Mechanical Strength Authors: Nan Li, Wei Chen, Guangxue Chen, Junfei Tian PII: DOI: Reference:

S0144-8617(17)30424-1 http://dx.doi.org/doi:10.1016/j.carbpol.2017.04.035 CARP 12223

To appear in: Received date: Revised date: Accepted date:

24-2-2017 31-3-2017 15-4-2017

Please cite this article as: Li, Nan., Chen, Wei., Chen, Guangxue., & Tian, Junfei., Rapid Shape Memory TEMPO-Oxidized Cellulose Nanofibers/Polyacrylamide/Gelatin Hydrogels with Enhanced Mechanical Strength.Carbohydrate Polymers http://dx.doi.org/10.1016/j.carbpol.2017.04.035 This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

1

Rapid Shape Memory TEMPO-Oxidized Cellulose

2

Nanofibers/ Polyacrylamide/ Gelatin Hydrogels with

3

Enhanced Mechanical Strength

4

Nan Li a, Wei Chen b, *, Guangxue Chen a, *, Junfei Tiana

5

a

6

Guangzhou 510640, P. R. China; b College of Engineering, Qufu Normal University, RiZhao 276826, China

7

* Corresponding authors at South China University of Technology, Guangzhou 510640, P. R. China and

8

College of Engineering, Qufu Normal University, RiZhao 276826, China.

9

E-mail addresses: [email protected] (N.Li); [email protected] (W.Chen); [email protected]

State Key Laboratory of Pulp and Paper Engineering, South China University of Technology,

10

(G.X.Chen); [email protected] (J.F.Tian)

11

TEL: (86)-020-22236485 ; Fax :( 86)-020-22236485

12

Abstract

13

TEMPO-oxidized cellulose nanofibers/ polyacrylamide/ gelatin shape memory hydrogels

14

were successfully fabricated through a facile in-situ free-radical polymerization method, and

15

double network was formed by chemically cross-linked polyacrylamide (PAM) network and

16

physically cross-linked gelatin network. TEMPO-oxidized cellulose nanofibers (TOCNs)

17

were introduced to improve the mechanical properties of the hydrogel. The structure, shape

18

memory behaviors and mechanical properties of the resulting composite gels with varied gel

19

compositions were investigated. The results obtained from those different studies revealed

20

that TOCNs, gelatin, and PAM could mix with each other homogeneously. Due to the

21

thermoreversible nature of the gelatin network, the composite hydrogels exhibited attractive

22

thermo-induced shape memory properties. In addition, good mechanical properties 1

23

(strength >200 kPa, strain >650%) were achieved. Such composite hydrogels with good

24

shape memory behavior and enhanced mechanical strength would be an attractive candidate

25

for a wide variety of applications.

26

Keywords: TEMPO-oxidized cellulose nanofibers (TOCNs), gelatin, hydrogels, mechanical

27

properties, shape memory.

28

1. Introduction

29

Shape memory hydrogels are kinds of special stimuli-responsive, soft, and wet material

30

that have the ability to fix a temporary deformation and memorize the original shape when

31

exposed to a specific external stimulus such as heat (Hao & Weiss, 2013), light (Xiao, Gong,

32

Kang, Jiang, Zhang & Li, 2016; Zhang, Wang, Hong, Sun, Liu & Tong, 2014), ultrasound (Li,

33

Yan, Xia & Zhao, 2015), solvent (Xiao et al., 2016), electric field (Zhao, Tan, Cui, Deng,

34

Huang & Guo, 2014). They have many potential applications in the fields of biomedical

35

(Kabir et al., 2014), tissue engineering (Huang et al., 2012), and smart devices (Harmon,

36

Tang & Frank, 2003). Shape memory hydrogels are most commonly triggered by heat.

37

Typically, shape memory hydrogels consist the permanent netpoints (hard domains) and

38

crystalline or glassy domains acting as temporary cross-links (switching domains).While the

39

netpoints determine the permanent shape of the hydrogel, the physical cross-links formed by

40

solidification of the switching domains below the glass or transition temperatures fix the

41

temporary shape gained by deformation of the gel sample at higher temperatures (Behl,

42

Razzaq & Lendlein, 2010). Nevertheless, the low mechanical properties of the shape memory

43

hydrogels limit their applications. In recent years, the filling of shape memory polymer

44

composites with different nanofillers broadens the varieties of shape memory hydrogels and 2

45

significantly expanding their applications. Nanofillers primarily include organic agents, such

46

as azobenzene derivatives, inorganic particles, such as nano clay, and polymers, such as poly

47

(ethylene glycol) (Meng & Li, 2013). Thanks to its biodegradability, renewability and

48

non-toxicity, nanocellulose is thereby considered as a promising bio-based material derived

49

from native cellulose sources with low impact on the environment.

50

At present, numerous studies have been carried out to reinforce hydrogels by

51

incorporating rigid, rod-like cellulose nanocrystals (CNCs), one class of nanocellulose that is

52

often isolated by strong acid hydrolysis (Habibi, Lucia & Rojas, 2010; Wang, Zhu, Reiner,

53

Verrill, Baxa & Mcneil, 2012). These studies indicate that CNCs exhibit good reinforcing

54

effect (Hebeish, Farag, Sharaf & Shaheen, 2014). As another type of nanocellulose materials,

55

cellulose nanofibers (CNFs) are marked by a three-dimensional network structure formed by

56

the entanglement of flexible nanofibers typically having a higher aspect ratio and flexibility

57

than CNCs (Xu, Liu, Jiang, Zhu, Haagenson & Wiesenborn, 2013). To this effect, CNFs are a

58

favorable reinforcing candidate for the hydrogel. CNFs can be produced by first pretreated

59

with TEMPO-mediated oxidation (2, 2, 6, 6, -tetrame-thylpipelidine-1-oxyl radical), and then

60

multipass high-pressure homogenization. As a kind of important “green” filler, TOCNs have

61

great potential in reinforcing various kinds of polymers (Wei, Chen, Liu, Du, Yu & Zhou,

62

2016). Gelatin, a partial derivative of collagen, possesses distinctive characteristics, such as

63

low cost, biocompatibility, and biodegradability. Besides, it could easily dissolve in water and

64

subsequently form a thermal reversible physical gel at low temperature, in which the

65

macromolecular chains recover the triple-helix structure of collagen (Bohidar & Jena, 1993;

66

Gornall & Terentjev, 2008). What’s more, constructing double network structures (DN gels) 3

67

is an ideal strategy to obtain high mechanical properties of hydrogels (Gong, 2010).

68

In this study, a type of protein/polymer-TOCNs shape memory hydrogel was successfully

69

synthesized through the in situ polymerization of acrylamide (Am) in a mixed suspension of

70

gelatin and TOCNs. This composite gels including two regions: the chemically crosslinked

71

polyacrylamide (PAM) function as the “permanent” cross-links; and physically crosslinked

72

gelatin as the “temporary” cross-links. TOCNs were introduced to improve the mechanical

73

properties of the hydrogel. The compatibility, morphology, shape memory properties and

74

mechanical properties of the resulting shape memory hydrogels with varying gel

75

compositions were investigated.

76

2. Experimental details

77

2.1 Materials

78

Bleached hardwood pulp was kindly supplied from the Guangzhou Chenhui Paper Co.,

79

Ltd. (China). (2, 2, 6, 6-Tetramethylpiperidin-1-yl) oxyl (TEMPO, 98 wt %), sodium bromide

80

(NaBr), sodium hypochlorite (NaClO), gelatin, acrylamide (AM), potassium peroxidisulfate

81

(KPS, K2S2O8, 99%, J&K), N, N-methylene bisacrylamide (MBA) and N, N, N′,

82

N′-tetramethyldiamine (TEMED) were purchased from Sinopharm Chemical Reagent Co.

83

Ltd (SCRC) and used without further purification. The concentrations of the solutions of KPS

84

and MBA were fixed at 20mg/mL. They were freshly prepared just before use.

85

2.2 Preparation of TOCNs

86

TEMPO-mediated oxidation was conducted according to a procedure described in the

87

literature (Saito, Kimura, Nishiyama & Isogai, 2007). Specifically, 20g dried hardwood pulp

88

was first dispersed in water with a mechanical stirrer until the solid content of 1 wt% was 4

89

achieved. Then 0.32g TEMPO and 2.32 g NaBr were dissolved into the slurry and the

90

reaction system was adjusted to pH 10 by the dropwise addition of 10% NaOH solution.

91

Oxidation began immediately after adding sodium hypochlorite (284g, 6.3% chlorine solution)

92

at 30 °C. A pH meter was used to monitor the pH and 10% NaOH solution was used to

93

maintain the pH at 10 during the reaction. The reaction time for the chemical oxidation was

94

approximately 3 h. After the TEMPO-mediated oxidation, the pulp suspensions were filtered

95

and thoroughly washed with distilled water until pH values of the filtrate reached neutral (i.e.,

96

unchanged). The TEMPO-oxidized fiber suspension mechanical treated by successively

97

passing it through a high-pressure homogenizer. The obtained transparent TOCNs was stored

98

at 4 °C without any treatment before future utilization.

99

2.3 Synthesis of TOCN/PAM/Gelatin Composite Hydrogels

100

The hydrogel was prepared through the in situ radical polymerizations of the monomer

101

(Am) in an aqueous suspension of TOCNs and gelatin with a small amount of MBA as a

102

co-cross-linker. Firstly, the desired amount of gelatin (0−10 mg/mL) was added into a 50mL

103

dried three-necked flask and dissolved in water at 45°C for 1 h. Successively, AM (6.4g),

104

MBA (76 μL, 20mg/mL), and TOCN (0−5 wt% to AM) were added. After being stirred at

105

45 °C for 2 h, the mixture became uniform. The oxygen was removed from the solution with

106

nitrogen gas for 10 min. Thereafter, KPS solution (1.8mL, 20mg/mL) and catalyst TEMED

107

(30 μL) were added to the solution and stirred for 1min. The mixture solution was rapidly

108

injected into a laboratory-made mold which consisted of two glass sheets with a 2mm rubber

109

spacer. Finally, the mold was cooled down to 6°C to form a gelatin network, followed by

110

polymerization at room temperature for 12 h to obtain the nanocomposite hydrogels. In this 5

111

article, the hydrogels are denoted as TOCNx/PAM/Gelatiny, which represents x wt% TOCN to

112

AM, and y w/v% gelatin. The total volume of the solution mixture was fixed at 30 mL with

113

deionized water. Detailed feed compositions are listed in Table S1.

114

2.4 Characterization.

115

Atomic force microscopy (AFM) was employed to investigate the morphology of TOCN

116

by using a Nanoscope III instrument (Veeco Co., Ltd., USA). X-ray diffraction (XRD)

117

patterns were recorded using an X-ray diffractometer at a range of 2 = 5−60 (D8

118

ADVANCE, Bruker, Germany). Attenuated total reflectance-fourier transform infrared

119

(ATR-FTIR) spectra of hydrogels were obtained on a VERTEX 70 FT-IR spectrometer

120

(Bruker, Germany) in the range of 600−3600cm-1. For SEM observations, the samples were

121

freeze-dried in a lyophilizer for 24 h at −50 °C and then coated with gold (S-3700, Hitachi,

122

Japan).

123

2.5 Shape Memory Behavior

124

A hydrogel strip (with 30mm length × 4mm width × 2mm thickness) was cut from the

125

as-prepared hydrogels. The gel strip was deformed into a helix after being immersed in 90 °C

126

water for 20 s. Then, it was immersed in ice water for 30 s to fix the shape. Quantitative

127

information on the shape memory properties was determined according to the reported

128

method (Chen, Peng, Liu, Wang, Shi & Wang, 2016; Huang, Zhao, Wang, Sun & Tong,

129

2016). The shape fixity ratios (Rf) were defined by the following equation (1):

130

=

× 100%

(1)

131

Where θi was the actually curled angle and θt was the maintained angle after fixing in ice

132

water and keeping it in air for 5 min. 6

133

The shape memory behavior was further observed by stretching a hydrogel strip. A gel

134

strip (with 36mm length × 10mm width × 2mm thickness) was first immersed in 90 ° C water

135

for 20 s and then stretched to an elongated shape. The fixation of the elongated gel strip was

136

conducted by keeping in ice water for 30s.

137

2.6 Equilibrium Swelling Degree of Hydrogels

138

Swelling experiments were performed at room temperature by immersing the dried

139

hydrogels in excess distilled water to obtain swelling equilibrium. The swelling ratio (SR) of

140

the hydrogel was calculated according to the following equation (2):

141

Swelling ratio =

142

Where

143

equilibrium state, respectively.

144

2.7 Dynamic Mechanical Analysis

and

(2) represent the weights of the dried hydrogel and the hydrogel at swelling

145

Dynamic mechanical analysis was carried out using a stress-controlled rheometer

146

RheoStress 600 (Thermo Fisher Instrument, USA) with a parallel plate fixture of the diameter

147

of 40 mm. Temperature sweeps were performed from 25 to 75 °C and then cooled down to

148

25 °C without waiting at a rate of 2 °C/min with a fixed strain amplitude γ = 0.5% and

149

angular frequency ω = 6.28 rad/s. Storage modulus (G′) and loss modulus (G″) were

150

calculated using the dynamic mechanical analysis software. Silicone oil was laid on the edge

151

of the plates to prevent water evaporation.

152

2.8 Measurements of Mechanical Properties

153

Tensile tests were performed on the as-prepared gel samples (with 30mm length × 4mm

154

width × 2mm thickness) using an Instron Universal Testing Machine 5565 with a crosshead 7

155

speed of 100mm/min at room temperature.

156

3. Results and discussion

157

3.1 Characterization of TOCNs.

158

AFM was used to investigate the morphology and dimension of TOCNs (Figure 1a). As

159

shown in Figure 1a, most TOCNs have diameters mainly distributed in the range of 15~25nm

160

and show entangled morphology. Upon oxidization of the surface hydroxyl groups, some

161

changes on the spectrum of TOCNs can be observed (Figure 1b). The most important change

162

is the appearance of the sodium carboxylate groups (−COONa) stretching band at 1605 cm−1.

163

Based on this analysis, the success of the chemical oxidization of cellulose was clearly

164

verified by FTIR.

(a)

(b) 200 nm

600 nm 165 166

Figure.1. (a) AFM image of TOCNs, and (b) FTIR spectrum of TOCNs

167

3.2 Preparation and Structural Characterization of the Formation of TOCN/PAM/Gelatin

168

Nanocomposite Gels

169

Scheme 1a shows the proposed mechanism for the reversible triple helices mechanism of

170

gelatin. It can be seen that during cooling, the gelatin chains underwent the coil-helix

171

transition followed by the aggregation of triple helices. As a result, the first, physically 8

172

cross-linked gelatin network was formed. Upon KPS initiation, the chemically cross-linked

173

polyacrylamide was formed as the second network through the radical polymerization

174

(Scheme 1b). Thus, the hydrogels with the double networks of polyacrylamide and gelatin

175

were obtained.

176 177

Scheme.1. (a) Heat-induced reversible triple helices mechanism of gelatin, (b) Synthesis scheme of

178

TOCN/PAAM/Gelatin gels.

179

ATR-FTIR spectra of the hydrogel samples are presented in Figure 2a. The PAM curve

180

had a broadband around 3339 cm−1 and a characteristic peak at 3189 cm−1 due to the

181

stretching vibration of N–H. The significant characteristic peaks at 1655 and 1612 cm−1 were

182

attributed to the carbonyl stretching vibration (amide I) and N–H bending vibration (amide II)

183

of the amide group, respectively (Zhou & Wu, 2011). In the ATR-FTIR spectra of the

184

composite

185

TOCN5/PAM/Gelatin10, the absorptions associated with characteristic groups of PAM were

186

clearly identified. Compared with pure PAM, there was no perceptible shift in positions

hydrogels

of

TOCN0/PAM/Gelatin10,

9

TOCN5/PAM/Gelatin0

and

187

occurred in the composite hydrogels. This may be due to the lower TOCN contents and its

188

characteristic absorption peaks were overlapped by those associated with PAM. Additionally,

189

there was no appearance of new vibration, indicating that no new covalent bond formed

190

between gelatin, PAM, and TOCN.

(a)

(b)

TOCN5/PAM/Gelatin10

Intensity (a.u.)

T (%)

TOCN5/PAM PAM/Gelatin10 PAM

3339

3600

191

3189

TOCN PAM TOCN5/PAM TOCN5/PAM/Gelatin10

1655 1612

3000 2400 1800 1200 Wavenumbers (cm-1)

600

5 10 15 20 25 30 35 40 45 50 55 60 2

192

Figure.2. Physicochemical characterization as elucidated by (a) ATR-FTIR spectra; and (b) XRD

193

patterns of the hydrogels.

194

Figure 2b shows the XRD patterns of TOCNs, PAM, TOCN5/PAM, and

195

TOCN5/PAM/Gelatin10 hydrogels. The diffraction pattern of PAM does not show an obvious

196

sharp peak of the crystalline structure; instead, a blunt peak centered at 2θ= 26.1° is observed.

197

It is noteworthy that in comparison with TOCN there is no characteristic TOCN peak at 22.6°

198

detected in the XRD patterns of TOCN5/PAM and TOCN5/PAM/Gelatin10 hydrogels, which

199

suggested that TOCNs were uniformly dispersed in the polymer gels without ordered

200

aggregation.

201

Lyophilization is known to preserve the structure and volume of the hydrogels even after

202

all of the solvent has been removed. Figure 3 shows the SEM images of the lyophilized PAM

203

and TOCN5/PAM/Gelatin10 hydrogels. It can be seen that the network of the net PAM

204

hydrogel was clear with a uniform macropore structure and the pore sizes ranging from ~30 10

205

to ~200 μm. For the network of TOCN5/PAM/Gelatin10 hydrogel (Figure 3c and d), it was

206

observed that the pore structure of the gels became more complex and with a smaller pore

207

size, thus implying a denser structure of the hydrogel. This can be attributed to the formation

208

of polyacrylamide-gelatin double network and the strong hydrogen-bonding interactions

209

between TOCNs and polar functional groups of the PAM side chains.

(b)

(a)

200 µ m

50 µ m

(c)

(d)

200 µ m

50 µ m

210 211

Figure.3. SEM micrographs of (a, b) PAM hydrogel and (c, d) TOCN5/PAM/Gelatin10 hydrogel.

212

3.3 Shape Memory Behavior

213

The hydrogels exhibited good shape memory behaviors. Figure 4b shows the thermally

214

triggered shape memory behavior of a composite hydrogel strip. The gel strip (Figure 4a) was

215

curled into a tight spiral after being immersed in 90 °C water for 20s. Afterwards, the

216

deformed gel was fixed in ice water for 30s to obtain a temporary deformed shape (Figure 4b).

217

When the deformed gel helix was immersed into 90 °C water, it can recover its original shape

218

rapidly (Movie S1). These results indicate that the hydrogels containing gelatin exhibit the 11

219

cooling-fixed shape memory capability.

(a)

(b)

(c)

220 221

Figure.4. Photos of (a) original and recovered shape of the gel strip;(b) fixed temporary shape of the

222

hydrogel. (c) Shape fixity ratios (Rf) of the gels prepared with different contents of TOCNs and gelatin

223

after fixing in ice water and keeping in the room temperature for 5 min. The numbers in the shape fixity

224

ratios picture indicate the content of the corresponding components in Table S1.

225

The shape fixity ratios (Rf) of the gels were obtained with the simple bending

226

experiments performed. The Rf of the gels was not significantly affected by TOCN contents,

227

while the contents of gelatin had an obvious effect on them (Figure 4c). The Rf of the

228

TOCN5/PAM/Gelatin10 gels were mostly around 65%, higher than those of the corresponding

229

gels that have lower contents of gelatin. It is noteworthy that the TOCN5/PAM/Gelatin0

230

sample without gelatin cannot be reshaped. All of the fixed shapes could rapid recover their

231

original shapes when immersing in 90 °C water.

12

Original Shape (a)

Deformed Shape (b)

Recovered Shape (c)

232 233

Figure.5. Photos showing the heat-shrinkable shape memory behavior of TOCN0/PAM/Gelatin10

234

hydrogel. (a−c): A gel strip of 36mm×10mm×2mm in the original (a), elongated (b), and recovered

235

shapes (c). The fixation of the elongated gel strip was conducted by keeping in ice water for 30s. This

236

shape memory behavior was performed with as-prepared wet TOCN0/PAM/Gelatin10 hydrogel.

237

The composite hydrogel also exhibited a heat-shrinkable shape memory behavior

238

(Figure 5). This behavior was observed by an elongated TOCN0/PAM/Gelatin10 hydrogel strip.

239

A gel strip was first heated to 90 ° C and then stretched to an elongated shape. The fixation of

240

the elongated gel strip was conducted by keeping in ice water for 30s. After being immersed

241

in 90 °C water for 5s, the elongated part shrank quickly to a remaining length of 39 mm

242

(Figure 5c), very close to the original length of 36mm (Figure 5a).

243

To examine the repeatability of the shape memory capability of the hydrogels, the cycle

244

of shape fixing and thermo-induced shape recovery was repeated 10 times (Figure S1). The

245

hydrogel can remain almost the same shape fixity ratio (Rf) and recover to its original shape

246

after 10 cycles showing a good reproducible capability of shape memory.

247

Temperature-dependent dynamic mechanical properties of the composite hydrogels were

248

measured. As shown in Figure 6, the storage modulus (G′) was always higher than the

249

corresponding loss modulus (G″) in the entire temperature cycle of 25 °C→75 °C→25 °C,

250

suggesting the elastic nature of the TOCN/PAM/Gelation hydrogels. It can be seen that the 13

251

TOCN5/PAM/Gelatin0 hydrogel keeps almost constant in G′ and G″ (Figure 6a) during a

252

cycle of heating and cooling process because of the absence of gelatin. However, G′ is not a

253

constant for the TOCN5/PAM/Gelatin10 hydrogel (Figure 6b). As temperature increases, G′ of

254

the TOCN5/PAM/Gelatin10 hydrogel decrease rapidly at 35 °C. This was induced by the

255

helix-coil transition of the gelatin chains at 35 °C. When temperature is higher than 45 °C, G′

256

of the TOCN5/PAM/Gelatin10 hydrogel keep almost a constant, no matter heating or cooling.

257

This constant G′ is attributed to the cross-linking density of the chemically cross-linked PAM

258

network, which is not changed with temperature. Moreover, G′ increases greatly at about

259

35 °C and could almost recover its initial value during the process of cooling, implying the

260

thermoreversible sol-gel transition of the gelatin network. These findings clearly indicate that

261

the thermoreversible sol-gel transition of gelatin network regulates the moduli of the

262

hydrogels, endowing the shape-memory capability of these composite hydrogels.

263 264

Figure.6. Dependence of storage modulus (G′) and loss modulus (G″) on temperature during a

265

temperature sweep at a heating and cooling rate of 2 °C/min for (a) TOCN5/PAM/Gelatin0, and (b)

266

TOCN5/PAM/Gelatin10 hydrogels. An angular frequency ω=6.28 rad/s and a strain amplitude γ=0.5%

267

were adopted.

268

3.3 Swelling and Mechanical Properties of TOCN/PAAM/Gelatin Gels. 14

269

The swelling behaviors were investigated to further reveal the structural features of the

270

TOCNx/PAM/Gelatiny nanocomposite hydrogels. The results are shown in Figure S2. It was

271

well-known that the swelling behavior of the hydrogel depended on the crosslinked density of

272

the gel network. The swelling ratio of the TOCNx/PAM/Gelatiny hydrogels decreased with the

273

increasing TOCN contents (Figure S2a). This was attributed to the fact that TOCNs could

274

interact with the PAM network by means of hydrogen bonding to form a closer network

275

structure, thus causing a decrease in the swelling capacity. Meanwhile, the swelling ratios of

276

TOCN5/PAM/Gelatiny hydrogels decreased from 21.79 to 11.4 g/g with the percent of gelatin

277

increasing from 3 to 10 w/v% (Figure S2b). This behavior confirmed that the

278

TOCN5/PAM/Gelatin10 double network (DN) hydrogel had a more robust network than that

279

of the TOCN5/PAM/Gelatin3, TOCN5/PAM/Gelatin5 and TOCN5/PAM/Gelatin7 DN

280

hydrogels.

281

As shown in Figure 7, TOCN/PAM/Gelatin nanocomposite hydrogels exhibited

282

excellent tensile properties. The gels can be easily stretched (Figure 7a), twisted stretched

283

(Figure 7b) and stretching with a knot (Figure 7c) to at least 4 times their original length. The

284

physically linked networks could be easily extended and contributed to the stretchability of

285

DN gels. Moreover, TOCN/PAM/Gelatin gels can be readily adapted to different complex

286

shapes (Figure 7d-f), demonstrating their free-shapeable property.

15

(d) (a) 1cm

(e) (b) 1cm

(f) (c) 1cm

287 288

Figure.7. The composite hydrogels are very tough and flexible, with the ability to withstand different

289

high-level deformations of (a) stretching (b) twisted stretching, as well as (c) stretching with a knot. The

290

gels are also free shapeable into different complex shapes (d-f, dyed with methyl orange). All these

291

behaviors were performed with the as-prepared wet TOCN5/PAM/Gelatin10 hydrogels.

292

The mechanical properties of TOCNx/PAM/Gelatiny gels at room temperature are

293

showed in Figure 8, including tensile stress-strain curves (Figure. 8a, c) and elastic moduli (E)

294

(Figure. 8b, d).

295

The positive effect of the TOCN amount on the mechanical properties of hydrogel is

296

shown in Figure 8a. As TOCN contents increased, the maximum tensile stress (b), and

297

elastic modulus (E) of the DN gels consistently increased. TOCN0/PAM/Gelatin10 had b, and

298

E of 91 kPa, and 20.8 kPa respectively. Compared with TOCN0/PAM/Gelatin10,

299

TOCN5/PAM/Gelatin10 exhibited much stronger mechanical properties (b of 237 kPa, and E

300

of 45.5 kPa), indicating that the TOCNs are able to effectively bear the tensile load applied on 16

301

the PAM matrix and thus enhance the mechanical properties of hydrogels.

302

The introduction of gelatin increases the elastic modulus and tensile strength of the

303

composite hydrogel, further revealing the enhanced compatibility of the gelatin and the

304

TOCN/PAM

305

TOCN5/PAM/Gelatin10 was superior to TOCN5/PAM/Gelatin0 single network (SN) gel and

306

demonstrated its high toughness in tensile tests. It is worth noting that the increment of

307

gelatin contents could not only enhances the hydrogel strength but also increases the

308

shape-fixing percentage Rf in Figure 4.

system.

From

Figure

8

it

can

be

seen

that

the

hydrogel

of

309

The elongation (εb) at break, however, of the samples do display a different tendency

310

compared with the above properties. It is highly dependent on gelatin contents. The

311

elongation (εb) significantly and monotonically decreased from 1590% to 680% as gelatin

312

content increased from 3 w/v% to 10 w/v% (Figure 8c). The results indicate that the

313

elongation (εb) at break decreases with the increasing crosslinking density as expected. It can

314

be attributed to the increase of crosslinking density having a negative effect on the toughness

315

of polymers. Therefore, we speculated that a combination of double network and nanofillers

316

facilitated such good elastic and mechanical properties.

17

(a)

TOCN0 TOCN1

Stress /kPa

TOCN3

150

TOCN5

100 50 TOCNx/PAM/Gelatin10

0

250

0

200

400 600 Strain / %

800

Gelatin3 Gelatin5

Stress /kPa

Gelatin7

150

Gelatin10

100 50

TOCN5/PAM/Gelatiny

0

317

400

800 1200 Strain / %

1600

2000

TOCNx/PAM/Gelatin10

30 20 10

50

(c)

200

0

(b) 40

0

1000

Elastic modulus (Kpa)

200

50

Elastic modulus (Kpa)

250

TOCN0

(d)

TOCN1

TOCN3

TOCN5

TOCN5/PAM/Gelatiny

40 30 20 10 0

Gelatin3 Gelatin5 Gelatin7 Gelatin10

318

Figure.8. (a) Tensile stress−strain curves of NC gels and (b) the corresponding elastic modulus. All

319

gels were tested at the as-prepared state without swelling equilibrium in water solution.

320

4. Conclusions

321

In this study, cellulose nanofibers were isolated through TEMPO-oxidation pre-treatment

322

followed by mechanical disintegration. And then a series kinds of shape memory

323

TOCN/PAM/Gelatin hydrogels with varying TOCNs and gelatin contents were prepared

324

through a facile in-situ free-radical polymerization method. After the polymerization, TOCNs,

325

gelatin, and PAM were distributed homogeneously in the hydrogels. Importantly, there was

326

no new covalent bond formed between gelatin, PAM, and TOCNs. As to the shape memory

327

properties of TOCN/PAM/Gelatin composite hydrogels, the reversible nature of the gelatin

328

network imparts the TOCN/PAM/Gelatin composite hydrogels with fast cooling-fixed (ice

329

water for 30s) and thermo-induced shape recovery (in 90 °C water for 5s) capability. Good 18

330

mechanical properties (strength >200 kPa, strain >650%) were achieved. As TOCNs content

331

increased, the maximum tensile stress (b), and elastic modulus (E) of the DN gels

332

consistently increased. The introduction of gelatin contributed to increases in the elastic

333

modulus and tensile strength, further revealing the enhanced compatibility of the gelatin and

334

the TOCN/PAM system. Such a composite hydrogel with good shape memory behavior and

335

enhanced mechanical strengths would be an attractive candidate for a wide variety of

336

applications.

337

Acknowledgements

338

The authors are extremely grateful to financial support from the Science and Technology

339

Project of Guangzhou City in China (No.2016070220045), State Key Laboratory of Pulp and

340

Paper Engineering (2016 c01), State Key Laboratory of Pulp and Paper Engineering

341

(2015TS01), Natural Science Foundation of China (No.201404042), Shandong Natural

342

Science Foundation (No. ZR2015PC016), Science and Technology Planning Project of

343

Higher Education of Shandong Province (No. J14LD51), Doctoral Starting up Foundation of

344

Qufu Normal University (No. BSQD2012058), and Science and Technology Planning Project

345

of Qufu Normal University (No. xkj201413).

346 347

References

348 349 350 351 352 353 354 355

Behl, M., Razzaq, M. Y., & Lendlein, A. (2010). Multifunctional Shape memory Polymers. Advanced Materials, 22(31), 3388-3410. Bohidar, H. B., & Jena, S. S. (1993). Kinetics of sol-gel transition in thermoreversible gelation of gelatin. Journal of Chemical Physics, 98(11), 8970-8977. Chen, Y. N., Peng, L., Liu, T., Wang, Y., Shi, S., & Wang, H. (2016). Poly(vinyl alcohol)-Tannic Acid Hydrogels with Excellent Mechanical Properties and Shape Memory Behaviors. ACS Applied Materials & Interfaces, 8(40). Gong, J. P. (2010). Why are double network hydrogels so tough? Soft Matter, 6(12), 2583-2590. 19

356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399

Gornall, J. L., & Terentjev, E. M. (2008). Helix-coil transition of gelatin: helical morphology and stability. Soft Matter, 4(3), 544-549. Habibi, Y., Lucia, L. A., & Rojas, O. J. (2010). Cellulose nanocrystals: chemistry, self-assembly, and applications. Chemical Reviews, 110(6), 3479. Hao, J., & Weiss, R. A. (2013). Mechanically Tough, Thermally Activated Shape Memory Hydrogels. ACS Macro Letters, 2(1), 86-89. Harmon, M. E., Tang, M., & Frank, C. W. (2003). A microfluidic actuator based on thermoresponsive hydrogels. Polymer, 44(16), 4547-4556. Hebeish, A., Farag, S., Sharaf, S., & Shaheen, T. I. (2014). Thermal responsive hydrogels based on semi interpenetrating network of poly(NIPAm) and cellulose nanowhiskers. Carbohydrate Polymers, 102(4), 159-166. Huang, J., Zhao, L., Wang, T., Sun, W., & Tong, Z. (2016). NIR Triggered Rapid Shape Memory PAM-GO-Gelatin Hydrogels with High Mechanical Strength. ACS Applied Materials & Interfaces, 8 (19), 12384–12392. Huang, W. M., Song, C. L., Fu, Y. Q., Wang, C. C., Zhao, Y., Purnawali, H., Lu, H. B., Tang, C., Ding, Z., & Zhang, J. L. (2012). Shaping tissue with shape memory materials. Advanced Drug Delivery Reviews, 65(4), 515-535. Kabir, M. H., Hazama, T., Watanabe, Y., Jin, G., Murase, K., Sunada, T., & Furukawa, H. (2014). Smart hydrogel with shape memory for biomedical applications. Journal of the Taiwan Institute of Chemical Engineers, Volume 45(Issue 6), 3134-3138. Li, G., Yan, Q., Xia, H., & Zhao, Y. (2015). Therapeutic-Ultrasound-Triggered Shape Memory of a Melamine-Enhanced Poly(vinyl alcohol) Physical Hydrogel. ACS Applied Materials & Interfaces, 7(22), 12067. Meng, H., & Li, G. (2013). A review of stimuli-responsive shape memory polymer composites. Polymer, 54(9), 2199-2221. Saito, T., Kimura, S., Nishiyama, Y., & Isogai, A. (2007). Cellulose nanofibers prepared by TEMPO-mediated oxidation of native cellulose. Biomacromolecules, 8(8), 2485-2491. Wang, Q. Q., Zhu, J. Y., Reiner, R. S., Verrill, S. P., Baxa, U., & Mcneil, S. E. (2012). Approaching zero cellulose loss in cellulose nanocrystal (CNC) production: recovery and characterization of cellulosic solid residues (CSR) and CNC. Cellulose, 19(6), 2033-2047. Wei, J., Chen, Y., Liu, H., Du, C., Yu, H., & Zhou, Z. (2016). Thermo-responsive and compression properties of TEMPO-oxidized cellulose nanofiber-modified PNIPAm hydrogels. Carbohydrate Polymers, 147, 201-207. Xiao, H., Lu, W., Le, X., Ma, C., Li, Z., Zheng, J., Zhang, J., Huang, Y., & Chen, T. (2016). A multi-responsive hydrogel with a triple shape memory effect based on reversible switches. Chemical Communications, 52(90). Xiao, Y. Y., Gong, X. L., Kang, Y., Jiang, Z. C., Zhang, S., & Li, B. J. (2016). Light-, pH- and thermal-responsive hydrogels with the triple-shape memory effect. Chemical Communications, 52(70). Xu, X., Liu, F., Jiang, L., Zhu, J. Y., Haagenson, D., & Wiesenborn, D. P. (2013). Cellulose nanocrystals vs. cellulose nanofibrils: a comparative study on their microstructures and effects as polymer reinforcing agents. ACS Applied Materials & Interfaces, 5(8), 2999-3009. Zhang, E., Wang, T., Hong, W., Sun, W., Liu, X., & Tong, Z. (2014). Infrared-driving actuation based on bilayer graphene oxide-poly(N-isopropylacrylamide) nanocomposite hydrogels. Journal of Materials Chemistry A, 2(37), 15633-15639. 20

400 401 402 403 404

Zhao, T., Tan, M., Cui, Y., Deng, C., Huang, H., & Guo, M. (2014). Reactive macromolecular micelle crosslinked highly elastic hydrogel with water-triggered shape-memory behaviour. Polymer Chemistry, 5(17), 4965-4973. Zhou, C., & Wu, Q. (2011). A novel polyacrylamide nanocomposite hydrogel reinforced with natural chitosan nanofibers. Colloids & Surfaces B Biointerfaces, 84(1), 155-162.

21