Accepted Manuscript Title: Starch and cellulose nanocrystals together into thermoplastic starch bionanocomposites Author: Kizkitza Gonz´alez Alo˜na Retegi Alba Gonz´alez Arantxa Eceiza Nagore Gabilondo PII: DOI: Reference:

S0144-8617(14)00965-5 http://dx.doi.org/doi:10.1016/j.carbpol.2014.09.055 CARP 9313

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

14-3-2014 11-9-2014 12-9-2014

Please cite this article as: Gonz´alez, K., Retegi, A., Gonz´alez, A., Eceiza, A., and Gabilondo, N.,Starch and cellulose nanocrystals together into thermoplastic starch bionanocomposites, Carbohydrate Polymers (2014), http://dx.doi.org/10.1016/j.carbpol.2014.09.055 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.

STARCH AND CELLULOSE NANOCRYSTALS TOGETHER INTO

2

THERMOPLASTIC STARCH BIONANOCOMPOSITES

3

Kizkitza Gonzáleza, Aloña Retegia, Alba Gonzálezb, Arantxa Eceizaa, Nagore

4

Gabilondoa*

ip t

1

5

a

6

Polytechnic School, University of the Basque Country. Pza. Europa 1. 20018 Donostia-

7

San Sebastián, Spain

cr

us

b

POLYMAT, Department of Polymer Science and Technology. Faculty of Chemistry,

an

8

‘Materials + Technologies’ Group, Dept. of Chemical and Environmental Engineering,

University of the Basque Country. P.O. Box 1072, 20080 Donostia-San Sebastián,

10

Spain

11 12

*To whom all correspondence should be addressed. Tel.: (+34)943017231/7162; fax:

13

(+34)943017200; e-mail address: [email protected] (Nagore Gabilondo)

15 16 17 18

d

te Ac ce p

14

M

9

19 20 21

-1Page 1 of 33

ABSTRACT

23

In the present work, thermoplastic maize starch based bionanocomposites were prepared

24

as transparent films, plasticised with 35% of glycerol and reinforced with both waxy

25

starch (WSNC) and cellulose nanocrystals (CNC), previously extracted by acidic

26

hydrolysis. The influence of the nanofiller content was evaluated at 1 wt.%, 2.5 wt.%

27

and 5 wt.% of WSNC. The effect of adding the two different nanoparticles at 1 wt.%

28

was also investigated. As determined by tensile measurements, mechanical properties

29

were improved at any composition of WSNC. Water vapour permeance values

30

maintained constant, whereas barrier properties to oxygen reduced in a 70%, indicating

31

the effectiveness of hydrogen bonding at the interphase. The use of CNC or CNC and

32

WSNC upgraded mechanical results, but no significant differences in barrier properties

33

were obtained. A homogeneous distribution of the nanofillers was demonstrated by

34

atomic force microscopy, and a shift of the two relaxation peaks to higher temperatures

35

was detected by dynamic mechanical analysis.

36

KEYWORDS

37

Bionanocomposite; Thermoplastic starch; Polysaccharide nanocrystals; Morphology;

38

Thermal, mechanical and barrier properties.

40

cr

us

an

M

d

te

Ac ce p

39

ip t

22

41 42 43

-2Page 2 of 33

1. INTRODUCTION

45

Starch, in form of its thermoplastic derivate (TPS), has been revealed as an appropriate

46

candidate to be employed as substitute of synthetic polymers traditionally used for

47

packaging, due to its versatility, availability, biodegradability, processability and low

48

cost. Starch is a hydrophilic semicrystalline biopolymer stored in form of granules in

49

the native state that consists of a mixture of two glycosidic macromolecules, i.e.

50

amylose and amylopectin (Carvalho, 2008; Le Corre, Bras, & Dufresne, 2010).

51

Amylose is a linear polysaccharide made up of (1-4) -D-glycopyranose, whereas

52

amylopectin is a highly branched macromolecule composed of both (1-4) and (1-6)

53

glycopyranosyl linkages (García, Ribba, Dufresne, Aranguren, & Goyanes, 2011; Le

54

Corre et al. 2010). The arrangement of the two polymers in form of alternant amorphous

55

and crystalline rings lead to the final semicrystalline granules, being amylopectin the

56

main responsible of the crystalline character (Le Corre et al. 2010).

57

Actually, starch is not a real thermoplastic polymer, but can be processed after its

58

gelatinization by mixing with enough water and/or plasticizers (Angellier, Molina-

59

Boisseau, Dole, & Dufresne, 2006; Carvalho, 2008; García et al., 2011; Viguié, Molina-

60

Boisseau, & Dufresne, 2007). The crystallinity of starch granules disappears during the

61

gelatinization process, and at the end, an amorphous material is obtained (Carvalho,

62

2008). In most investigations the plasticisation of the material is carried out by casting a

63

dispersion of starch with glycerol (Angellier et al., 2006; Bertuzzi, Vidaurre, Armada, &

64

Gottifredi, 2007; Carvalho, 2008; García, Ribba, Dufresne, Aranguren, & Goyanes,

65

2009; Yan, Hou, Guo, & Dong, 2012), but other water soluble plasticizers such as

66

sorbitol, xylitol, maltitol, urea or ethylene glycol have been also used (Abdorreza,

67

Cheng, & Karim, 2011; Da Róz, Carvalho, Gandini, & Curvelo, 2006; Vigué et al.,

68

2007; Wang, Cheng, & Zhu, 2014). Attempting to reduce the undesired well-known

Ac ce p

te

d

M

an

us

cr

ip t

44

-3Page 3 of 33

sensitivity to moisture and temperature of TPS, several strategies have been developed

70

in the literature (Averous, Moro, Dole, & Fringant, 2000; Xie, Pollet, Halley, &

71

Avérous, 2013).

72

Bionanocomposites have already been identified as useful materials for achieving the

73

mentioned objectives. The addition of different nanocrystals to the starch based matrix,

74

would lead to both mechanical and barrier properties improvement of the material. In

75

this context, nanoclays and polysaccharide derived nanocrystals have been tried in the

76

most recent researches (Angellier et al., 2006; García et al., 2011; Xie et al., 2013).

77

Actually, a large amount of work has been published mainly with cellulose whiskers

78

and, in a less extent, chitin or starch nanocrystals (Wang, Tian, & Zhang, 2010).

79

Disordered amorphous domains of natural polysaccharides can be removed by acid

80

hydrolysis (Habibi, Lucia, & Rojas, 2010; Lin, Huang, Chang, Anderson, & Yu, 2011a;

81

Lin, Huang, & Dufresne, 2012), under controlled conditions, leaving the crystalline

82

regions intact (Lin et al., 2012), and thus, nanometrical high crystallinity particles are

83

obtained. However, little work has been reported using different polysaccharide

84

nanocrystals simultaneously (Wang et al., 2010). Nanocrystals extracted from different

85

sources have a variety of geometrical structures (Wang et al., 2010), and thus, they are

86

supposed to have their own properties and applications as reinforcing agent. While

87

cellulose nanocrystals were found to be rod like or fibrillar (Angellier, Molina-

88

Boisseau, Belgacem, & Dufresne, 2005; Lin et al., 2012; Wang et al., 2010), starch

89

nanocrystals isolated from waxy maize presented a platelet like morphology (Habibi et

90

al., 2010; Lin et al., 2012; Saralegi, et al. 2013) and would be preferred for barrier

91

properties enhancement (LeCorre, Bras, & Dufresne, 2012). However, the expected

92

improvement of mechanical, thermal or barrier properties by the addition of

Ac ce p

te

d

M

an

us

cr

ip t

69

-4Page 4 of 33

nanocrystals extracted from different origins would depend on both a correct dispersion

94

and the generation of an active interphase nanoreinforcement/matrix.

95

One of the most common applications of TPS based biocomposites is their use as films

96

in food packaging products. A critical issue in food packaging is that of migration of

97

certain components and vapour and gases permeability (Duncan, 2011). Many authors

98

have investigated the improvement of water vapour and oxygen permeability in

99

reinforced TPS materials (Bertuzzi et al., 2007; Chivrac, Angellier-Coussy, Guillard,

100

Pollet, & Avérous, 2010; García et al., 2009; Kelnar, Kaprálková, Brozová,

101

Hromádkova, 2013; Xie et al., 2013). The barriers to oxygen and water vapour are two

102

essential properties to consider in starch-based material because these can deteriorate

103

food properties. The gas transport trough the film depends hardly on the diffusivity.

104

Tortuous pathways for gas molecules diffusion, created by dispersed nanocrystals,

105

increase molecule migration and consequently limit permeability (Le Corre et al., 2010).

106

It is important to take into account that permeability is also affected by solubility, which

107

is related with the chemical nature and crystallinity of the polymer, and sample

108

homogeneity. Polysaccharides, such as TPS, are well known to be effective barrier

109

materials to oxygen transport, whereas they present high water permeability values

110

(Bertuzzi et al., 2007).

111

The main objective of the present work was to investigate the effect of different

112

contents of waxy maize starch nanocrystals on a normal maize starch matrix plasticized

113

by glycerol. In addition, the influence of using together nanocrystals from similar

114

chemical nature with different geometries, i.e. waxy maize starch and cellulose

115

nanocrystals, was studied in order to investigate possible synergistic effects. The

116

thermoplastic starch was obtained by casting process and both polysaccharide

117

nanocrystals were extracted by acid hydrolysis. The nanometrical dimensions and the

Ac ce p

te

d

M

an

us

cr

ip t

93

-5Page 5 of 33

degree of crystallinity of nanocrystals were investigated. Resulting nanobiocomposites

119

were characterized in terms of thermal stability, mechanical properties and barrier

120

properties.

121

2. EXPERIMENTAL SECTION 2.1. Materials

cr

122

ip t

118

Waxy maize starch and normal corn starch were purchased from Sigma-Aldrich and

124

glycerol (99% purity) from Panreac. The former, almost completely amylopectin, was

125

employed for the preparation of starch nanocrystals (WSNC) and the latter, with higher

126

content of amylose, nearly 25%, was used for the preparation of thermoplastic starch

127

(TPS).

an

M

128

us

123

2.2. Preparation of waxy starch nanocrystals (WSNC) The preparation of the waxy starch nanocrystals (WSNC) was carried out by acidic

130

hydrolysis according to the method described elsewhere (Angellier, Choisnard, Molina-

131

Boisseau, Ozil, & Dufresne, 2004; García et al., 2011). 36.725 g of waxy starch maize

132

were mixed with 250 ml of H2SO4 solution (3.16M) during 5 days at 40 °C under

133

continuous magnetic stirring. The suspension was washed and centrifuged with distilled

134

water until neutralization. The obtained nanocrystals were stored at 4 °C after adding

135

some drops of chloroform.

te

Ac ce p

136

d

129

2.3. Preparation of cellulose nanocrystals (CNC)

137

The extraction of the crystalline part of cellulose was also performed by acidic

138

hydrolysis as explained by Saralegi et al. (2013). 5 g of microcrystalline cellulose

139

(MCC) were treated with a H2SO4 64 wt.% aqueous solution. The reaction was

140

completed during 30 min at 45 °C with continuous stirring. After that, the acidity was

-6Page 6 of 33

141

reduced by centrifugation (twice, 4500 rpm, 20 min) and the precipitate was neutralized

142

by dialysis (4-5 days). Finally, the suspension was ultrasonicated for 15 min.

143

2.4. Preparation of thermoplastic starch (TPS) The thermoplastic starch and the subsequent nanocomposites were obtained by casting

145

based on the method described by Angellier et al. (2006) with some modifications. A

146

mixture of 3.58 g of normal maize starch, 1.93 g of glycerol and 35 g of distilled water

147

was gelatinized at 90 °C with continuous stirring until viscosity increased (about 20

148

min). After that, the material was spread into glass petri and dried in an oven at 55 °C

149

for 2 hours (CSg35). For the preparation of nanocomposites, the desired amount of

150

nanocrystals was added, after their previous dispersion in distilled water by

151

ultrasonication. In order to avoid the gelatinization of starch nanocrystals, the mixture

152

was cooled down below 50 °C before adding the nanoreinforcements. Thereafter, the

153

processing was carried out as described above (CSg35+wt.%). All the samples were

154

stored at 43% relative humidity (RH) with a K2CO3 saturated solution for two weeks

155

before all characterization tests. A glycerol content of 35 wt.% was used for all samples,

156

i.e. relative to the starch plus water mixture weight. The nanofiller content was relative

157

to the starch plus glycerol weight and WSNC or CNC contents were relative to the total

158

nanocrystals weight. The samples were named as CSg35 plus the nanocrystals content,

159

followed by the WSNC and/or CNC proportion. Bionanocomposites are referred

160

hereafter as shown in Table 1.

Ac ce p

te

d

M

an

us

cr

ip t

144

Nanocrystal (wt.%) 0

WSNC (%) 0

CNC (%) 0

CSg35+1 (100/0)

1

100

0

CSg35+1 (50/50)

1

50

50

CSg35+1 (0/100)

1

0

100

CSg35

-7Page 7 of 33

CSg35+2.5 CSg35+5 161

163

100

0

5

100

0

Table 1. Compositions of obtained bionanocomposites.

2.5. Characterization 2.5.1.

Atomic force microscopy (AFM)

ip t

162

2.5

Atomic force microscopy (AFM) measurements were performed in tapping mode using

165

a Veeco Multimode scanning probe microscope equipped with a Nanoscope IIIa

166

Controller. All measurements were recorded using TESP type silicon tips having a

167

resonance frequency of approximately 340 kHz and a cantilever spring constant about

168

40 N/m.

169

analyzed.

an

us

2.5.2.

Both nanocrystals and cryofractured surfaces of nanocomposites were

M

170

cr

164

Optical microscopy (OM)

The study of the gelatinization process, i.e. the transformation of starch into a

172

thermoplastic and amorphous material, was performed by optical microscopy, analyzing

173

the change of the crystalline structure and the evolution of the birefringence. The

174

analyses were carried out with a Nicon Eclipse E600 instrument in reflectance and with

175

a magnification of x500.

te

Ac ce p

176

d

171

2.5.3.

X-ray diffraction (XRD)

177

All samples were submitted to X-ray radiation using an Xpert diffractometer, 40kV,

178

40mA. Scattered radiation was detected in the angular range 1-40° (2). The XRD

179

curves were used to determine the crystallinity percentage of WSNC and CNC. For

180

WSNC an appropriate method is not so widely established, for this reason the

181

crystallinity degree was determined by the ratio of the crystalline area to the total area

182

(Jivan, Madadlou, &Yarmand, 2013; Rosa et al., 2010). -8Page 8 of 33

WSNC, Crystallinity percentage (%) =

183

Area under the peaks × 100 Total area

Crystallinity for CNC was calculated according to the method described in literature

185

(Kargarzadeh et al., 2012; Segal, Creely, Martin, & Conrad, 1959).

186

  002 −   × 100  002

cr

CNC, Crystallinity percentage (%) =

ip t

184

Where I002 is the peak intensity of the (002) crystallographic plane (Kargarzadeh et al.,

188

2012), i.e. the intensity peak near 2= 22.5° (Lin, Huang, Chang, Feng, & Yu, 2011b)

189

and Iam is the intensity scattered by the amorphous part of the sample measured around

190

2= 18.0° (Kargarzadeh et al., 2012).

an

2.5.4.

Thermogravimetric analysis (TGA)

M

191

us

187

Thermogravimetric analysis was carried out in order to investigate the degradation

193

process of all nanocomposites using a Mettler Toledo TGA-SDTA 851 instrument. The

194

samples were heated from room temperature to 800 °C, with a heating rate of 10 °C

195

min-1 and under nitrogen.

te

Ac ce p

196

d

192

2.5.5.

Tensile tests

197

The mechanical behaviour of the nanocomposites was analyzed using a MTS Insight 10

198

instrument. Dumbbell shaped specimens 5 mm wide and 15 mm long (L0) were used.

199

The initial gap between jaws was adjusted to 22 mm.

200

2.5.6.

Dynamic mechanical analysis (DMA)

201

Dynamic mechanical analysis was carried out using an Eplexor 100N instrument from

202

Gabo Qualimeter. The specimen was a rectangular strip (25×3×0.2 mm3), it was

203

subjected to a 0.05% constant strain and the setup measured the stored modulus E’ the -9Page 9 of 33

204

loss modulus E’’ and the ratio between the two components (tan=E’’/E’).

205

Measurements were performed at 1 Hz, working from -100 °C to 150 °C with a heating

206

rate of 2 °C min-1.

207

2.5.7.

ip t

Water vapour permeability (WVP)

Water vapour permeation experiments were performed at 25 °C on a gravimetric cell in

209

which a small amount of liquid water was sealed by a membrane. The cell was put on a

210

balance with a readability of 10-5 g, and the weight loss of the cell, solely due to the

211

permeation of the water vapour through the membrane, was followed by means of a

212

computer connected to the balance.

us

an

2.5.8.

Oxygen permeability (OP)

M

213

cr

208

The measurements were carried out using a MOCON OX-TRAN Model 2/21 gas

215

permeability tester in accordance with the ASTM standard D3985. The permeability of

216

samples was tested at 760 mmHg, HR= 50% and 23 °C.

217

3. RESULTS AND DISCUSSION

te

Ac ce p

218

d

214

3.1. Characterization of polysaccharide nanocrystals

219

Geometrical structure and dimensions at the nanoscale of the nanocrystals were studied

220

by Atomic force microscopy (AFM) (Figure 1a and 1b). Assuming the cylindrical shape

221

of CNC (Saralegi et al., 2013) mean diameter values were obtained by the evaluation of

222

AFM height profiles. For WSNC, since they trend to agglomerate, less single

223

nanocrystals were possible to measure. Nanometrical dimensions of both waxy starch

224

and cellulose nanocrystals were confirmed. Waxy starch nanoparticles were found to be

225

platelet like, with 28.2  7.4 nm in width and 35  7.9 nm in length. Cellulose

226

nanocrystals presented the typical fibril like geometry with 9.1  2.6 nm in diameter and -10Page 10 of 33

150.6  29.1 nm in length. These results agreed with those obtained in literature

228

(Angellier et al., 2005; Wang et al., 2010; Saralegi et al., 2013).

cr

ip t

227

us

229

Figure 1. Characterization of polysaccharide nanoreinforcements: AFM images of extracted a) WSNC and b) CNC and c) X-ray diffraction patterns for nanocrystals.

232

X-ray diffraction (XRD) measurements were carried out in order to observe the

233

crystalline polymorphism for each type of nanocrystal (Figure 1c). XRD pattern of

234

WSNC showed the typical A-type polymorphism with two peaks of low intensity

235

located at 2=10.1° and 2=11.5°, a peak at 2=15.2° of higher intensity, a double peak

236

near 2=17.3°/18.1° and an intense peak at 23.2°. Cellulose nanocrystals pattern

237

exhibited strong peaks at 2=14.5°, 2=16.8° and 2=22.7° associated to the typical

238

cellulose I crystalline structure. As explained by Lin et al. (2012), the crystallinity

239

degree of polysaccharide nanocrystals should be 100%. However, disorder or

240

amorphous domains cannot be completely removed by hydrolysis, leading to a lower

241

crystallinity degree. Following the method described elsewhere (Jivan et al., 2013; Rosa

242

et al., 2010), a crystallinity percentage of 22.8% for WSNC was established. Besides,

243

according to the method reported previously (Kargarzadeh et al., 2012; Segal et al.,

244

1959) a crystallinity percentage of 83.6% was calculated for CNC. Other authors

245

(Buléon, Colonna, Planchot, & Ball, 1998; Lin et al., 2012) determined similar values

246

for CNC obtained from MCC (54-88%), but higher crystallinity degree percentages for

247

WSNC (38-48%). Differences could be related to the different extraction protocol

Ac ce p

te

d

M

an

230 231

-11Page 11 of 33

employed, along with the method used for calculating crystallinity degrees (Jivan et al.,

249

2013; Rosa et al., 2010). However, even if the crystallinity percent of the WSNC could

250

be affected slightly by the calculation method, the results for CNC were, as expected,

251

significantly higher. Attending to AFM and XRD results, it was concluded that the

252

employed hydrolysis methods were satisfactory. 3.2. Characterization of bionanocomposites

cr

253

ip t

248

3.2.1. Influence of the WSNC content

255

Prior to the preparation of the bionanocomposites, the gelatinization of the matrix was

256

optimized following the disappearance of the birefringence by optical microscopy. The

257

effect of the nanoreinforcement content on thermal, mechanical and barrier properties

258

was investigated for nanocomposites of TPS reinforced with WSNC. XRD results for

259

the TPS/WSNC bionanocomposites showed that the A-type polymorphism pattern of

260

the native normal maize starch was replaced by a typical broad peak centred at

261

2=19.6° in the case of the amorphous thermoplastic derivative. However, as WSNC

262

content increased, peaks associated to the nanoreinforcements were progressively

263

detected, in concordance with results published previously in literature (Angellier et al.,

264

2006).

265

Thermogravimetric measurements were carried out to assess the thermal decomposition

266

of all nanocomposites (Figure 2). Un-plasticized normal maize starch showed a single

267

degradation step at 300 °C. On the contrary, the thermal decomposition process for as-

268

prepared nanocomposites occurs in three main steps (García et al., 2009, 2011). The

269

first step (25-100 °C) related to the loss of sorbed humidity, the second one (100-200

270

°C) associated to the decomposition of the glycerol-rich phase and the last one (below

271

340 °C) due to the oxidation of the partially decomposed starch. The degradation

Ac ce p

te

d

M

an

us

254

-12Page 12 of 33

stability decrease of the bionanocomposites may be maximized by the affinity between

273

glycerol and waxy starch nanocrystals. Acidic sulphate ester groups attached to waxy

274

starch nanocrystals surface during the extraction seemed to act as catalyzers of the

275

thermal decomposition, resulting in a clear decrease of degradation temperature for

276

nanocomposites compared with that for the unfilled matrix. At the same time, the final

277

char percentage increased with the nanoreinforcement content.

278

d

M

an

us

cr

ip t

272

Figure 2. TGA curves for a) Normal maize starch b) CSg35, c) CSg35+1 (100/0), d) CSg35+2.5 and e) CSg35+5.

281

The mechanical properties of bionanocomposites reinforced with different contents of

282

WSNC were determined by means of tensile tests. Obtained results are collected in

283

Table 2. The addition of WSNC resulted in a significant improvement of the mechanical

284

properties of the TPS. Indeed, the Young’s Modulus increased from 2.5 MPa for the

285

plasticized starch to 7.5 MPa for the bionanocomposites, the tensile strength was

286

improved in a 70%, with only a slight variation in deformability at break. Due to the

287

identical chemical nature of the matrix and the nanofiller, a good affinity between both

288

components was expected, allowing strong TPS/WSNC hydrogen bonding at the

289

interfaces and the subsequent favourable reinforcing effect. However, it is worth noting

290

that mechanical properties of bionanocomposites were not further improved from 2.5

Ac ce p

te

279 280

-13Page 13 of 33

wt.% to 5 wt.% of WSNC content. The reinforcing effect of starch nanocrystals is

292

generally ascribed to the formation of a hydrogen bonded percolating filler network

293

above a given starch content corresponding to the percolation threshold (Le Corre et al.,

294

2010). Our results seemed to indicate that starch nanocrystals were correctly dispersed

295

in an effective percolated network up to a 2.5 wt.% of reinforcement.

ip t

291

Water Vapour Permeance (kg/m2 s Pa)1010 56.2  2.0

Oxygen Permeance (cm3/m2 day atm)

66.7  4.5

26.9  1.6

Strain at break (%)

Tensile strength (MPa)

CSg35 CSg35+1 (100/0) CSg35+2.5

2.5  0.9

52.7  9.4

1.0  0.2

5.8  0.9

48.8  24.4

1.7  0.4

7.4  1.1

41.8  8.7

1.6  0.3

57.3  3.8

15.5  4.7

CSg35+5

7.5  1.1

47.9  10.3

1.6  0.1

57.3  1.4

7.7  1.5

an

us

cr

Young’s Modulus (MPa)

22.7  3.9

Table 2. Tensile properties and water vapour and oxygen permeance values for plasticized starch and TPS-WSNC nanocomposites.

298

Gas and vapour permeability of polymeric materials is independent of film thickness in

299

ideal systems. However, TPS based films present non-ideal behaviour with increasing

300

permeability values proportionally to the film thickness providing an increasing

301

resistance to mass transfer trough the film (Bertuzzi et al., 2007). In our study such

302

deviation was detected, being the relationship between thickness and water vapour

303

permeability almost linear (R2>0.98). For this reason, both for water vapour and oxygen

304

mass transfer, the calculation of permeance values was preferred. Obtained results are

305

collected in Table 2. Nanocomposites development respond to the hypothesis that

306

nanometric particles incorporated into a polymeric matrix would generate a tortuous

307

path for vapour or gas and, consequently, the diffusive component of the permeability

308

would decrease. Besides the diffusion, polymers barrier properties depends on other

309

factors as polarity and structural features of polymeric side chains, hydrogen bonding

310

characteristics, molecular weight and polydispersity, degree of branching or cross-

Ac ce p

te

d

M

296 297

-14Page 14 of 33

linking, processing methodology and degree of crystallinity (Duncan, 2011). Platelet-

312

like starch nanocrystals are supposed to have good attributes in the improvement of

313

barrier properties but, according to water vapour permeability, due to their hydrophilic

314

nature, low crystallinity and agglomeration opposite conclusions have been reported for

315

their nanocomposites (Le Corre et al., 2010). In our particular investigation, whereas

316

water permeance values were not reduced by the addition of starch nanocrystals, oxygen

317

permeance was significantly improved for 2.5 wt.% and 5 wt.% of WSNC contents.

318

Bertuzzi et al. (2007) and Forssell et al. (2002) correlated these results with sorbed

319

water-starch interactions that are closely related with water transport phenomena. These

320

interactions were not generated with oxygen molecules, and that is why oxygen

321

permeance values were improved. As it could be noted, whereas CSg35 and CSg35+1

322

(100/0) presented similar permeance to O2, it was reduced to 7.7  1.5 cm3/m2 day atm,

323

for 5 wt.% of WSNC content, suggesting that the presence of a minimum concentration

324

was necessary to create an actual tortuous pathway. This significant improvement for

325

higher WSNC contents suggests that those were effectively dispersed in the TPS matrix

326

and points out the close interface adhesion between nanocrystals and matrix.

327

The viscoelastic behaviour of the bionanocomposites was assessed by Dynamic

328

mechanical analysis (Figure 3). Two-step modulus drop relative to the tan  peaks

329

temperatures near -60 °C (Tα1) and -20 °C (Tα2) was observed. García et al. (2009)

330

attributed them to the glycerol-rich phase and starch-rich phase relaxation temperatures,

331

respectively. As it can be noted, the addition of 1 wt.% of WSNC resulted in a

332

significant shift of Tα1 which is related to an increase in the rigidity of the glycerol-rich

333

phase. As the WSNC content increased, the peak shifted to temperatures close to -50 ºC.

334

At the same time, Tα2 also moved to higher temperatures mainly in bionanocomposites

335

filled with 2.5 wt.% and 5 wt.%, revealing that a good dispersion of the nanocrystals

Ac ce p

te

d

M

an

us

cr

ip t

311

-15Page 15 of 33

was achieved and the reinforcement effect was homogeneously distributed in both

337

phases. Thus, DMA results were in good agreement with results obtained from both

338

tensile and permeance measurements.

an

us

cr

ip t

336

339

Figure 3. Evolution of E’ and tan  with temperature for a) CSg35, b) CSg35+1 (100/0), c) CSg35+2.5, d) CSg35+5.

M

340 341

3.2.2. Influence of nanocrystals’ origin on bionanocomposites

343

The influence of nanocrystal nature was analysed when WSNC and CNC were

344

combined together in TPS. 1 wt.% composition was chosen in order to better analysing

345

any synergistic effect resulting from the simultaneous action of both nanocrystals. This

346

assumption lays on the fact that the platelet shape of WSNC would enhance barrier

347

properties improvement (Duncan, 2010), whereas rod-like high crystalline CNC would

348

contribute in a higher extent to the mechanical reinforcement as they present higher

349

aspect ratio, reducing the percolation threshold. Four different WSNC-CNC nanocrystal

350

contents were studied (Table 1).

351

Mechanical behaviour was evaluated by means of tensile tests and obtained results are

352

shown in Table 3. Some remarkable changes were detected when cellulose or waxy

353

starch nanocrystals were added into the matrix. Firstly, higher values of Modulus were

354

obtained for CNC filled nanocomposites at the same composition. This finding agrees

Ac ce p

te

d

342

-16Page 16 of 33

with the fact that higher mechanical reinforcement would be achieved by using high

356

aspect ratio rod-like nanocrystals (Flauzino Neto, Silvério, Dantas, & Pasquini, 2013;

357

Wang et al., 2010). The opposite effect was detected for deformation at break, as it was

358

reduced in a larger extent. It is worth noting that the Young’s Modulus was

359

considerably improved up to 12.0 MPa for the CSg35+1 (50/50) system, i.e. the one

360

where WSNC and CNC were combined. Wang et al. (2010) obtained similar results

361

using waterborne polyurethane as matrix. As they explained, the combination of both

362

nanometric reinforcements would lead to a much jammed hydrogen bonding network

363

between nanofiller and the matrix. However, taking into account the variability of the

364

measurements, in our opinion these findings should be corroborated in further

365

investigations.

366 367

M

Oxygen Permeance (cm3/m2 day atm)

1.0  0.2

Water Vapour Permeance (kg/m2 s Pa)1010 56.2  2.0

Strain at break (%)

Tensile Strength (MPa)

2.5  0.9

52.7  9.4

te

d

Young’s Modulus (MPa)

22.7  3.9

5.8  0.9

48.8  24.4

1.7  0.4

66.7  4.5

26.9  1.6

12.0  2.0

41.7  4.2

1.6  0.1

54.2  1.1

26.6  1.9

8.5  2.2

41.9  8.5

1.5  0.2

57.9  2.7

22.1  1.1

Ac ce p

CSg35 CSg35+1 (100/0) CSg35+1 (50/50) CSg35+1 (0/100)

an

us

cr

ip t

355

Table 3. Tensile properties and water vapour and oxygen permeance values for TPS and TPS+WSNC/CNC.

368

Table 3 summarized the results of water vapour and oxygen permeance obtained for the

369

bionanocomposites containing WSNC and/or CNC. Large nanoparticle aspect ratios are

370

required to reduce the gas permeability by an appreciable degree (Duncan, 2011).

371

Attending to the results, it could be concluded that the incorporation of CNC to the

372

bionanocomposite did not result in any improvement of the barrier properties. On one

373

hand, water vapour permeability seemed to be again governed by the strong chemical

374

affinity of the material with water molecules. On the other hand, the higher aspect ratio -17Page 17 of 33

and crystallinity of CNC did not reduce the oxygen permeance values (only slight

376

decrease is noting for the CSg35+1 (0/100) sample), revealing that in studied

377

polysaccharide-based films, barrier properties improvement was more correlated to the

378

nanofiller content than to the origin of it.

379

Nanocomposites reinforced with waxy starch and cellulose nanocrystals showed the

380

same rheological behaviour in comparison with films reinforced only with WSNC

381

(Figure 4). Indeed, almost identical DMA curves were obtained with the characteristic

382

two-step modulus drop related to the two different phases. The addition of

383

polysaccharide nanocrystals resulted in higher relaxation temperatures (T) values

384

whatever its origin was. This effect could be ascribed to the fact that both starch and

385

cellulose nanocrystals interacted strongly with the glycerol-rich phase. However, a

386

remarkable higher shift of the Tα1 value to higher temperatures was observed for

387

bionanocompostes containing WSNC. On the contrary, Tα2 maintained almost constant.

388

These results confirmed the interaction between glycerol-rich phase domains and the

389

reinforcing nanocrystals, especially with waxy starch nanocrystals.

Ac ce p

te

d

M

an

us

cr

ip t

375

390 391 392

Figure 4. Evolution of the E’ and tan  with temperature for a) CSg35, b) CSg35+1 (100/0), c) CSg35+1 (50/50), d) CSg35+1 (0/100).

-18Page 18 of 33

AFM was used for the morphological characterization of the cryofractured

394

nanocomposite samples. Figure 5a and 5b show the obtained images for CSg35+1

395

(100/0) and CSg35+1 (0/100), respectively. In case of the CSg35+1 (100/0) sample,

396

starch nanocrystals were observed as ill-defined nanoparticles due to their extremely

397

reduced dimensions. However, focusing on the amplified image where the dimensions

398

of the nanofillers were possible to evaluate individually, it could be concluded that

399

nanocrystals were correctly dispersed, in good agreement with the conclusions extracted

400

previously.

401 402 403

Ac ce p

te

d

M

an

us

cr

ip t

393

Figure 5. AFM height images of a) CSg35+1 (100/0) and b) CSg35+1 (0/100).

404

4. CONCLUSIONS

405

AFM and XRD results demonstrate that the acid hydrolysis protocols used to obtain

406

both waxy starch and cellulose nanocrystals were satisfactory. Also TOM and XRD

407

corroborated that method employed in the gelatinization to produce thermoplastic starch -19Page 19 of 33

and resulting nanocomposites was successful. TGA results defined that the thermal

409

degradation of nanocomposites occurs in three main steps. Also was demonstrated that

410

the addition of NC reduces the thermal stability of all samples. The mechanical

411

properties and oxygen permeance values of the TPS-WSNC nanocomposites are clearly

412

higher than those of the unfilled matrix. The addition of 2.5 wt.% of WSNC to the

413

starch thermoplastic matrix represents a significant improvement of the mechanical and

414

oxygen permeance properties. Mechanical properties were also enhanced with the

415

addition of CNC for CSg35+1 (50/50) and CSg35+1 (0/100) samples, due to the

416

reinforcement ability of CNC. However, DMA curves show an increased of the

417

relaxation temperature since the addition of nanocrystals result in a stronger affinity

418

between glycerol and NC phases.

419

5. ACKNOWLEDGMENTS

420

Financial support from the Basque Country Government in the frame of Saiotek S-

421

PE12UN036 and Grupos Consolidados (IT-776-13) and from the University of the

422

Basque Country (UPV/EHU) in the frame of EHUA12/19 is gratefully acknowledged.

423

Technical and human support provided by SGIker (UPV/EHU, MINECO, GV/EJ,

424

ERDF and ESF) is also gratefully acknowledged.

425

6. REFERENCES

426

Abdorreza, M. N., Cheng, L. H., & Karim, A. A. (2011). Effects of plasticizers on

427

thermal properties and heat sealability of sago starch films. Food Hydrocolloids, 25, 56–

428

60.

429

Angellier, H., Choisnard, L., Molina-Boisseau, S., Ozil, P., & Dufresne, A. (2004)

430

Optimization of the preparation of aqueous suspensions of waxy maize starch

431

nanocrystals using a response surface methodology. Biomacromolecules, 5, 1545–1551.

Ac ce p

te

d

M

an

us

cr

ip t

408

-20Page 20 of 33

Angellier, H., Molina-Boisseau, S., Belgacem, M. N., & Dufresne, A. (2005). Surface

433

chemical modification of waxy maize starch nanocrystals. Langmuir, 21, 2425–2433.

434

Angellier, H., Molina-Boisseau, S., Dole, P., & Dufresne, A. (2006). Thermoplastic

435

starch-waxy maize starch nanocrystals nanocomposites. Biomacromolecules, 7(2), 531–

436

539.

437

Angellier-Coussy, H., Putaux, J.-L., Molina-Boisseau, S., Dufresne, A., Bertoft, E., &

438

Perez, S. (2009). The molecular structure of waxy maize starch nanocrystals.

439

Carbohydrate Research, 344, 1558-1566.

440

Averous, L., Moro, L., Dole, P., & Fringant, C. (2000). Properties of thermoplastic

441

blends: starch–polycaprolactone. Polymer, 41(11), 4157–4167.

442

Bertuzzi, M. A., Vidaurre, E. F. C., Armada, M., & Gottifredi, J. C. (2007). Water vapor

443

permeability of edible starch based films. Journal of Food Engineering, 80, 972–978.

444

Buléon, A., Colonna, P., Planchot, V., & Ball, S. (1998). Starch granules: structure and

445

biosynthesis. International Journal of Biological Macromolecules, 23, 85-112.

446

Carvalho, A. J. F. (2008). Starch: Major sources, properties and applications as

447

thermoplastic materials. In M. N. Belgacem & A. Gandini (Eds.), Monomers, polymers

448

and composites from renewable resources (pp. 321-342).

449

Chivrac, F., Angellier-Coussy, H., Guillard, V., Pollet, E., & Averous, L. (2010). How

450

does

451

Carbohydrate Polymers, 82(1), 128–135.

Ac ce p

te

d

M

an

us

cr

ip t

432

water

diffuse

in

starch/montmorillonite

nano-biocomposite

materials?.

-21Page 21 of 33

Da Róz, A., Carvalho, A., Gandini, A., & Curvelo, A. (2006). The effects of plasticizers

453

on thermoplastic starch compositions obtained by melt processing. Carbohydrate

454

Polymers, 63, 417–424.

455

Duncan V. T. (2011). Applications of nanotechnology in food packaging and food

456

safety: Barrier materials, antimicrobials and sensors. Journal of Colloid and Interface

457

Science, 363, 1-24.

458

Flauzino Neto, W. P., Silvério, H. A., Dantas, N. O., & Pasquini, D. (2013). Extraction

459

and characterization of cellulose nanocrystals from agro-industrial residue-soy hulls.

460

Industrial Crops and Products, 42, 480- 488.

461

Forssell, P., Lahtinen, R., Lahelin, M., & Myllärinen, P. (2002). Oxygen permeability of

462

amylose and amylopectin films. Carbohydrate Polymers, 47, 125–129.

463

García, N. L., Ribba, L., Dufresne, A., Aranguren, M. & Goyanes, S. (2009). Physico-

464

Mechanical properties of biodegradable starch nanocomposites. Macromolecular

465

Materials and Engineering, 294, 169–177.

466

García, N. L., Ribba, L., Dufresne, A., Aranguren, M., & Goyanes, S. (2011). Effect of

467

glycerol on the morphology of nanocomposites made from thermoplastic starch and

468

starch nanocrystals. Carbohydrate Polymers, 84, 203–210.

469

Habibi, Y., Lucia, L. A., & Rojas, O. J. (2010). Cellulose nanocrystals: Chemistry, self-

470

assembly, and applications. Chemical Reviews, 110(6), 3479–3500.

471

Jivan, M. J., Madadlou, A., & Yarmand, M. (2013). An attempt to cast light into starch

472

nanocrystals preparation and cross-linking. Food Chemistry, 141, 1661-1666.

Ac ce p

te

d

M

an

us

cr

ip t

452

-22Page 22 of 33

Kargarzadeh, H., Ahmad, I., Abdullah, I., Dufresne, A., Zainudin, S. Y., & Sheltami, R.

474

M. (2012). Effects of hydrolysis conditions on the morphology, crystallinity, and

475

thermal stability of cellulose nanocrystals extracted from kenaf bast fibers. Cellulose,

476

19(3), 855–866.

477

Kelnar, I., Kaprálková, L., Brozová, L., Hromádková, J., Kotek, J. (2013). Effect of

478

chitosan on the behaviour of the wheat B-starch nanocomposite. Industrial Crops and

479

Products, 46, 186– 190.

480

Le Corre, D., Bras, J., & Dufresne, A. (2010). Starch nanoparticles: A review.

481

Biomacromolecules, 11, 1139-1153.

482

LeCorre, D., Bras, J., & Dufresne, A. (2012). Influence of native starch’s properties on

483

starch nanocrystals thermal properties. Carbohydrate Polymers, 87, 658–666.

484

Lin, N., Huang, J., Chang, P. R., Anderson D. P., & Yu, J. (2011a). Preparation,

485

Modification, and Application of Starch Nanocrystals in Nanomaterials: A Review.

486

Journal of Nanomaterials, doi: 10.1155/2011/573687.

487

Lin, N., Huang, J., Chang, P. R., Feng, L., & Yu, J. (2011b). Effect of polysaccharide

488

nanocrystals on structure, properties, and drug release kinetics of alginate-based

489

microspheres. Colloids and Surfaces B: Biointerfaces, 85, 270–279.

490

Lin, N., Huang, J., & Dufresne, A. (2012). Preparation, properties and applications of

491

polysaccharide nanocrystals in advanced functional nanomaterials: a review. Nanoscale,

492

4(11), 3274–3294.

493

Rosa, M. F., Medeiros, E. S., Malmonge, J. A., Gregorski, K. S., Wood, D. F., Mattoso,

494

L. H. C., Glenn, G., Orts, W. J., & Imam, S. H. (2010). Cellulose nanowhiskers from

Ac ce p

te

d

M

an

us

cr

ip t

473

-23Page 23 of 33

coconut husk fibers: Effect of preparation conditions on their thermal and

496

morphological behaviour. Carbohydrate Polymers, 81(1), 83–92.

497

Saralegi, A., Rueda, L., Martin, L., Arbelaiz, A., Eceiza, A., & Corcuera, M.A. (2013).

498

From elastomeric to rigid polyurethane/cellulose nanocrystal bionanocomposites.

499

Composites Science and Technology, 88, 39-47.

500

Segal, L., Creely, J. J., Martin, A. E., Jr., & Conrad, C. M. (1959). An empirical method

501

for estimating the degree of crystallinity of native cellulose using the X-ray

502

diffractometer. Textile Research Journal, 29(10), 786–794.

503

Viguié, J., Molina-Boisseau, S., & Dufresne, A. (2007). Processing and characterization

504

of waxy maize starch films plasticized by sorbitol and reinforced with starch

505

nanocrystals. Macromolecular Bioscience, 7(11), 1206-1216.

506

Wang, J.-L., Cheng, F., Zhu, P-X. (2014). Structure and properties of urea-plasticized

507

starch films with different urea contents. Carbohydrate Polymers, 101, 1109-1115.

508

Wang, Y. X., Tian, H., & Zhang, L. (2010). Role of starch nanocrystals and cellulose

509

whiskers in synergistic reinforcement of waterborne polyurethane. Carbohydrate

510

Polymers, 80, 665–671.

511

Xie, F., Pollet, E., Halley, P. J., Avérous, L. (2013) Starch-based nano-biocomposites.

512

Progress in Polymer Science, 38, 1590– 1628.

513

Yan, Q., Hou, H., Guo, P., & Dong H (2012). Effects of extrusion and glycerol content

514

on properties of oxidized and acetylated corn starch-based films. Carbohydrate

515

Polymers, 87, 707– 712.

Ac ce p

te

d

M

an

us

cr

ip t

495

516

-24Page 24 of 33

516

WSNC (wt.%) 0

CNC (wt.%) 0

CSg35+1 (100/0)

1

100

0

CSg35+1 (50/50)

1

50

50

CSg35+1 (0/100)

1

0

100

2.5

100

0

5

100

CSg35+2.5 CSg35+5

0

Table 1. Compositions of obtained bionanocomposites.

us

517

cr

CSg35

ip t

Nanocrystal (wt.%) 0

Ac ce p

te

d

M

an

518

-25Page 25 of 33

Water Vapour Permeance (kg/m2 s Pa)1010 56.2  2.0

Oxygen Permeance (cm3/m2 day atm) 22.7  3.9

48.8  24.4

1.7  0.4

66.7  4.5

26.9  1.6

7.4  1.1

41.8  8.7

1.6  0.3

57.3  3.8

7.5  1.1

47.9  10.3

1.6  0.1

57.3  1.4

7.7  1.5

Strain at break (%)

Tensile strength (MPa)

CSg35 CSg35+1 (100/0) CSg35+2.5

2.5  0.9

52.7  9.4

5.8  0.9

CSg35+5

15.5  4.7

cr

Young’s Modulus (MPa)

Table 2. Tensile properties and water vapour and oxygen permeance values for plasticized starch and TPS-WSNC nanocomposites.

us

519 520

1.0  0.2

ip t

518

521

Ac ce p

te

d

M

an

522

-26Page 26 of 33

522

523 524

48.8  24.4

1.7  0.4

66.7  4.5

26.9  1.6

12.0  2.0

41.7  4.2

1.6  0.1

54.2  1.1

8.5  2.2

41.9  8.5

1.5  0.2

57.9  2.7

2.5  0.9

52.7  9.4

5.8  0.9

26.6  1.9

cr

Tensile Strength (MPa)

ip t

1.0  0.2

Oxygen Permeance (cm3/m2 day atm) 22.7  3.9

Strain at break (%)

22.1  1.1

us

CSg35 CSg35+1 (100/0) CSg35+1 (50/50) CSg35+1 (0/100)

Water Vapour Permeance (kg/m2 s Pa)1010 56.2  2.0

Young’s Modulus (MPa)

Table 3. Tensile properties and water vapour and oxygen permeance values for TPS and TPS+WSNC/CNC.

Ac ce p

te

d

M

an

525

-27Page 27 of 33

Highlights (for review)

STARCH AND CELLULOSE NANOCRYSTALS TOGETHER INTO THERMOPLASTIC STARCH BIONANOCOMPOSITES K. González et al.

ip t

HIGHLIGHTS Bioanocomposites were reinforced with combined polysaccharide nanocrystals.



Increasing waxy starch nanocrystals content mechanical properties were enhanced.



An important improvement of oxygen permeance was obtained at certain

us

cr



percentages.

Barrier to water vapour was maintained almost insensitive to nanoreinforcement.



Mechanical properties were improved by the incorporation of cellulose

an



Ac

ce pt

ed

M

nanocrystals.

Page 28 of 33

Ac

ce

pt

ed

M

an

us

cr

i

Figure 1

Page 29 of 33

Ac

ce

pt

ed

M

an

us

cr

i

Figure 2

Page 30 of 33

Ac

ce

pt

ed

M

an

us

cr

i

Figure 3

Page 31 of 33

Ac

ce

pt

ed

M

an

us

cr

i

Figure 4

Page 32 of 33

Ac ce p

te

d

M

an

us

cr

ip t

Figure 5

Page 33 of 33

Starch and cellulose nanocrystals together into thermoplastic starch bionanocomposites.

In the present work, thermoplastic maize starch based bionanocomposites were prepared as transparent films, plasticized with 35% of glycerol and reinf...
755KB Sizes 0 Downloads 7 Views