Effect of different polyol-based plasticizers on thermal properties of polyvinyl alcohol:starch blends

Effect of different polyol-based plasticizers on thermal properties of polyvinyl alcohol:starch blends

Accepted Manuscript Title: Effect of Different Polyol-Based Plasticizers on Thermal Properties of Polyvinyl Alcohol (PVA):Starch Blends Films Author: ...

829KB Sizes 2 Downloads 94 Views

Accepted Manuscript Title: Effect of Different Polyol-Based Plasticizers on Thermal Properties of Polyvinyl Alcohol (PVA):Starch Blends Films Author: Ahmet Alper Aydın Vladimir Ilberg PII: DOI: Reference:

S0144-8617(15)00841-3 http://dx.doi.org/doi:10.1016/j.carbpol.2015.08.093 CARP 10296

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

20-5-2015 12-8-2015 29-8-2015

Please cite this article as: Aydin, A. A., and Ilberg, V.,Effect of Different Polyol-Based Plasticizers on Thermal Properties of Polyvinyl Alcohol (PVA):Starch Blends Films, Carbohydrate Polymers (2015), http://dx.doi.org/10.1016/j.carbpol.2015.08.093 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

Effect of Different Polyol-Based Plasticizers on Thermal Properties of Polyvinyl Alcohol

2

(PVA):Starch Blends Films

3

Ahmet Alper Aydın1,* and Vladimir Ilberg2

4

1

5

Istanbul Technical University, 34469 Maslak, Istanbul, Turkey. Email: [email protected]

6

2

7

Triesdorf, Am Staudengarten 11, D-85350, Freising, Germany.

8

Abstract

9

A series of gelatinized polyvinyl alcohol (PVA):starch blends were prepared with various polyol-

10

based plasticizers in 5 wt%, 15 wt% and 25 wt% ratios via solution casting method. The obtained

11

films were analyzed by Fourier transform infrared (FT-IR) spectroscopy, differential scanning

12

calorimetry (DSC) and thermogravimetric analysis (TGA). Remarkable changes have been

13

observed in glass-transition temperature (Tg) and thermal stability of the samples containing

14

varying concentrations of different plasticizers and they have been discussed in detail with

15

respect to the conducted thermal and chemical analyses. The observed order of Tg point

16

depression of the samples with containing 15 wt% plasticizer content is 1,4-butanediol - 1,2,6-

17

hexanetriol - pentaerythriyol - xylitol - mannitol, which is similar to the sequence of the thermal

18

stability changes of the samples. In the presence of 25 wt % 1,4-butanediol, the Tg point of

19

PVA:starch films reduce from 76.1oC to 37.2oC.

20

Keywords: starch; polyvinyl alcohol; plasticizer; polyol; thermal properties; glass transition

21

Chemical compounds studied in this article: Corn Starch (PubChem CID: 24836924);

22

Polyvinyl Alcohol (PubChem CID: 11119); 1,4-Butanediol (PubChem CID: 8064); 1,2,6-

23

Hexanetriol (PubChem CID: 7823); Pentaerythritol (PubChem CID: 8285); Xylitol (PubChem

24

CID: 6912); Mannitol (PubChem CID: 6251)

ip t

Chemical Engineering Department, Faculty of Chemical and Metallurgical Engineering,

Ac ce p

te

d

M

an

us

cr

Faculty of Gardening and Food Technology, University of Applied Sciences Weihenstephan-

1 Page 1 of 24

1. Introduction

26

Petroleum-based synthetic polymers with high molecular weight and hydrophobic character have

27

been extensively used in different areas. However, their extensive use and ecological concerns

28

have increased interest in biodegradable alternatives from renewable sources (Petersen et al.,

29

1999; Weber et al., 2002).However, these polymers show very high chemical stability and

30

degrade very slowly in the environment. In the 21st century, environmental and health issues

31

associated with disposal and recycling of these products indicate major problems and extensive

32

research activities continue for developing their biodegradable alternatives.

33

There have been intensive attempts to develop biodegradable plastics with the purpose of

34

protecting the environment. Biodegradable polymers offer environmentally friendly substitutes

35

which can be degraded by microorganisms in soil or water. The approaches toward development

36

of biodegradable polymers contain modification of non-degradable polymers by introducing

37

degradable linkages or blending biodegradable polymers and tailoring their formulations so that

38

the resulting materials have properties superior to the individual components (Tang, & Alavi,

39

2011).

40

Starch has been considered as one of the most suitable materials among natural biopolymers due

41

to its low cost and abundant availability (Wang, Chang, &Zhang, 2010; Gross & Kalra, 2002).

42

Polysaccharides such as starch, chitosan and cellulose are typical examples of natural

43

biodegradable polymers with relatively good biocompatibility. However, starch-based products

44

exhibit several disadvantages such as brittleness, strong hydrophilic character, poor mechanical

45

properties and processability which limit their applications in material engineering. Therefore, the

46

weakness of starch products must be compensated by blending starch with other synthesized

47

degradable polymers (Shi et al., 2008). In this sense, blends with PVA are well suited to improve

48

the strength and flexibility of starch-based products as a result of chemical resistance, optical and

Ac ce p

te

d

M

an

us

cr

ip t

25

2 Page 2 of 24

physical properties, good film-forming capability, water solubility and biocompatibility of PVA

50

(Nakayama, Takatsuka, & Matsuda, 1999; Cinelli, Chiellini, Gordon, & Chiellini, 2005). the

51

making biodegradable blends with natural polymers due to its good physical properties, film-

52

forming capability and water solubility

53

Although PVA is a suitable solution to improve the properties of starch-based products, further

54

improvement is still needed. As a result of the formed hydrogen bonds between PVA and starch

55

molecules, the molecular movement and processability of the blend are restrained. At this point,

56

plasticizers play an important role to reduce the strong interactions between PVA and starch

57

molecules by forming new hydrogen bonds with PVA and starch. In this way, PVA:starch blends

58

are processed into mouldable thermoplastic materials in the presence of suitable plasticizers

59

(Chin & Te, 2008; Forssell et al., 1997; Raj, Siddaramaiah & Somashekar, 2004). Besides, the

60

polarity of PVA increases the hydrolytic attack to the blend by atmospheric moisture and

61

accelerates the break down in sugar molecules (Raj, Siddaramaiah, & Somashekar, 2004). Since

62

starch and PVA molecules contain many hydroxyl groups, their blend films also have hydrophilic

63

nature. Therefore, their thermal and mechanical properties and water resistance must be improved

64

and tailored for different applications.

65

Although PVA:starch blend films have been produced by means of solution casting, it is not an

66

economic and efficient way compared to thermoplastic processing in larger production scale.

67

Glass-transition temperature (Tg) plays an important role in production of thermoplastic products

68

via extrusion. At the Tg point, the cohesive forces drastically decrease and the polymer expands.

69

The increase in free volume reaches an extent that there is room for migration of segments, which

70

provides flow of the polymer. Addition of plasticizers reduces molecular interactions, lowers the

71

Tg point and makes the polymer more rubber-like.

Ac ce p

te

d

M

an

us

cr

ip t

49

3 Page 3 of 24

The degree of conversion of PVA:starch blend into mouldable thermoplastic is a measure of Tg

73

point depression of the sample. The cohesive forces drastically decrease at the Tg point and the

74

polymer expands to create room for migration of segments, which provides flow of the polymer.

75

In the case of PVA and starch, strong molecular interactions are disturbed in the presence of

76

plasticizers and the Tg point is reduced which makes the polymer more rubber-like (Jayasekara et

77

al., 2003; Yoon, Chough, & Park, 2006a; Sreedhar et al., 2006).

78

Different plasticizers like glycerol (Sreekumar, Al-Harthi, & De, 2012; Luo, Li, & Lin, 2012),

79

polyethylene glycol (Sreedhar et al., 2005), sorbitol (Arvanitoyannis, Kolokuris, Nakayama,

80

Yamamoto, & Aiba, 1997), sucrose (Arvanitoyannis, Kolokuris, Nakayama, Yamamoto, & Aiba,

81

1997), urea (Luo, Li, & Lin, 2012), ascorbic acid (Yoon, 2014), citric acid (Shi et al., 2008;

82

Yoon, Chough, & Park, 2006a), succinic acid (Yoon, Chough, & Park, 2006b), malic acid (Yoon,

83

Chough, & Park, 2006b), tartaric acid (Yoon, Chough, & Park, 2006b) or more complex

84

plasticizers (Zhou, Cu, Jia, & Xie, 2009) have been successfully employed to improve flexibility

85

of the blends.

86

Liu et al. (1999) compounded glycerol and water with PVA:starch blend in a single screw

87

extruder. They reported that glycerol was much more effective than water as a plasticizer. In

88

addition, Park et al. (2005) examined glycerol, sorbitol and citric acid as additives with different

89

functional groups. They concluded that due to the increased hydrogen bonding in the presence of

90

both hydroxyl group and carboxyl group, citric acid provided better film forming compared to

91

sorbitol and glycerol. Similarly, Shi et al. (2008) reported another role of citric acid in the blend.

92

According to their reported data, partial esterification takes place between starch (or PVA) and

93

citric acid during processing at 140oC in the extruder. While partial esterification initially

94

increases the tensile strength and Tg of the blend in the presence of 5% citric acid, further

95

increase in residual citric acid amount provides plasticizing effect.

Ac ce p

te

d

M

an

us

cr

ip t

72

4 Page 4 of 24

Other effective ways of improving the properties of PVA:starch blends include: (i) chemically

97

modifying starch or PVA (Kim, & Lee, 2002; Kim, 2003), (ii) physically modifying by forming

98

PVA/starch based nanocomposites (Dean, Do, Petinakis, Yu, 2008; Majdzadeh-Ardakani, &

99

Nazari, 2010; Vasile et al., 2008; Tang, Alavi, & Herald, 2008), (iii) chemically modifying the

100

PVA and starch during or after blending process with crosslinking agents such as boric acid (Yin,

101

Li, Liu, & Li, 2005), borax (Sreedhar et al., 2005), epichlorohydrin (Sreedhar et al., 2006) and

102

hexamethoxymethylmelamine (Chen et al., 1997).

103

The aim of this paper is to introduce the influence of different polyol-based plasticizers on

104

plasticization of neat PVA:starch blend with respect to the Tg point depression and thermal

105

stability changes in the presence of plasticizers. For this purpose, 1,4-butanediol, 1,2,6-

106

hexanetriol, pentaerythritol, xylitol and D-mannitol have been investigated as polyol-based new

107

plasticizers and the changes in Tg point and thermal stability of the blends containing varying

108

amounts of plasticizers have been discussed in detail with respect to the conducted thermal and

109

chemical analyses.

110

2. Materials and Methods

111

2.1 Materials

112

Corn starch was supplied by Sigma-Aldrich and it was composed of 73% amylopectin and 27%

113

amylose. PVA was also obtained from Sigma-Aldrich, which was 99% hydrolyzed with an

114

average molecular weight of 130.000 amu.

115

The conducted strategy in choosing plasticizer was based on obtaining a series compounds with

116

varying hydroxyl group functionality on the carbon chain. In this manner, the series of 1,4-

117

butanediol (99%, Sigma), 1,2,6-hexanetriol (96%, Sigma), pentaerythritol (>99%, Sigma), xylitol

118

(>99%, Sigma) and D-mannitol (>98%, Sigma), which contains polyols with different number of

119

hydroxyl groups as shown in Fig. 1, was used as received from the supplier.

Ac ce p

te

d

M

an

us

cr

ip t

96

5 Page 5 of 24

ip t cr us

120

Fig. 1. Chemical structures of the investigated plasticizers

122

2.2 Preparation of PVA:Starch blend films

123

PVA:starch (1:1, w/w) blend film samples were prepared by solution casting method according to

124

the mass ratios given in Table 1. First, PVA was dissolved in 300 mL hot water at 97±2oC. After

125

complete dissolution of PVA, calculated amount of plasticizer was added and stirred for 15

126

minutes in order to maintain homogeneous distribution of plasticizer in hot water. Afterwards,

127

dispersed starch in 100 mL water was added and the mixture was stirred for another 45 minutes.

128

The mixtures were then casted onto petri dishes with similar weight (30 ± 2 g) and dried at 65oC

129

to constant weight. In order to comparatively investigate the blends, Samples containing neat

130

corn starch and PVA:starch were also prepared as reference.

131

The peeled film samples were directly taken into sealed plastic bags and stored at room

132

temperature in desiccator containing fresh silica-gel beads to maintain constant low-humidity.

133

Samples exceeding 2 days were withdrawn and prepared again to guarantee low humidity of the

134

samples.

Ac ce p

te

d

M

an

121

135 136 137

6 Page 6 of 24

138

Table 1. Abbreviations and corresponding sample compositions

PSG-M5 PSG-M15 PSG-M25 Abbreviation PSG-P5 PSG-P15

PSG-H5 PSG-H15 PSG-H25

(wt% of the dry weight)

5 15 25 Pentaerythritol

Abbreviation PSG-X5 PSG-X15 PSG-X25

(wt% of the dry weight)

5 15

1,2,6-Hexanetriol (wt% of the dry weight)

5 15 25 Xylitol

ip t

Abbreviation

5 15 25 Mannitol

Abbreviation

(wt% of the dry weight)

cr

PSG-B5 PSG-B15 PSG-B25

1,4-Butanediol (wt% of the dry weight)

us

Abbreviation

5 15 25

an

139

2.3 Fourier transform infrared spectroscopy (FT-IR)

141

FT-IR spectra of the synthesized high-chain fatty acid esters were recorded on a Perkin Elmer

142

FT-IR Spectrum 100 spectrometer with universal ATR accessory between 4000 and 650 cm-1

143

wavelength.

144

2.4 Differential scanning calorimeter (DSC)

145

A TA Instruments Q200 DSC was used for the calorimeter analyses of the samples. The

146

measurements were carried out under inert nitrogen atmosphere at 50 ml/min flow rate. All the

147

DSC thermal analyses were conducted at 5oC/min rate for the determination of glass transition

148

points.

149

DSC analyses were conducted according to the ASTM standard test methods with designation

150

numbers E 793-06 and E 1269-11, explaining the determination of enthalpies of fusion and

151

freezing and specific heat of liquids and solids, respectively. The temperature and heat

152

calibrations of the instrument were systematically performed with sapphire and indium references

153

prior to the analyses on each workday.

Ac ce p

te

d

M

140

7 Page 7 of 24

Every presented DSC data in this paper is calculated according to the results of at least 3

155

individual analyses in order to minimize the uncertainty.

156

2.5 Thermo-gravimetric analyses (TGA)

157

A NETZSCH STA 409 PC/PG was used for determination of thermo-gravimetric decomposition

158

of the film samples, including decomposition behavior, onset temperature and weight losses at

159

different temperatures of the samples. The analyses were carried out under inert argon

160

atmosphere at 60 ml/min flow and 10oC/min heating rate between 30 and 600oC.

161

The analyses were conducted according to the general principles given in BS EN ISO

162

11358:1997. The weight and temperature calibrations of the instrument were made using the

163

reference weight and according to the sensor calibration of the instrument, respectively. The

164

calibration of the instrument was performed systematically prior to the first analysis of each

165

workday.

166

Every presented TGA data in this paper is calculated according to the results of at least 2

167

individual analyses.

168

3. Results and Discussion

169

3.1 Plasticizers and hydrogen bonding (FT-IR)

170

The hydrogen bonding plays an important role in formation of strong interactions between starch,

171

PVA and plasticizer molecules and the FT-IR spectra enable the interactions to be identified. The

172

changes in peak bandwidth, strength and frequency provide valuable data for interpretations on

173

the formation of intermolecular hydrogen bonding in the samples.

174

In general, the changes in band formations are clearly seen in the overlapped spectra of the blends

175

given in Fig. 2(ii) compared to the polyol-based plasticizers in Fig. 2(i). In Fig. 2(ii)-a, the broad

176

transmittance band of starch at 3293 cm-1 is assigned to the stretching vibration of –OH and the

177

band at 2922 cm-1 is due to C–H stretching in the molecule. The band peak at 1638 cm-1 indicates

Ac ce p

te

d

M

an

us

cr

ip t

154

8 Page 8 of 24

the bound water in starch with the formed hydrogen bonds. The bands located at 1411 cm-1 and

179

858 cm-1 are assigned to the vibrations associated with the CH2 group. The peak at 1011 cm-1 is

180

related to C–O bond stretching of C–O–C groups in anhydroglucose ring. The C–O bond

181

stretching of C–O–H group is seen at 1150 cm-1. The given data are in good agreement with the

182

literature (Fang et al., 2002; Pavlovic & Brandao, 2003; Ma & Yu, 2004).

Ac ce p

te

d

M

an

us

cr

ip t

178

183 184 185 186

Fig. 2. FTIR spectra: i – (a) 1,4-Butanediol, (b) 1,2,6-Hexanetriol, (c) Pentaerythritol, (d) Xylitol, (e) Mannitol; ii – (a) Starch, (b) PVA:Starch, (c) PSG-B15, (d) PSG-H15, (e) PSG-P15, (f) PSG-X15, (g) PSG-M15

187 9 Page 9 of 24

The hydrogen bonding between –OH groups of starch, PVA and polyol-based plasticizers in the

189

blends shifts the –OH stretching vibration bands of all neat polyol-based plasticizers to lower

190

frequency region by around 4-14 cm-1 and the corresponding bands of blends get broader and

191

stronger. Additionally, the positions of the C–O stretching bands of the C–O–H group located at

192

around 1150 cm-1 significantly shift to lower frequencies as a result of the changed hydrogen

193

bonding ratio in the blends. The single C–O stretching band of the C–O–C group in starch and

194

PVA:starch samples transforms into double peak with the added plasticizers. The bound H2O

195

shifts from 1638 cm-1 to higher frequencies by the addition of PVA and polyol-based plasticizers,

196

indicating that the water molecules are more strongly hydrogen bonded in the presence of PVA

197

and plasticizers (Wolkers et al., 2004). The details of the FT-IR spectra given in Fig. 2(ii) are

198

tabulated in Table 2.

199

Table 2 indicates that comparative changes are observed between spectra of polyol-based

200

plasticizer containing samples and neat starch and PVA:starch samples, whereas spectra of

201

plasticizer containing samples have generally close band frequencies as a result of their similar

202

chemical structures. In PSG-X15 and PSG-M15 samples, C-H stretching is cleaved into two

203

bands at around 2937-2932 and 2908-2910 cm-1. Besides, C-O stretching of C-O-C is located at

204

higher frequencies compared to other polyol-based plasticizer containing samples.

206

cr

us

an

M

d

te

Ac ce p

205

ip t

188

207 208 209 210 211 10 Page 10 of 24

Table 2. Summary of FT-IR results PSGB15 (cm-1)

PSGH15 (cm-1)

PSGP15 (cm-1)

PSGX15 (cm-1)

PSGM15 (cm-1)

3293

3296

3292

3295

3304

3270

3287

2922

2931

2932

2933

2932

1638

1654

1654

1653

1653

1411, 858

1424, 844

1417, 845

1417, 847

1011

988

1014, 996

1150

1149

1147

ip t

PVA:Starch (cm-1)

2932, 2910

1654

1654

1409, 848

1417, 842

1417, 843

1016, 996

1009, 991

1087, 1044

1077, 1024

1145

1146

1142

1143

cr

2937, 2908

us

OH stretching C-H streching Bound water Vibrations associated to CH2 group C-O stretching of C-O-C C-O stretching of C-O-H

Starch (cm-1)

an

Functional group

M

212

3.2 The effects of plasticizers on thermal properties

214

3.2.1 DSC analyses and changes in Tg points

215

As it is given in Table 3 and Fig. 3-i, the casted sample of pure PVA has a clear Tg at 70.5oC.

216

However, such a clear Tg cannot be observed for gelatinized starch, which could be because of its

217

amorphous and hygroscopic nature. It has been stated in literature that the absence of its Tg may

218

be due to amorphous chains surrounded by crystalline domains, presence of moisture, physical

219

crosslinks inhibiting the movements of the amorphous chain segments or presence of

220

intercrystalline phases (Sreedhar et al., 2005; Shi et al., 2008). Although there are no discernible

221

changes in the DSC thermogram of gelatinized starch film, the PVA:starch (1:1, w/w) blend has a

222

clear Tg at 76.1oC, which is higher than that of neat PVA as a result of the hydrogen bonding

223

interactions between starch and PVA.

Ac ce p

te

d

213

224 225 11 Page 11 of 24

Table 3. Glass transition points (Tg) of the samples with 95% confidence interval Abbreviation PVA:Starch PSG-H5 PSG-H15 PSG-H25 PSG-M5 PSG-M15 PSG-M25

Tg (oC) ± 95% Conf. Int. 76.1 ± 1.4 49.9 ± 0.6 45.6 ± 1.2 43.0 ± 2.3 52.5 ± 1.3 52.2 ± 0.4 51.6 ± 0.4

ip t

Tg (oC) ± 95% Conf. Int. 70.5 ± 1.2 49.2 ± 1.7 45.2 ± 0.5 37.2 ± 2.2 51.5 ± 0.1 51.4 ± 1.8 46.6 ± 0.4 54.1 ± 0.2 48.7 ± 1.8

us

Abbreviation PVA PSG-B5 PSG-B15 PSG-B25 PSG-X5 PSG-X15 PSG-X25 PSG-P5 PSG-P15

cr

226

Plasticizers are low molecular weight substances, which reduce the Tg point of polymers and

228

provide their flow by improving flexibility and processability. In this sense, the effects of

229

plasticizers in different polymer blend compositions can be observed by investigating the degree

230

of changes in Tg points at different concentrations (Jayasekara et al., 2003; Yoon, Chough, &

231

Park, 2006a; Sreedhar et al., 2006).

232

In the case of PVA:starch films, two types of hydrogen bonding interactions take place: (1)

233

hydrogen bonding between hydroxyl groups of PVA and gelatinized starch and (2) hydrogen

234

bonding between functional groups of PVA, starch and plasticizer. In the second type,

235

plasticizers function as agents penetrating the matrix and interrupting the hydrogen bonding

236

sequence between PVA and starch. They form hydrogen bonding bridges via functional groups in

237

their chemical structure and spread throughout the matrix with their small molecular size and low

238

molecular weight. In this way, polar attractive forces are established between the plasticizer and

239

chain segments, which are responsible for the reduction of Tg point and simultaneously,

240

enhancement of chain segment mobility with better flexibility and processibility of the matrix

241

(Chin & Te, 2008; Forssell et al., 1997; Raj, Siddaramaiah, & Somashekar, 2004).

242

The hydroxyl groups are hydrogen bonding quarters on the investigated plasticizers (Fig. 1)

243

which might possibly form new hydrogen bonding bridges and enable better segment mobility in

Ac ce p

te

d

M

an

227

12 Page 12 of 24

the PVA:starch matrix. In this way, 5 wt% addition of these plasticizers into PVA:starch blend

245

significantly drops the Tg point in Table 3. The Tg points shift significantly to lower temperature

246

region in DSC data presented in Fig. 3.

247

Ac ce p

te

d

M

an

us

cr

ip t

244

248 249 250

Fig. 3. DSC data: i – (a) Starch, (b) PVOH, (c) Starch:PVOH; ii – (a) PSG-B5, (b) PSG-H5, (c) PSG-P5, (d) PSG-X5, (e) PSG-M5; iii – (a) PSG-B15, (b) PSG-H15, (c) PSG-P15, (d) PSG-X15, (e) PSG-M15; iv – (a) PSG-B25, (b) PSG-H25, (c) PSG-X25, (d) PSG-M25

251

Unlike xylitol and mannitol, the Tg points continue decreasing in the presence of 15 wt% and 25

252

wt% 1,4-butanediol, 1,2,6-hexanetriol and pentaerythritol as a result of increasing number of

253

available hydroxyl groups. However, Tg points do not significantly decrease in the samples

254

containing 15 wt% xylitol and mannitol. Only for xylitol, increase in its mass ratio to 25 wt%

255

leads to additional increase in segment mobility and decrease in Tg point down to 46.6oC.

13 Page 13 of 24

The observed different plasticizing efficiencies of xylitol and mannitol can be attributed to their

257

molecular structures. Compared to 1,4-butanediol, 1,2,6-hexanetriol and pentaerythritol, they are

258

relatively larger molecules with 5 and 6 hydroxyl groups, respectively. Probably, further

259

penetration into chain segments of PVA and starch is prevented due to their larger molecular

260

geometry after the initial plasticization of 5 wt% addition. Consequently, it can be mentioned that

261

mannitol molecules tend to interact with each other and do not significantly improve the thermal

262

properties of the blend, whereas smaller xylitol molecules can further play role in plasticization

263

when they are found in high amount.

264

3.2.2 Thermal-decomposition behaviors and thermogravimetric analyses

265

In the given figures, three distinct regions of mass loss can be seen. There are generally three

266

distinct mass loss regions in PVA:starch samples reported in literature. The first region (between

267

75oC and 200oC) is attributed to the loss of loosely bound water, accompanied by the formation

268

of volatile disintegrated products such as dislocation of the plasticizers in the blend. The second

269

region is described as the main decomposition step and the final stage is the carbonization of the

270

organic matter (above 500oC) (Galdeno et al., 2009; Luo, X., Li, J., & Lin, X, 2012; Saiah et al.,

271

2009; Shi et al., 2011). The thermal decomposition behaviors of various PVA:starch films

272

with/without polyol-based plasticizers are given in Fig. 4.

273

In addition to the overlapped thermal decomposition curves in Fig. 4, the thermal decomposition

274

data are also tabulated in Table 4 to present the 10 % weight loss temperature and residual

275

weights at 300oC, 400oC and 500oC for reference samples (starch, PVA, PVA:starch) and 15 wt%

276

polyol-based plasticizer containing samples.

Ac ce p

te

d

M

an

us

cr

ip t

256

277 278 279 14 Page 14 of 24

280

Table 4. Thermal decomposition values of the samples with 15 wt% plasticizer Residual weight percentage at 400oC

at 500oC

88.84 52.12 80.32 72.22 68.11 74.55 74.87 89.47

21.82 27.99 33.16 28.17 24.69 26.72 23.88 30.85

16.61 10.54 20.85 14.09 11.79 11.37 11.52 14.27

cr

ip t

at 300 C

Ac ce p

te

d

M

an

Starch PVA PVA:Starch PSG-B15 PSG-H15 PSG-P15 PSG-X15 PSG-M15

o

us

Abbreviation

Temperature at 10 % weight loss 298.4oC 258.8oC 285.6oC 244.9oC 238.3oC 269.4oC 275.2oC 298.7oC

281 282 283

Fig. 4. TGA data: i – (a) Starch, (b) PVOH and (c) Starch:PVOH; ii – (a) PSG-B15, (b) PSGH15, (c) PSG-M15, (d) PSG-P15, (e) PSG-X15 15 Page 15 of 24

The PVA:starch films show better thermal stability than neat PVA as a result of the thermal

285

resistive cyclic hemiacetal in starch structure and development of blending towards high-energy

286

stability of the mixture (Sin et al., 2011). Therefore, PVA:starch sample has closer 10 %

287

decomposition temperature to gelatinized neat starch instead of PVA and less weight loss than

288

neat PVA at 300oC, 400oC and 500oC compared to PVA.

289

When the thermal decomposition data of polyol-based plasticizer containing samples are

290

examined, reductions in thermal resistivity are generally observed, except for mannitol containing

291

sample, compared to the PVA:starch sample. Although the mass loss order of different

292

plasticizers is attributed to their volatility difference in literature (Ma, Yu, & Wan, 2006), the

293

mass loss order of samples containing different plasticizers is also related to intermolecular

294

interactions and mobility between the plasticizer, PVA and starch chains according to the

295

interpreted data below.

296

As PVA:starch sample has closer thermal stability to starch due to the development of blending

297

towards high-energy stability, it is expected that higher the plasticization effect, lower the thermal

298

stability of the blend sample.

299

Other than the exception of 1,4-butanediol and 1,2,6-hexanetriol containing samples, the

300

increasing hydroxyl number and molecular size of the plasticizer enhance the thermal stability

301

with a maximum in the presence of mannitol. Although the increase in molecular size of the

302

plasticizers decreases their penetration ability into PVA and starch chain segments and limit the

303

Tg depletion (plasticization effect), it provides convergence of the thermal decomposition

Ac ce p

te

d

M

an

us

cr

ip t

284

16 Page 16 of 24

resistance of the blend samples to neat PVA:starch film as a result of the developed blending with

305

high-energy stability. The exceptional behavior of 1,4-butanediol and 1,2,6-hexanetriol

306

containing samples might be attributed to their volatility difference. Therefore, the thermal

307

stability of PVA:starch:plasticizer samples is related to not only the volatility difference of the

308

plasticizers, but also to the mobility between the plasticizer, PVA and starch chains.

309

The mannitol containing sample shows better thermal endurance than neat starch and PVA:starch

310

samples at 300 oC and its 10 % weight loss temperature is at 298.7 oC, which is the highest

311

among the analyzed samples. However, its decomposition rate later gets higher and at 400oC and

312

500oC, the neat PVA:starch sample is distinguished as the most durable sample with 33.16 % and

313

20.85 % of remaining weight, respectively.

314

4. Conclusion

315

As a result of the increasing reservations in environmental and health issues, development of new

316

biodegradable products attracts attention of researchers. However, biodegradable polymers

317

cannot be solely used and must be blended with other biodegradable additives to enhance their

318

mechanical properties and shape stability. Although the properties of PVA:starch blend films

319

containing different plasticizers have been investigated in literature, new plasticizers are still

320

needed for better thermoplastic processability of the blends in production scale. In this sense,

321

several polyol-based plasticizers have been investigated in different mass ratios for gelatinized

322

PVA:starch blend films.

323

5 wt% addition of the investigated plasticizers lowers the Tg point significantly, and different

324

plasticizing behaviors are observed with increasing plasticizer amount. The observed decrease

325

continue in a descending manner with 15 wt% 1,4-butanediol, 1,2,6-hexanetriol and

326

pentaerythritol addition. Xylitol leads to additional increase in segment mobility with 25 wt%

327

addition and mannitol stays almost the same despite its increasing amount in the blend. The

Ac ce p

te

d

M

an

us

cr

ip t

304

17 Page 17 of 24

difference of xylitol and mannitol is attributed to lower penetration capability of the molecules

329

into chain segments after the initial plasticization of 5 wt% addition.

330

The thermal decomposition data of polyol-based plasticizer containing samples indicate that

331

thermal resistivity is generally lower than neat PVA:starch blend film samples. However, the

332

orders of the plasticizers according to the 10 % weight loss temperature and residual weight

333

percentages at 300 oC generally indicate that the increasing hydroxyl number and molecular size

334

of the plasticizer enhance the thermal stability with a maximum in the presence of mannitol.

335

Based on the reported data, it can be consequently stated that although the number of hydrogen

336

groups on plasticizers are hydrogen bonding quarters for starch and PVA, molecular structure and

337

molecular geometry of plasticizers can prevent their penetration into chain segments and reduce

338

the intermolecular interactions, which in turn limit the expected plasticizing effect. Among the

339

investigated polyol-based plasticizers, 1,4-butanediol shows the highest plasticizing effect for

340

PVA:starch. It reduces the Tg point by approximately 27 oC, 31 oC and 39 oC, and 1,2,6-

341

hexanetriol reduces the Tg point by approximately 26 oC, 30 oC and 33 oC with 5 wt%, 15 wt%

342

and 25 wt% addition in the samples, respectively.

343

References

344

Arvanitoyannis, I., Kolokuris, I., Nakayama, A., Yamamoto, N., & Aiba, S. (1997). Physico-

345

chemical studies of chitosan-poly (vinyl alcohol) blends plasticized with sorbitol and sucrose.

346

Carbohydrate Polymers, 34, 9-19.

347

Chen, L., Iman, S. H., Stein, T. M., Gordon, S. H., Hou, C. T., & Greene, R. V. (1997). Starch-

348

polyvinyl alcohol cast film-performance and biodegredation. Polymer Preprints, 37, 461-462.

349

Chin, A. L., & Te, H. K. (2008). Shear and elongational flow properties of thermoplastic

350

polyvinyl alcohol melts with different plasticizer contents and degrees of polymerization. Journal

351

of Materials Processing Technology, 200, 331-338.

Ac ce p

te

d

M

an

us

cr

ip t

328

18 Page 18 of 24

Cinelli, P., Chiellini, E., Gordon, L. S. H., & Chiellini, L. E. (2005). Characterization of

353

biodegredable composite films prepared from blends of poly(vinyl alcohol), cornstarch, and

354

lignocellulosic fibre. Journal of Polymer Environment, 13, 47-59.

355

Dean, K. M., DO, M. D., Petinakis, E., & Yu, L. (2008). Key interactions in biodegradable

356

thermoplastic

357

Composites Science and Technology, 68(6), 1453-1462.

358

Fang, J. M., Fowler, P. A., Tomkinson, J., & Hill, C. A. S. (2002). The preparation and

359

characterization of a series of chemically modified potato starch. Carbohydrate Polymers, 47,

360

245-252.

361

Forssell, P. M., Mikkila, J. M., Moates, G. K., & Parker, R. (1997). Phase and glass transition

362

behaviour of concentrated barley starch-glycerol-water mixtures, a model for thermoplastic

363

starch. Carbohydrate Polymers, 34, 275-282.

364

Galdeano, M. C., Grossmann, M. V. E., Mali, S., Bello-Perez, L. A., Garcia, M. A., & Zamudio-

365

Flores, P. B. (2009). Effects of production process and plasticizers on stability of films and sheets

366

of oat starch. Materials Science and Engineering: C, 29, 492–498.

367

Gross, R.A., & Kalra, B. (2002). Biodegradable polymers for the environment. Science, 297, 803-

368

807.

369

Jayasekara, R., Harding, I., Bowater, I., Christie, G. B. Y., Lonergan, G. T. J. (2003).

370

Biodegradation by composting of surface modified starch and PVA blended films. Journal of

371

Polymers and the Environment, 11, 49-56.

372

Kim, M. (2003). Evaluation of degradability of hydroxypropylated potato starch/polyethylene

373

blend films. Carbohydrate Polymers, 54, 173-181.

374

Kim, M., & Lee, S-J. (2002). Characteristics of crosslinked potato starch and starch-filled linear

375

low-density polyethylene films. Carbohydrate Polymers, 50, 331-337.

micro-

and

nanocomposites.

cr

alcohol)/montmorillonite

Ac ce p

te

d

M

an

us

starch/poly(vinyl

ip t

352

19 Page 19 of 24

Liu, Z. Q., Feng, Y., & Yi, X. S. (1999). Thermoplastic starch/PVAI compounds: Preparation,

377

processing and properties. Journal of Applied Polymer Science, 74, 2667-2673.

378

Luo, X., Li, J., & Lin, X. (2012). Effect of gelatinization and additives on morphology and

379

thermal behavior of corn starch/PVA blend films. Carbohydrate Polymers, 90, 1595-1600.

380

Ma, X. F., Yu, J. G., & Wan, J. J. (2006). Urea and ethanolamine as a mixed plasticizer for

381

thermoplastic starch. Carbohydrate Polymers, 64, 267-273.

382

Ma, X., Yu, J. (2004). The effects of plasticizers containing amide groups on the properties of

383

thermoplastic starch. Starch/Stärke, 56, 545–551.

384

Majdzadeh-Ardakani, K., & Nazari, B. (2010). Improving the mechanical properties of

385

thermoplastic starch/poly (vinyl alcohol)/clay nanocomposites. Composites Science and

386

Technology, 70, 1557-1563.

387

Nakayama, Y., Takatsuka, M., & Matsuda, T. (1999). Surface hydrogelation using photolysis of

388

dithiocarbamate or xanthate: Hydrogelation, surface fixation, and bioactive substance

389

immobilization. Langmuir, 15, 1667-1672.

390

Park, H., Chough, S., Yun, Y., & Yoon, S. (2005). Properties of starch/PVA blend films

391

containing citric acid as additive. Journal of Polymers and the Environment, 13(4), 375-382.

392

Pavlovic, S., & Brandao, P. R. G. (2003). Adsorption of starch, amylose, amylopectin and

393

glucose monomer and their effect on the flotation of hematite and quartz. Minerals Engineering,

394

16, 1117-1122.

395

Petersen, K., Nielsen, P.V., Bertelsen, G., Lawther, M., Olsen, M., Nilsson, N.H., & Mortensen,

396

G. (1999). Potential of biobased materials for food packaging. Trends in Food Science &

397

Technology, 10, 52-68.

398

Raj, B., Sidddaramaiah, S., & Somashekar, R. (2004). Structure-property relation in polyvinyl

399

alcohol/starch composites. Journal of Applied Polymer Science, 91, 630-635.

Ac ce p

te

d

M

an

us

cr

ip t

376

20 Page 20 of 24

Saiah, R., Sreekumar, P. A., Leblanc, N., & Saiter, J.-M. (2009). Structure and thermal stability

401

ofthermoplasticfilms basedon wheatflourmodifiedbymonoglyceride. Industrial Crops and

402

Products, 29, 241–247.

403

Shi, Q. F., Chen, C., Gao, L., Jiao, L., Xu, H. Y., & Guo, W. H. (2011). Physical and degradation

404

properties of binary or ternary blends composed of poly (lactic acid), thermoplastic starch and

405

GMA grafted POE. Polymer Degradation and Stability, 96, 175–182.

406

Shi, R., Bi, J., Zhang, Z., Zhu, A., Chen, D., Zhou, X., Zhang, L., & Tian, W. (2008). The effect

407

of citric acid on the structural properties and cytotoxicity of the polyvinyl alcohol/starch films

408

when molding at high temperature. Carbohydrate Polymers, 74, 763-770.

409

Sin, L. T., Rahman, W. A. W. A., Rahmat, A. R., & Mokhtar, M. (2011). Determination of

410

thermal stability and activation energy of polyvinyl alcohol–cassava starch blends. Carbohydrate

411

Polymers, 83, 303-305.

412

Sreedhar, B., Chattopadhyay, D. K., Karunakar, M. S. H., & Sastry, A. R. K. (2006). Thermal

413

and surface characterization of plasticized starch polyvinyl alcohol blends crosslinked with

414

epichlorohydrin. Journal of Applied Polymer Science, 101, 25-34.

415

Sreedhar, B., Sairam, M., Chattopadhyay, D. K., Syamala Rathnam, P. A., & Mohan Rao, D. V.

416

(2005). Thermal, mechanical and surface characterization of starch-poly(vinyl alcohol) blends

417

and borax-crosslinked films. Journal of Applied Polymer Science, 96, 1313-1322.

Ac ce p

te

d

M

an

us

cr

ip t

400

21 Page 21 of 24

Sreekumar, P. A., Al-Harthi, M. A., & De, S. K. (2012). Effect of glycerol on thermal and

419

mechanical properties of polyvinyl alcohol/starch blends. Journal of Applied Polymer Science,

420

123, 135-142.

421

Tang, X. Z., Alavi, S., & Herald, T. (2008). Effects of plasticizers on the structure and properties

422

of starch-clay nanocomposite films. Carbohydrate Polymers, 74, 552-558.

423

Tang, X., & Alavi, S. (2011). Recent advances in starch, polyvinyl alcohol based polymer blends,

424

nanocomposites and their biodegradability. Carbohydrate Polymers, 85, 7-16.

425

Vasile, C., Stoleriu, A., Popescu, M. C., Duncianu, C., Kelnar, I., & Dimonie, D. (2008).

426

Morphology and thermal properties of some green starch/poly (vinyl alcohol)/montmorillonite

427

nanocomposites. Cellulose Chemistry and Technology, 42(9-10), 549-568.

428

Wang, Y., Chang, C., & Zhang, L. (2010). Effects of Freezing/Thawing Cycles and Cellulose

429

Nanowhiskers

430

Macromolecular Materials and Engineering, 295, 137-145.

431

Weber, C.J., Haugaard, V., Festersen, R., & Bertelsen, G. (2002) Production and application of

432

biobased packaging materials for the food industry. Food Additives & Contaminants, 19, 172-

433

177.

434

Wolkers, W. F., Oliver, A. E., Tablin, F., & Crowe, J. H. (2004). A Fourier-transform infrared

435

spectroscopy study of sugar glasses. Carbohydrate Research, 339 (6), 1077-1085.

436

Yin, Y., Li, J., Liu, Y., & Li, Z. (2005). Starch crosslinked with poly(vinyl alcohol) by boric acid.

437

Journal of Applied Polymer Science, 96, 1394-1397.

438

Yoon, S. D., Chough, S. H., & Park, H. R. (2006a). Properties of starch-based blend films using

439

citric acid as additive. Journal of Applied Polymer Science, 100, 2554-2560.

Structure

and

Properties

of

Biocompatible

Starch/PVA

Sponges.

Ac ce p

te

d

on

M

an

us

cr

ip t

418

22 Page 22 of 24

Yoon, S. D., Chough, S. H., & Park, H. R. (2006b). Effects of additives with different functional

441

groups on the physical properties of starch/PVA blend film. Journal of Applied Polymer Science,

442

100, 3733-3740.

443

Yoon, S. D. (2014). Cross-Linked Potato Starch-Based Blend Films Using Ascorbic Acid as a

444

Plasticizer. Journal of Agricultural and Food Chemistry, 62, 1755-1764.

445

Zhou, X. Y., Cu, Y. F., Jia, D. M., & Xie, D. (2009). Effect of a complex plasticizer on the

446

structure and properties of the thermoplastic PVA/starch blends. Polymer Plastics Technology

447

and Engineering, 48(5), 489-495.

us

cr

ip t

440

Ac ce p

te

d

M

an

448

23 Page 23 of 24

448

Highlights -

PVA:starch blend films were prepared using a series of polyol-based plasticizers

450

-

5% (wt.) addition of the investigated plasticizers lowers the Tg point significantly

451

-

Different plasticization behaviors are observed with increasing plasticizer amount

452

-

Molecular geometry of plasticizers can prevent penetration into chain segments

453

-

1,4-butanediol shows the highest plasticizing effect for PVA:starch

cr

ip t

449

Ac ce p

te

d

M

an

us

454

24 Page 24 of 24