Accepted Manuscript Title: Effect of structure and viscosity of the components on some properties of starch-rich hybrid blends Author: Willian H. Ferreira Marwin M.I.B. Carmo Ana L´ucia N. Silva Cristina T. Andrade PII: DOI: Reference:

S0144-8617(14)01023-6 http://dx.doi.org/doi:10.1016/j.carbpol.2014.10.018 CARP 9371

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

19-8-2014 11-10-2014 13-10-2014

Please cite this article as: Ferreira, W. H., Carmo, M. M. I. B., Silva, A. L. N., and Andrade, C. T.,Effect of structure and viscosity of the components on some properties of starch-rich hybrid blends, Carbohydrate Polymers (2014), http://dx.doi.org/10.1016/j.carbpol.2014.10.018 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  2  3 

Effect of structure and viscosity of the components on some properties of starch-rich



hybrid blends

ip t



cr



Willian H. Ferreira, Marwin M. I. B. Carmo, Ana Lúcia N. Silva, Cristina T. Andrade*

us

7  8 

Instituto de Macromoléculas Professora Eloisa Mano, Universidade Federal do Rio de

10 

Janeiro, Centro de Tecnologia, Bloco J, P.O. Box 68525, 21941-598, Rio de Janeiro, RJ,

11 

Brazil

M

an



12 

d

13 

te

14 

Ac ce p

15  16  17  18  19  20  21  22 

*Corresponding author

23 

Telephone number: + 55 21 2562 7208

24 

Fax number: + 55 21 2270 1317

25 

E-mail address: [email protected] 1

 

Page 1 of 32

26 

Abstract Glycerol-plasticized cornstarch and poly(lactic acid) (PLA) were melt-blended alone and

28 

at a constant 70:30 (m/m) composition, in the present of an organoclay. The effect of

29 

increasing contents of the organoclay on extruded and compression-molded samples was

30 

evaluated by X-ray diffraction (XRD), capillary rheometry, thermogravimetric analysis

31 

(TGA), scanning electron microscopy (SEM), and tensile tests. XRD and shear viscosity

32 

results obtained for the hybrid components (TPS/organoclay and PLA/organoclay) were

33 

correlated with the hybrid blends properties. XRD and TGA results suggested that the

34 

organoclay was similarly dispersed within both phases. SEM images revealed improved

35 

adhesion

36 

compatibilization as the organoclay content was increased. Some of the extruded materials

37 

were also submitted to injection molding, and characterized by SEM and by tensile tests.

38 

For the extruded and compression-molded samples, improved mechanical properties were

39 

obtained for the samples with higher contents of the organoclay. For the injection-molded

40 

samples, the mechanical properties seemed to be dependent on the organoclay dispersion.

phases.

Shear

viscosities

results

indicated

improved

te

d

M

the

Ac ce p

41  42 

between

an

us

cr

ip t

27 

Keywords: Thermoplastic starch; Poly(lactic acid); Hybrid blends; Extrusion, Injection molding

43  44 

Chemical compounds studied in this article

45 

Starch (PubChem CID: 24836924); Poly(lactic acid) (not found)

46  47  48 

1. Introduction Poly(lactic acid) (PLA) and starch are considered promising biobased materials to

49  50 

substitute traditional petroleum-based polymers. PLA is a biocompatible and bioresorbable 2

 

Page 2 of 32

polyester, with mechanical properties similar to poly(ethylene terephathalate) and

52 

polystyrene. Until the late 1980s, the application of PLA was mainly focused on the

53 

biomedical field, as a biomaterial for hard tissue replacement or in controlled drug delivery

54 

systems. However, the high hydrophobic character and, consequently, slow hydrolytic

55 

degradation rate, was appointed as a disadvantage to some applications (Luckachan &

56 

Pillai, 2011).

cr

ip t

51 

The interest in starch-based materials has grown because of their relatively low cost,

us

57 

renewability and biodegradability in most environments. Starch consists of two

59 

polysaccharides, amylose and amylopectin, both formed by D-glucose repeating units. The

60 

semicrystalline structure of starch granules is destroyed by thermomecanical treatment in

61 

the presence of suitable plasticizers, leading to thermoplastic starch (TPS). However, the

62 

sensitivity of TPS to ambient humidity practically invalidates its commercial use. Because

63 

of this drawback, the development of TPS blends and composites has been investigated.

64 

Blending TPS with PLA would maintain the biodegradability of the resulting materials,

65 

and might improve their properties.

Ac ce p

te

d

M

an

58 

It was postulated that blend properties are determined by the structural characteristics of

66  67 

each component and the phase morphology developed during melt processing. The final

68 

morphology is controlled by blend composition, rheological properties of the components,

69 

interfacial tension and viscosity ratio between the components, and processing conditions

70 

(Sundararaj & Macosko, 1995; Wu, 1987). Much effort has been made to develop TPS/PLA blends. The low interfacial adhesion

71  72 

between hydrophilic starch and hydrophobic PLA has been recognized as the major

73 

problem to be overcome. To promote the compatibility of these materials, maleic

74 

anhydride was added in one or two stages (Huneault & Li, 2007; Wang, Yu, & Ma, 2007;

75 

Zhang & Sun, 2004). On the other hand, some authors followed three strategies; in situ 3

 

Page 3 of 32

formation of urethane linkages, in situ coupling of starch and PLA with peroxide, and

77 

addition of previously formed PLA-grafted amylose as compatibilizing agent (Schwarch,

78 

Six, & Avérous, 2008). Improvements in mechanical properties were obtained for the

79 

blends with PLA in excess, substituting PLA for maleated PLA or by using PLA-grafted

80 

amylose. Additionally, to prepare PLA/TPS blends, other authors used maleated PLA and

81 

sodium montmorillonite (MMT) at 2% content, as a compatibilizing and a reinforcing

82 

agent, respectively. Before processing, MMT was added either to a starch water-glycerol

83 

suspension or to PLA. For the blends with 60% TPS, the location of MMT at the TPS

84 

phase led to higher modulus and strength enhancement (Arroyo, Huneault, Favis, &

85 

Bureau, 2010). In another report, small amounts of MMT (0.5 and 1% based on potato

86 

starch dry mass) were also incorporated to a TPS/PLA blend at 60/40 composition, in a

87 

multi-stage process, which consisted of potato starch gelatinization, ultrasonic dispersion

88 

of MMT in water and melt mixing at 185ºC for 10 min. Reduction in water absorption and

89 

improvements in mechanical and dynamic-mechanical properties were observed (Ayana,

90 

Suin, & Khatua, 2014).

Ac ce p

te

d

M

an

us

cr

ip t

76 

Similarly to the stabilization of foams and oil-in-water emulsions (Horozov, 2008;

91  92 

Hunter, Pugh, Franks, & Jameson, 2008), nanoparticles may inhibit coalescence of

93 

dispersed phase droplets in immiscible polymer blends, mainly when located at their

94 

interface. However, the location of nanoparticles is dictated by thermodynamic

95 

interactions, kinetic and viscosity effects (Feng, Chan, & Li, 2003; Fenouillot, Cassagnau,

96 

& Majesté, 2009). Also, as postulated previously by these authors, the stabilization of

97 

morphology is a requisite to maintain constant properties (Fenouillot at al., 2009). In this work, glycerol-plasticized cornstarch was melt-blended with PLA at a 70:30

98  99  100 

constant composition without the addition of a compatibilizing agent, under one-step extrusion processing. Differently from the works cited before (Arroyo et al., 2010; Ayana 4

 

Page 4 of 32

et al., 2014), an organically-modified clay was also incorporated with increasing contents

102 

(1.0 to 10.0 mass%) to the polymeric components. Both polymeric components were also

103 

extruded and compression-molded alone and in the presence of the organoclay at the same

104 

contents, and were characterized by X-ray diffraction and capillary rheometry. The

105 

objectives of these experiments were to investigate the organoclay dispersion within TPS

106 

and PLA matrices, determine changes in flow behavior resulting from the filler addition,

107 

and mainly investigate the influence of these features on some properties of the resulting

108 

hybrid blends. For this, TPS/PLA/organoclay blended materials were also characterized by

109 

thermogravimetric analysis. The mechanical and morphological properties were

110 

investigated for the extruded and compression-molded samples and for some of the

111 

extruded materials , which were injection-molded .

112 

Experimental

113 

1.1 Materials

d

M

an

us

cr

ip t

101 

Regular cornstarch (CS) was supplied by Ingredion Brasil - Ingredientes Industriais

te

114 

Ltda. (São Paulo, SP, Brazil). According to the producer, this material is composed of 26–

116 

30 mass% amylose and 74–70 mass% amylopectin, with less than 0.5 mass% gluten. The

117 

gravimetric method was used to determine the humidity content (12 mass%). Analytical

118 

grade glycerol was purchased from Vetec Química Fina (Rio de Janeiro, RJ, Brazil) and

119 

was used as received. Poly(lactic acid) (PLA, NatureWorks LLC, Ingeo 2002D, with melt

120 

flow index of 4-8 g/10min at 190°C / 2.16 kg) was supplied as pellets by Cargill Agrícola

121 

S.A. (São Paulo, SP, Brazil). This PLA sample was previously characterized by gel

122 

permeation chromatography in chloroform, and the results reported as Mn = 121.100 and

123 

Mw = 197.000 (Souza, Dahmouche, Andrade, & Dias, 2013). PLA pellets were

124 

cryogenically ground into powder to ensure better mixing with the other components.

125 

Cloisite® 30B (C30B), a montmorillonite modified with dihydroxyethyl alkyl

Ac ce p

115 

5

 

Page 5 of 32

126 

methylammonium ions, was supplied by Southern Clay Products (Gonzales, TX, USA),

127 

and was used as received.

128 

1.2 Preparation of samples The mixture of CS and glycerol, added at 25 mass% based on starch dry mass, was

ip t

129 

homogenized in a conventional mixer (Ika Works, Wilmington, NC, USA) for 30 min, and

131 

maintained in tightly sealed polyethylene bags for 48 h at 4°C before processing to obtain

132 

TPS. Glycerol-plasticized CS and PLA were mixed separately with C30B at 1.0, 2.5, 5.0,

133 

7.5 and 10.0 mass% contents to obtain TPS/C30B and PLA/C30B materials. To obtain

134 

TPS/PLA/C30B hybrid blends, glycerol-plasticized CS and PLA were homogenized at a

135 

constant 70:30 (m/m) ratio and the organoclay was added at 1.0, 2.5, 5.0, 7.5 and 10.0

136 

mass% contents, over the total mass of the polymers.

137 

2.3. Extrusion and compression molding

M

an

us

cr

130 

Extrusion processings were carried out in a Coperion ZSK 18 (Werner & Pfleiderer,

139 

Stuttgart, Germany) co-rotating twin-screw extruder, with a L/D ratio of 40 and seven

140 

heating zones. For glycerol-plasticized CS with various amounts of C30B, the seven

141 

heating zones were maintained at 110, 110, 115, 115, 115, 110, 110°C, and the screw

142 

speed was set at 120 rpm. For PLA and the blends (obtained by one-step extrusion

143 

processings) with varying amounts of C30B, the seven heating zones were maintained at

144 

160, 160, 165, 165, 165, 160, 160°C, and the screw speed was set at 120 rpm. The

145 

PLA/C30B materials showed free-flowing behavior and were cooled in a water bath after

146 

extrusion. All extruded materials were pelletized and compression-molded by heating at

147 

different temperatures (TPS at 120ºC, PLA and the hybrid blends at 165ºC) under 68.9

148 

MPa for 10 min, and cooling for 5 min in a cold press. The resulting sheets, with 12 cm x

149 

10 cm x 2 cm dimensions, were used for the materials characterization.

150 

2.4. X-ray diffraction (XRD)

Ac ce p

te

d

138 

6

 

Page 6 of 32

The samples were analyzed with a Ultima IV diffractometer (Rigaku Corporation) in the

151 

angular region 0.9 to 10° (2θ) at a 0.01º/s rate, in reflection mode. The CuKα1 radiation

153 

with wavelength of 1.542 Å was generated at 40 kV and 20 mA. The d-values were

154 

calculated by the Bragg's law (Equation 1). nλ = 2 d sen θ

155 

Equation 1

ip t

152 

where n is an integer, λ is the wavelength of incident wave (1.5418 Å), d is the spacing

157 

between any two atomic planes in the crystal lattice, and θ is the angle of reflection.

158 

2.5. Capillary viscosity measurements

us

cr

156 

A Göttffert capillary rheometer model Rheograph 25 (Buchen, Germany), equipped

an

159 

with a round-hole capillar of D = 1 mm (L/D = 30), was used to measure the flow

161 

properties of TPS/C30B, PLA/C30B extruded materials and the hybrid blends. All tests

162 

were performed over a shear rate range of 50 – 3000 s-1, at 165ºC with a prior melting

163 

time of 5 min. End corrections were not applied; thus, the viscosity values are reported as

164 

apparent viscosities. The setting of parameters and evaluation of raw data, as well as the

165 

Rabinowitsch’s correction, were performed with the software LabRheo, provided with the

166 

equipment. The consistency index and the flow behavior index were obtained according to

167 

the power law model (Equations 2 and 3).

Ac ce p

te

d

M

160 

168 

169 

τ=K

Equation 2

η=K

Equation 3

170 

where τ is the shear stress, is the shear rate at the capillary wall, η is the viscosity, K is

171 

the consistency index and n is the flow behavior index of the materials analyzed.

172 

2.6. Thermogravimetric analysis

7

 

Page 7 of 32

Thermogravimetric analyses were carried out for the extruded materials under nitrogen

173 

atmosphere on a TGA Q-500 equipment from TA Instruments (New Castle, USA).

175 

Approximately, 10 mg of sample were heated from 25 to 700ºC, at a 10ºC/min rate.

176 

2.7. Phase morphology

ip t

174 

The phase morphology of the samples was examined with a Jeol electron microscope,

178 

model JSM-6460LV (Akishima-shi, Japan), generally at the acceleration voltage of 20 kV.

179 

Samples were cryogenically fractured before observation. The fractured surfaces were

180 

vacuum-coated with gold before measurements. X-ray photoelectron spectroscopy (XPS)

181 

was used to analyze the cryofractured surfaces of injection-molded samples in a

182 

NORAN™ System 6 X-ray Microanalysis System, model C10015 (Thermo Fisher

183 

Scientific Inc., Middleton, WI, USA), linked to the Jeol electron microscope.

184 

2.8. Tensile tests of extruded and compression-molded samples

M

an

us

cr

177 

Tensile tests were carried out at 21ºC in a Q-800 DMA equipment from TA Instruments

d

185 

(New Castle, DE, USA), according to the ISO 527-3 method, with extruded and

187 

compression-molded specimens with 13 ± 0.2 mm x 6 ± 0.2 mm x 0.8 ± 0.1 mm

188 

dimensions, conditioned at 21ºC, and 50% relative humidity, for a period of at least 48 h.

189 

The average value from a total of three measurements was taken. The samples were placed

190 

between a fixed and a moveable clamp, and the experiments were performed in tensile

191 

mode.

192 

2.9. Injection molding

Ac ce p

te

186 

The pelletized samples were subsequently dried in an oven at 100ºC for 2 h. Samples for

193  194 

tensile tests were prepared in an injection-molding machine Arburg, model Allrounder

195 

270S ( Lossburg, Germany) to obtain specimens according to ASTM D638, type I

196 

(thickness 1.8 ± 0.1 mm). The five heating zones of the injection molding machine were

197 

maintained at 170, 180, 190 and 200°C and 210ºC (from feed to die zone). The mold 8

 

Page 8 of 32

temperature was set at 30°C. The injection and the back pressures were 90 MPa and 57

199 

MPa, respectively. The injection speed was 6 x 10-5 m3/s and the back time was 4 s. Before

200 

performing mechanical tests, the specimens were conditioned at 21ºC and 50% relative

201 

humidity for 48 h.

202 

2.10. Tensile tests of extruded and injection-molded samples

ip t

198 

Tensile tests were performed in a Instron Universal Testing Machine model 4204

204 

(Norwood, MA, USA) equipped with a 1 kN load cell, at a speed of 1 mm/min. The

205 

average value from a total of seven measurements was taken. All data were recorded and

206 

processed by the software provided with the Instron equipment.

207 

Results and discussion

an

us

cr

203 

A diffractometer with a high resolution at low angles was used to investigate C30B

M

208 

dispersion within TPS and PLA extruded materials. In Figure 1(a-c) (trace I), the 001

210 

reflection for the neat C30B was observed at 2θ = 4.9º, and revealed a d001 value of 1.8 nm.

211 

For TPS/C30B materials up to 5.0 mass% organoclay, no reflection was detected within

212 

the 0.9 – 10º (2θ ) region (Figure 1a). This result indicated that the 001 reflection of C30B

213 

was shifted to angles bellow 0.9° (2θ), with interlayer spacings higher than 9.8 nm. For the

214 

TPS samples with C30B at 7.5 and 10 mass% contents, broad reflections were observed

215 

around 4.5º and 4.3º (2θ), corresponding to d-values of approximately 2.0 nm. This result

216 

may be attributed to the presence of a fraction of C30B aggregates, in which starch

217 

molecules were partially intercalated. Although some authors reported on the

218 

incompatibility of hydrophilic thermoplastic starch with C30B (Chiou et al., 2006), the

219 

results of this work may be compared to those obtained by compounding potato starch with

220 

C30B in an internal mixer (Park, Lee, Park, Cho, & Ha, 2003). In this latter work, no

221 

reflection was found when C30B was incorporated at 2.5 mass% content, and a slight shift

222 

in the organoclay 001 reflection, from around 4.71º (2θ) to around 4.30º (2θ), was

Ac ce p

te

d

209 

9

 

Page 9 of 32

observed for the hybrids with 5 and 10 mass% C30B. The presence of two hydroxyethyl

224 

moieties in the intercalant of C30B explains its interaction with starch molecules. On the

225 

other hand, for all PLA/C30B materials (Figure 1b), low-intensity reflections around 2.6º

226 

(2θ) were observed. The intensity of these reflections increased slightly with the increase

227 

in organoclay content, and indicated that the organoclay had interlayer spacings of

228 

approximately 3.4 nm, intercalated with PLA molecules. Similar results were obtained by

229 

other authors for PLA and C30B (Araújo, Botelho, Oliveira, & Machado, 2014; Duan,

230 

Thomas, & Huang, 2013; Fukushima, Tabuani, & Camino, 2012; Lai, Wu, Lin, & Don,

231 

2014; Molinaro et al., 2013; Pluta, 2006; Zaidi et al., 2010), and attributed to hydrogen

232 

bonding interaction between the hydroxyethyl groups of the oraganoclay intercalant and

233 

carbonyl ester groups of PLA macromolecules. Anyway, the difference between PLA and

234 

TPS intercalation might be ascribed to the bulky ring structure of TPS, because in both

235 

cases hydrogen bonding interactions are favored.

d

M

an

us

cr

ip t

223 

It would be worth observing the diffratograms obtained for the TPS/PLA/C30B

te

236 

materials. In Figure 1c (traces II – IV), following the results for TPS/C30B, no reflection

238 

was observed in the diffratograms for TPS/PLA hybrids with 1.0, 2.5 and 5.0 mass%

239 

C30B, which indicated that the 001 reflection of C30B was shifted to angles bellow 0.9°

240 

(2θ), with interlayer spacings higher than 9.8 nm. For the hybrids with 7.5 and 10.0 mass%

241 

C30B, shallow reflections around 3º and 4.5 - 5º (2θ) were envisaged. In this case, the

242 

reflection around 3º (2θ) was attributed to C30B layers intercalated with starch molecules,

243 

and the other broader and less intense reflection at 4.5 – 5º (2θ) was attributed to a very

244 

small amount of C30B aggregates. Anyway, these diffractograms corroborated the good

245 

affinity of C30B towards both TPS and PLA phases, and suggested an even better

246 

dispersion of the organoclay in the extruded blends in relation to TPS or PLA alone.

Ac ce p

237 

10

 

Page 10 of 32

The melt flow characteristics of all extruded materials considered in this work were

248 

investigated by capillary rheometry at 165ºC, and the resulting curves are shown in Figure

249 

2 (a-c), as apparent viscosity versus shear rate. Shear thinning behavior was observed for

250 

all samples. Comparing TPS and PLA alone, the virgin PLA presented higher viscosity

251 

values, but a less pronounced shear thinning behavior compared to TPS, as indicated by the

252 

higher flow index (n) value, shown in Table 1. However, for the extruded PLA sample, a

253 

severe reduction of viscosity was observed. In general, processing of PLA is preceded by

254 

humidity elimination, but even when this procedure was followed, degradation was

255 

observed (Souza et al., 2013). Viscosity reduction of PLA was attributed to both hydrolysis

256 

reaction and thermal degradation (Speranza, De Meo, & Pantani, 2014). Since the

257 

objective of the extrusion processing was to produce TPS/PLA hybrid blends, in which all

258 

components are hygroscopic, drying seemed an unrealistic procedure.

M

an

us

cr

ip t

247 

For TPS/C30B materials (Figure 2a), the increase in organoclay content, from 1.0 to 10

d

259 

mass% content, led to almost a half-decade increase of apparent viscosity, and indicated

261 

that both exfoliated (at a low C30B content), and particularly intercalated organoclay

262 

contributed positively to viscosity. For these materials, the flow index n (Table 1) showed

263 

typical values, within the range (0.27 – 0.56) previously reported for cornstarch submitted

264 

to thermomechanical treatments at different moisture contents (Vergnes & Villemaire,

265 

1987). On the other hand, a substantial reduction of viscosity and n values was observed

266 

for PLA/C30B materials with increasing organoclay content (Figure 2b, Table 1).

267 

Recently, the thermal decomposition process of a similar system composed of PLA and

268 

C30B at 0.5 and 2.5 mass% content was reported, and the nanocomposites were found

269 

more stable than PLA alone (Carrasco, Pérez-Maqueda, Santana, & Maspoch, 2014). In the

270 

last case, a dehumidifier was used prior to extrusion. When comparing the two hybrids, the

271 

authors observed that the incorporation of 2.5 mass% C30B to PLA led to a less thermally

Ac ce p

te

260 

11

 

Page 11 of 32

stable sample, attributed to the presence of C30B aggregates. In this study, the

273 

diffractograms of Figure 1b revealed that the extrusion conditions led to intercalated C30B

274 

hybrids. Thus, the reduction of viscosity, mainly at 7.5 and 10.0 mass% C30B contents,

275 

and at high shear rates, might be indicating hydrolytic degradation to such an extent that

276 

topological constraints were extensively minimized.

ip t

272 

Viscosity curves were also obtained for the 70:30 TPS/PLA blends, with or without the

cr

277 

addition of C30B (Figure 2c, Table 1). As expected, the melt viscosity for the neat

279 

TPS/PLA blend resulted in intermediary values between those determined for TPS and

280 

PLA, and the incorporation of C30B contributed to increase the blends viscosity. The flow

281 

index values seemed to be governed by the properties of TPS, the component added at a

282 

larger composition. However, no clear trend was followed by the consistency index K. It

283 

might be inferred that the more pronounced Newtonian character of the TPS/PLA hybrid

284 

with C30B at 10.0 mass% content (Table 1) would worsen conditions for further

285 

processing, such as injection molding.

te

d

M

an

us

278 

The thermal stability of TPS, PLA, and TPS/PLA blends with and without the addition

287 

of C30B was evaluated by thermogravimetric analysis (TGA) under nitrogen atmosphere.

288 

The TGA curves are shown in Figure 3 (a,b). Up to 100ºC, mass loss is mainly attributed

289 

to humidity loss. From 100ºC to decomposition temperature, the mass loss observed for

290 

TPS and TPS/PLA/C30B at 1.0 mass% might be attributed to glycerol evaporation. With

291 

increasing organoclay content, a reduction of plasticizer loss was observed. The neat

292 

TPS/PLA blend showed the lowest onset degradation temperature (To = 284ºC), whereas

293 

TPS (To = 297ºC) and PLA (To = 332ºC) had higher thermal stabilities. For the hybrid

294 

blends, To decreased moderately, from 303ºC to 288ºC, as the C30B content was increased.

295 

It is worth observing in Figure 3 (c,d) (DTG curves) that the maximum degradation rate

296 

was reached for PLA at Tmax = 361ºC, and for TPS at Tmax = 327ºC. While a single peak at

Ac ce p

286 

12

 

Page 12 of 32

Tmax = 301ºC was observed for the neat TPS/PLA blend, two peaks were observed for the

298 

hybrid samples. The first peak, attributed to the maximum degradation rate of the TPS

299 

phase, varied from Tmax = 302ºC to Tmax = 313ºC with increasing C30B content. Lower

300 

Tmax values for TPS/calcium montmorillonite hybrids in relation to TPS were previously

301 

observed, and were attributed to the catalytic effect of acidic sites of montmorillonite on

302 

the thermal degradation of the starch matrix (Magalhães & Andrade, 2010). Similar result

303 

was also observed for other hybrid systems (Ramos Filho, Mélo, Rabello, & Silva, 2005).

304 

The second peak, attributed to the maximum degradation rate of the PLA phase, varied

305 

from Tmax = 357ºC to Tmax = 369ºC with increasing C30B content. This result is in

306 

accordance with some authors, to whom the preparation of nanocomposites by adding

307 

organo-modified montmorillonites should be considered as a promising method to reduce

308 

PLA thermal degradation (Carrasco et al, 2014). In relation to PLA, other authors observed

309 

a slight increase in Tmax for the PLA/C30B nanocomposites at 1.0, 3.0 and 5.0 phr

310 

organoclay content (Lai et al., 2014).

te

d

M

an

us

cr

ip t

297 

In this case, the TGA data also corroborated viscosity results for the TPS/PLA/C30B

312 

hybrids (Figure 2), which suggested the protecting action of TPS against PLA degradation.

313 

More important, DTG curves gave additional evidence on the location of clay mineral

314 

particles in both TPS and PLA phases, as suggested by XRD results (Figure 1). Also, as

315 

observed by some authors (Lai et al., 2014), the nucleating effect of the organoclay

316 

particles should not be neglected. XRD data in the 2θ region of 2º to 35º revealed PLA

317 

characteristic reflections with increasing intensities, as the content of C30B was increased

318 

(Figure 1, Supplementary data).

Ac ce p

311 

As in most polymer blends, TPS and PLA are immiscible. Melt blending was performed

319  320 

without the addition of an organic compatibilizer and, as expected, the image of Figure 4a,

321 

obtained for the neat TPS/PLA blend, shows a coarse morphology. Immiscibility was 13

 

Page 13 of 32

evidenced by the size, and the lack of adhesion between the two phases. For the extruded

323 

and compression-molded TPS/PLA/C30B materials in Figure 4 (b-f), the micrographs

324 

revealed that, as the organoclay content was increased from 2.5 to 10 mass%, at 10

325 

mass% the PLA dispersed phase became undistinguished from the TPS matrix. This

326 

result was related to the role of the organoclay in the improvement of the system

327 

morphology. . Considering that the organoclay interacted similarly with both components,

328 

and considering the viscosity results obtained, a rough estimation of viscosity ratios might

329 

be envisaged for the 70:30 TPS/PLA blends with added C30B. With increasing C30B

330 

contents, the viscosity ratio p and the capillary number might be expected to decrease, and

331 

p might approach the value p = 1. Contrarily to the morphology improvement, previously

332 

reported for C30B nanocomposites of TPS and poly(3-hydroxybutyrate-co-3-

333 

hydroxyvalerate) (PHBV) (Magalhães, Dahmouche, Lopes, & Andrade, 2013), no clear

334 

visualization of size reduction of the PLA dispersed phase was observed, as the content of

335 

C30B was increased from 2.5 to 10 mass%. Thus, no indication on the location of the

336 

organoclay nanoparticles at the interface between the components could be suggested.

337 

Also, it was claimed that clay mineral particles would be incorporated preferentially into

338 

the polymer with the lowest melting temperature (Fenouillot et al., 2009). In this case, TPS

339 

has a lower melting temperature than PLA. However, as shown in Figure 2 for TPS/C30B

340 

and PLA/C30B hybrids, with increasing contents of the organoclay, while the viscosity of

341 

TPS was increased, the viscosity of PLA was decreased, and the melting pattern might be

342 

changed even during the extrusion step. As the viscosity of the polymeric components

343 

seems to interfere on the location of the clay minerals particles (Naderi, Lafleur, & Dubois,

344 

2008), it might be presumed that C30B tended to be almost equally incorporated in both

345 

TPS and PLA phases during processing, and eventually at the interface, as its content was

346 

increased. However, it seems as if the amount of organoclay was not enough to saturate

Ac ce p

te

d

M

an

us

cr

ip t

322 

14

 

Page 14 of 32

347 

the interface between the TPS matrix and PLA droplets, probably because exfoliated

348 

organoclay layers or organoclay aggregates were preferentially located within both

349 

polymers. The mechanical properties of the TPS/PLA hybrid blends were investigated by two

351 

methods, using compression-molded (Figure 5) and injection-molded specimens (Figure

352 

6). These results were analyzed independently, since they were acquired by different

353 

methods. In Figure 5a, for compression-molded samples, the rigid character of PLA, to

354 

which the elastic modulus was determined as E = 3146.3 ± 311.8 MPa, and the soft nature

355 

of TPS were once more revealed. In Figure 5b, even for the neat TPS/PLA blend, the

356 

elastic modulus (190.2 ± 18.1 MPa) increased in relation to TPS alone (E = 82.3 ± 7.9

357 

MPa). For the compression-molded hybrids, the moduli varied from E = 167.5 ± 14.8 MPa

358 

to 501.4 ± 52.2 MPa, by incorporating 1.0 to 7.5 mass% C30B. This result was expected

359 

because of the rigid nature of the filler. However, a substantial decrease in tensile stress at

360 

break (σmax) was observed for the neat TPS/PLA blend (σmax = 1.5 ± 0.1 MPa) and the

361 

C30B hybrids, in relation to TPS and PLA. For TPS, this value was σmax = 4.4 ± 0.4 MPa

362 

and, for PLA, σmax = 5.3 ± 0.5 MPa. For the TPS/PLA hybrids, σmax varied from 1.7 ± 0.2

363 

MPa for C30B incorporated at a 1.0 mass% content to σmax = 3.4 ± 0.4 MPa for the hybrid

364 

with C30B at a 7.5 mass% content. As for the elongation at break, εmax, their values varied

365 

for the same hybrids, respectively, from 0.8 ± 0.1% to 1.3 ± 0.01%, in relation to εmax =

366 

15.2 ± 1.1% for TPS.

Ac ce p

te

d

M

an

us

cr

ip t

350 

For the hybrid prepared with C30B at 10.0 mass% content, the E and σmax values

367  368 

decreased to E = 383.5 ± 41.2 MPa and σmax = 2.9 ± 0.3 MPa, respectively. This result was

369 

not expected, as long as increasing values were obtained with increasing contents of C30B.

370 

However, the plasticizing effect of the C30B intercalant, also incorporated at a higher 15

 

Page 15 of 32

371 

amount, might have contributed to the less rigid character of this hybrid, for which εmax

372 

reached 1.9 ± 0.01%. It is worth noting that the relatively better results of tensile tests (in relation to the neat

373 

TPS/PLA blend) for the hybrids, to which C30B was incorporated at 5.0 to 10.0 mass%

375 

contents, followed the previously rough estimation for viscosity ratios. Because of this

376 

result, these extruded samples as well as the neat 70:30 TPS/PLA blend were chosen to be

377 

submitted to injection molding, to investigate possible changes of morphology and

378 

mechanical properties.

us

cr

ip t

374 

The mechanical properties of the injection-molded samples were evaluated (Figure 6).

380 

The highest stress at break (σmax = 18.9 MPa) was obtained for the neat TPS/PLA blend.

381 

Differently from the previous result obtained for the compression-molded sample, the

382 

injection-molded hybrid with 10.0 mass% C30B did not show the highest ductibility. The

383 

highest elastic modulus (E = 11.4 ± 3.8 MPa) and lowest elongation at break (εmax = 0.55 ±

384 

0.16%) were determined for the hybrid with 7.5 mass% C30B.

te

d

M

an

379 

The best ductibility was achieved for the TPS/PLA/C30B hybrid with 5.0 mass%

Ac ce p

385  386 

organoclay (E = 1091.5 ± 181.3 MPa, σmax = 11.2 ± 2.2 MPa, εmax = 2.33 ± 1.4%). Bearing

387 

in mind XRD results, the incorporation of 5.0 mass% C30B led to a higher degree of

388 

exfoliation, which certainly contributed to the better fluidity during injection molding and

389 

the highest ductibility. Moreover, the more pronounced Newtonian character of the

390 

TPS/PLA hybrid with C30B at 10.0 mass%, as shown before (Table 1), indeed made

391 

injection molding more difficult. To verify the effect of injection molding on morphology, cryofractured surfaces of the

392  393 

injection-molded samples were analyzed by SEM (Figure 7). Although more sensitive to

394 

the electron beam, the neat TPS/PLA blend in Figure 7a showed a similar image to that

395 

taken for the compression-molded sample (Figure 4a). The SEM images for the hybrids 16

 

Page 16 of 32

with 5.0 and 7.5 mass% C30B in Figure 7 (b,c) were characterized by the appearance of

397 

microvoids much larger in diameter than those observed for the corresponding

398 

compression-molded samples. The appearance of these microvoids might be related to the

399 

loss of some droplets of PLA dispersed phase during cryofracturing. If this supposition

400 

were true, it indicated a very small amount of C30B at the interface. No particular feature

401 

seems to distinguish the image observed for the injection-molded hybrid with C30B at 10

402 

mass% content from that obtained for the compression-molded sample with the same

403 

content of the filler. This result might be probably related to a higher amount of the

404 

organoclay at the interface.

an

us

cr

ip t

396 

Using XPS to analyze the cryofractured surfaces of injection-molded samples, the

406 

results revealed widespread distribution of Mg, Al and Si (Supplementary data, Figures 2-

407 

4), according to the previous supposition on incorporation of C30B within both

408 

components phases.

409 

Conclusions

te

d

M

405 

The favorable interaction between the organoclay intercalant and the molecules of both

Ac ce p

410  411 

components, starch and PLA, suggested by XRD and TGA results, seemed to dictate the

412 

properties of extruded TPS/PLA/C30B hybrids. XRD data suggested no specific

413 

preference for one of the blend components. TGA indicated the stabilizing effect of C30B

414 

on the PLA phase, as the content of the organoclay increased. Shear viscosities,

415 

determined for each component in the presence of the same organoclay contents, roughly

416 

indicated the tendency for achieving a viscosity ratio around p = 1, as higher contents of

417 

organoclay were incorporated. For the extruded and compression-molded samples, the

418 

results of tensile tests followed this estimation. The SEM images revealed that the

419 

morphology, even after injection molding, remained the same for the TPS/PLA hybrid

420 

blend, prepared with C30B at the highest content (10.0 mass%). The high degree of 17

 

Page 17 of 32

organoclay dispersion, associated with the highest pseudoplasticity of the 5.0 mass%

422 

hybrid blend, and the structuring characteristic of the process, certainly contributed to the

423 

highest ductibility of this TPS/PLA hybrid material after processing by injection molding.

424 

Acknowledgments

ip t

421 

The authors acknowledge Conselho Nacional de Desenvolvimento Científico e

426 

Tecnológico (CNPq), Fundação de Amparo à Pesquisa do Estado do Rio de Janeiro

427 

(FAPERJ), and Coordenação para o Aperfeiçoamento de Pessoal de Nível Superior

428 

(CAPES) for financial support.

429 

References

430 

Araújo, A., Botelho, G., Oliveira, M., & Machado, A.V. (2014). Influence of clay organic

431 

modifier on the thermal-stability of PLA based nanocomposites. Applied Clay Science,

432 

88-89, 144-150.

us an

M

Arroyo, O. H., Huneault, M. A., Favis, B. D., Bureau, M. N. (2010). Processing and

d

433 

cr

425 

properties of PLA/thermoplastic starch/montmorillonite nanocomposites. Polymer

435 

Composites, 31, 114–127.

Ac ce p

436 

te

434 

Ayana, B., Suin, S., & Khatua, B. B. (2014). Highly exfoliated eco-friendly thermoplastic starch (TPS)/poly (lactic acid)(PLA)/clay nanocomposites using unmodified

437 

nanoclay. Carbohydrate Polymers, 110, 430–439.

438  439 

Carrasco, F., Pérez-Maqueda, L. A., Santana, O.O., & Maspoch, M. Ll. (2014). Enhanced general analytical equation for the kinetics of the termal degradation of poly(lactic

440  441 

acid)/montmorillonite nanocomposites driven by random scission. Polymer

442 

Degradation and Stability, 101, 52-59.

443 

Chiou, B.-S., Yee, E., Wood, D., Shey, J., Glenn, G., & Orts, W. (2006). Effects of

444 

processing conditions on nanoclay dispersion in starch-clay nanocomposites. Cereal

445 

Chemistry, 83, 300-305. 18

 

Page 18 of 32

446 

Duan, Z., Thomas, N. L., & Huang, W. (2013). Water vapour permeability of poly(lactic acid) nanocomposites. Journal of Membrane Science, 445, 112-118.

447  448 

Feng, J., Chan, C. M., & Li, J. X. (2003). A method to control the dispersion of carbon black in an immiscible polymer blend. Polymer Engineering and Science, 43, 1058–

450 

1063.

Fenouillot, F., Cassagnau, P., & Majesté, J.-C. (2009). Uneven distribution of

cr

451 

ip t

449 

nanoparticles in immiscible fluids: morphology development in polymer blends.

453 

Polymer, 50, 1333–1350.

Fukushima, K., Tabuani, D., & Camino, G. (2012). Poly(lactic acid)/clay nanocomposites:

an

454 

us

452 

effect of nature and content of clay on morphology, thermal and thermo-mechanical

456 

properties. Materials Science and Engineering C 32, 1790-1795.

457 

M

455 

Horozov, T. S. (2008). Foams and foam films stabilised by solid particles. Current Opinion in Colloid & Interface Science, 13, 134–140. Huneault, M. A., Li, H. (2007). Morphology and properties of compatibilized

te

459 

d

458 

polylactide/thermoplastic starch blends. Polymer, 48, 270–280.

Ac ce p

460  461 

Hunter, T. N., Pugh, R. J., Franks, G. V., & Jameson, G. J. (2008). The role of particles in

462 

stabilising foams and emulsions. Advances in Colloid and Interface Science, 137, 57–

81.

463  464 

Lai, S.-M., Wu, S.-H., Lin, G.-G, & Don, T.-M. (2014). Unusual mechanical properties of melt-blended poly(lactic acid) (PLA)/clay nanocomposites. European Polymer

465 

Journal, 52, 193-206.

466  467 

Luckachan, G.E., & Pillai, C.K.S. (2011). Biodegradable polymers – a review on recent

468 

trends and emerging perspectives. Journal of Polymers and the Environment, 19, 637-

469 

676.

19

 

Page 19 of 32

470 

Magalhães, N. F., & Andrade, C. T. (2010). Calcium bentonite as reinforcing nanofiller for thermoplastic starch. Journal of the Brazilian Chemical Society, 21, 202-208.

471  472 

Magalhães, N. F., Dahmouche, K., Lopes, G. K., & Andrade, C. T. (2013). Using an organically-modified montmorillonite to compatibilize a biodegradable blend. Applied

474 

Clay Science, 72, 1–8.

ip t

473 

Molinaro, S., Romero, M. C., Boaro, M., Sensidoni, A., Lagazio, C., Morris, M., & Kerry,

476 

J. (2013). Effect of nanoclay-type and PLA optical purity on the characteristics of

477 

PLA-based nanocomposite films. Journal of Food Engineering, 117, 113-123.

478 

Naderi, G., Lafleur, P. G., & Dubois, C. (2008). The influence of matrix viscosity and

an

us

cr

475 

composition on the morphology, rheology, and mechanical properties of thermoplastic

480 

elastomer nanocomposites based on EPDM/PP. Polymer Composites, 29, 1301-1309. Park, H.-M., Lee, W., Park, C.-Y., Cho, W.-J., & Ha, C.-S. (2003). Environmentally

d

481 

M

479 

friendly hybrids. Part I Mechanical, thermal, and barrier properties of thermoplastic

483 

starch/clay nanocomposites. Journal of Materials Science, 38, 909-915.

te

482 

Pluta, M. (2006). Melt compounding of polylactide/organoclay: structure and properties of

485 

nanocmposites. Journal of Polymer Science Part B: Polymer Physics, 44, 3392-3405.

Ac ce p

484 

486 

Ramos Filho, F. G., Mélo, T. J. A., Rabello, M. S., & Silva, S. M. L. (2005). Thermal

487 

stability of nanocomposites based on polypropylene and bentonite. Polymer Degradation and Stability, 89, 383-392.

488  489 

Schwach, E., Six, J.-L., & Avérous, L. (2008). Biodegradable blends based on starch and

490 

poly(lactic acid): comparison of different strategies and estimate of compatibilization.

491 

Journal of Polymers and the Environment, 16, 286–297.

20

 

Page 20 of 32

492 

Souza, D. H. S., Dahmouche, K., Andrade, C. T., & Dias, M. L. (2013). Synthetic

493 

organofluoromica/poly(lactic acid) nanocomposites: structure, rheological and thermal

494 

properties. Applied Clay Science, 80-81, 259-266. Speranza, V., De Meo, A., & Pantani, R. (2014).Thermal and hydrolytic kinetics of PLA in the molten state. Polymer Degradation and Stability, 100, 37-41.

497 

cr

496 

ip t

495 

Sundararaj, U., & C. W. Macosko, C. W. (1995). Drop breakup and coalescence in polymer blends: the effects of concentration and compatibilization. Macromolecules,

499 

28, 2647-2657.

Vergnes, B., & Villemaire, J. P. (1987). Rheological behaviour of low moisture molten

an

500 

maize starch. Rheological Acta, 26, 570-576.

501 

Wang, N., Yu, J., & Ma, X. (2007). Preparation and characterization of thermoplastic

M

502 

us

498 

starch/PLA blends by one‐step reactive extrusion. Polymer International, 56, 1440–

Ac ce p

te

d

503 

1447.

504  505 

Wu, S. (1987). Formation of dispersed phase in incompatible polymer: interfacial and rheological effects. Polymer Engineering and Science, 27, 335-343.

506  507 

Zaidi, L., Bruzaud, S., Bourmaud, A., Médéric, P., Kaci, M., & Grohens, Y. (2010). Journal of Applied Polymer Science, 116, 1357-1365.

508  509 

Zhang, J. - F., & Sun, X. (2004). Mechanical properties of poly(lactic acid)/starch composites compatibilized by maleic anhydride. Biomacromoleules, 5, 1446-1451.

510  511  512 

21

 

Page 21 of 32

513  514  515 

ip t

516  517 

cr

518 

us

519  520 

an

521  522 

M

523  524 

d

525 

te

526 

Ac ce p

527  528  529 

Table 1: Power law constants for glycerol-plasticized thermoplastic starch, poly(acid acid)

530 

and their blends, with or without the incorporation of the C30B.

531 

K (Pa.s)

n

R2

TPS

10 399

0.27

0.998

TPS/ C30B (1.0 mass%)

9 727

0.31

0.999

TPS/ C30B (2.5 mass%)

9 332

0.36

0.999

TPS/ C30B (5.0 mass%)

13 489

0.31

0.999

Blend composition

22

 

Page 22 of 32

21 281

0.28

0.998

TPS/C30B (10.0 mass%)

25 577

0.31

0.999

PLA (as received)

44 668

0.40

0.995

964

0.32

0.996

PLA/C30B (1.0 mass%)

100 025

0.17

PLA/C30B (2.5 mass%)

51 286

0.25

PLA/C30B (5.0 mass%)

41 686

0.15

PLA/C30B (7.5 mass%)

125 892

PLA/C30B (10.0 mass%)

25 150

TPS/PLA

0.998

us

cr

0.995 0.996

0.999

-0.04

0.998

13 423

0.37

0.998

TPS/PLA/C30B (1.0 mass%)

25 118

0.29

0.997

TPS/PLA/C30B (2.5 mass%)

40 738

0.34

0.994

81 234

0.28

0.999

51 121

0.39

0.996

21 579

0.60

0.998

an

-0.44

M

PLA (after extrusion)

ip t

TPS/C30B (7.5 mass%)

te

TPS/PLA/C30B (7.5 mass%)

d

TPS/PLA/C30B (5.0 mass%)

TPS/PLA/C30B (10.0 mass%)

Ac ce p

532  533  534 

Caption to the figures

535 

Figure 1: X-ray diffractograms in the 2θ region of 0.6º to 10º for the neat C30B (traces I);

536 

(a) TPS, (b) PLA, (c) 70:30 TPS/PLA blend with the incorporation of C30B at 1.0 (traces II), 2.5 (traces III), 5.0 (traces IV), 7.5 (traces V), 10.0 (traces VI) mass%.

537  538 

Figure 2: Variation of apparent viscosity as a function of shear rate at 165ºC for

539 

(a) TPS alone (■) and hybrids, (b) PLA alone (■) and hybrids, and (c) neat 70:30

540 

TPS/PLA blend (■) and hybrids, with the incorporation of C30B at 1.0 (U), 2.5 (…),

541 

5.0 ({), 7.5 (V), and 10.0 (‘) mass%.

23

 

Page 23 of 32

Figure 3: TGA curves for (a) TPS (¯) and PLA ( ), and (b) neat 70:30 TPS/PLA blend

543 

(■) and hybrids, (c) DTG curves for TPS (¯) and PLA ( ), and (d) neat 70:30

544 

TPS/PLA blend (■) and hybrids with the incorporation of C30B at 1.0% (…), 2.5%

545 

(U), 5.0 ({) 7.5 % (V), and 10.0 (‘) mass%.

ip t

542 

Figure 4: SEM images for the extruded samples; neat 70:30 TPS/PLA blend (a), and for

547 

the hybrids with the incorporation of C30B at 1.0 (b), 2.5 (c), 5.0 (d), 7.5 (e) and 10.0

548 

(f) mass%.

us

cr

546 

Figure 5: Stress versus strain curves for the extruded samples; (a) TPS ( ) and PLA (¯);

550 

(b) neat 70:30 TPS/PLA blend (■) and hybrids with the incorporation of C30B at 1.0%

551 

(…), 2.5% (U), 5.0 ({) 7.5 % (V), and 10.0 (‘) mass%.

an

549 

Figure 6: Stress versus strain curves for the injection-molded samples; neat 70:30

553 

TPS/PLA blend (■) and for the hybrids the incorporation of C30B at 5.0 ({) 7.5 %

554 

(V), and 10.0 (‘) mass%.

d

Figure 7: SEM images for the injection-molded samples; neat 70:30 TPS/PLA blend (a),

te

555 

M

552 

and for the hybrids the incorporation of C30B at 5.0 (b) 7.5 % (c), and 10.0 (d) mass%.

Ac ce p

556  557  558  559 

 

24

 

Page 24 of 32

ip t cr M

an

us

560 

Ac ce p

te

d

561 

562  563  564 

 

565  566  567  568 

Figure 1

569 

25

 

Page 25 of 32

cr

ip t

570 

M

an

us

571 

Ac ce p

te

d

572 

573  574  575  576  577 

Figure 2

26

 

Page 26 of 32

ip t M

an

us

cr

578  579 

Ac ce p

te

d

580  581 

582  583 

584  585 

Figure 3

27

 

Page 27 of 32

ip t cr M

an

us

586 

Ac ce p

te

d

587 

588 

589  28

 

Page 28 of 32

590  591  592 

ip t

Figure 4

an

us

cr

593 

 

Ac ce p

te

d

M

594 

595  596  597  598  599 

 

600  601  602  603 

29

 

Page 29 of 32

604  605 

Figure 5

606 

ip t

607 

M

an

us

cr

608 

609  610 

te

d

611  612 

Ac ce p

613  614  615  616  617  618  619  620  621  622  623 

30

 

Page 30 of 32

Figure 6

cr

ip t

624 

d

M

an

us

625 

Ac ce p

te

626 

627  628 

629 fig. 7

31

 

Page 31 of 32

638 

Highlights

639  640 

• Melt-blending of cornstarch hybrid blends were investigated.

ip t

641 

643 

cr

642 

• Hybrids of both components were prepared at the same contents of the organoclay.

645 

us

644 

• Results indicated the incorporation of the organoclay in both phases.

647 

an

646 

• .

• The organoclay dispersion and pseudoplasticity affected hybrid materials properties.

d

649 

M

648 

te

650  651 

Ac ce p

652  653  654  655  656  657  658  659 

 

660 

33

 

Page 32 of 32

Effect of structure and viscosity of the components on some properties of starch-rich hybrid blends.

Glycerol-plasticized cornstarch and poly(lactic acid) (PLA) were melt-blended alone and at a constant 70:30 (m/m) composition, in the present of an or...
2MB Sizes 0 Downloads 6 Views