You are on page 1of 24

1 Characterization of Tapioca Starch Biopolymer Composites Reinforced with Micro Scale Water

2 Hyacinth Fibers

3 Hairul Abrala*, Maro Hagabean Dalimunthea, Joko Hartonoa, Rice Putra Efendia, Mochamad

4 Asrofia, Eni Sugiartib, SM Sapuanc,d

a
6 Department of Mechanical Engineering, Andalas University, 25163 Padang, Sumatera Barat,

7 Indonesia

b
8 Laboratory of High Temperature Coating, Research Center for Physics Indonesian Institute of

9 Sciences (LIPI) Serpong, Indonesia

c
10 Department of Mechanical and Manufacturing Engineering, Faculty of Engineering, Universiti

11 Putra Malaysia, 43400 UPM Serdang, Selangor, Malaysia

d
12 Laboratory of Biocomposite Technology, Institute of Tropical Foresty and Forest Products

13 (INTROP) Universiti Putra Malaysia, 43400 UPM Serdang, Selangor, Malaysia

14 * Corresponding author. Tel.: +62-8126720355

15

16 E-mail address: abral@ft.unand.ac.id

17

18

19

20

21
22 Abstract

23 This paper informs characterization of micro scale pulp water hyacinth fibers (WHF) filled tapioca

24 starch biopolymer (TSB) composites. The volume fraction of WHF in TSB matrix was varied of

25 1, 3, 5, 10% respectively. Casting method was used for making the film samples. Some

26 characterizations of the present studied samples observed were mechanical properties, thermal and

27 moisture resistance, Fourier transform infrared spectroscopy (FTIR) and scanning electron

28 microscopy (SEM). The results showed that SEM fracture surface of biocomposite for tensile

29 sample displayed good fibers distribution in the matrix, and interface bonding between WHF and

30 TSB. 10% fibers content biocomposite delivered the highest value of tensile strength (TS) of 6.68

31 MPa, and tensile modulus (TM) of 210.95 MPa, however, had the lowest fracture strain of 7.30%.

32 In this case, there was 549% improvement of TS and 973% of TM in comparison to TSB.

33 Biocomposite with 10% WHF content showed also the highest thermal resistance and the lowest

34 moisture absorption.

35

36 Keywords: Starch biopolymer; Pulp water hyacinth fibers; Biocomposites; Mechanical properties;

37 Thermal and moisture resistance.

38

39

40

41

42

43

44
45 1. Introduction

46 In last two decades, attention has been given to the development of starch based thermoplastics

47 as a way of reducing non-degradable synthetic polymer use [1]. These thermoplastics are

48 environmentally friendly, cheap to produce from widely available raw materials and biodegradable

49 [2]. The cassava or tapioca plant is one extensively researched source of starch because it grows

50 abundantly in tropical countries like Indonesia. This plant contains a high amount of water along

51 with starch and celluloses fibers [3]. However, compared with conventional synthetic

52 thermoplastics, tapioca starch thermoplastics are brittle, has poor mechanical properties, and is

53 water sensitive [4]. Extensive research has been conducted to attempt to enhance these properties

54 in recent years [3,57]. Brittleness has been reduced by mixing in a plasticizer such as glycerol

55 [8]. To cope with low tensile strength and low water resistance, fillers obtained from natural fibers

56 have been added to starch based biopolymers to produce biocomposites [5]. Chang et al. reported

57 an improvement of mechanical properties, thermal and moisture resistance of biocomposites

58 consisting of cellulose nanoparticles from natural sources [9].

59 Generally, the natural fiber has chemical compositions consisting of celluloses,

60 hemicelluloses, and lignin [10]. Strength of the fibers depends on the cellulose content [11]. If this

61 is high the mechanical properties are better [12]. One possible natural fibers to be used as a

62 reinforcement in starch based composites is WHF with a high cellulose content of 3050% [13].

63 Water hyacinths reproduce prolifically in the tropics, creating problems in the environment as they

64 block irrigation channels. They are easily harvested as they float on the water surface. Commercial

65 use of the plant is still limited, however. Therefore, considering its high cellulose content it could

66 be used as a reinforcement material for biocomposites [14].


67 Size of natural fiber affect the mechanical properties of a starch based thermoplastic [4].

68 Decrease in fiber diameter improves mechanical properties and resistance to moisture absorption

69 [4]. Intensive studies have been conducted to improve the surface area ratio between fibers and the

70 matrix interface by reducing the diameter size of the fibers [15]. Nanocrystallite of ramie cellulose

71 in a bioplastic of wheat starch has been shown to improve the properties of the biocomposite [5].

72 Such cellulose nano-crystals from plant sources can be produced mechanically or chemically [16

73 20]. However, producing nano sized dimension fibers is expensive and can use environmentally

74 damaging chemicals. It is economically advantageous if micro-sized fibers from abundantly

75 available plants could be utilized as the reinforcing material for starch based thermoplastics.

76 Therefore it is necessary to develop methods to lower the cost and optimize the process

77 performances of bio-based products with efficient packaging materials and techniques providing

78 surety in terms of safety and best quality of the food items [21].

79 Some previous work using WHF has been published [14,2225]. However, as far as we are

80 aware, there are no previous studies investigating the microscale effects of WHF on the physical

81 and mechanical properties of starch tapioca based thermoplastic. Therefore, this research was

82 designed to study the properties of this combination. Mechanical properties, thermal resistance,

83 moisture absorption, FTIR, XRD, SEM fracture surface of the studied samples were measured.

84 Results of this investigation could form a basis for further research to provide better understanding

85 of the properties of this biocomposite and the potential development of a commercially attractive

86 solution to a pressing environmental problem.

87

88 2. Materials and Experimental

89 2.1. Extraction of WHF


90 WHF was extracted from freshwater hyacinths growing in a local swamp in 50 Kota

91 district, Indonesia, cut into 1 cm lengths and cleaned with fresh water. Then, it was placed in a

92 plastic covered shelter under sunlight for 5 days to dry. The dried WHF samples were mixed with

93 a 25% NaOH solution and placed in a digester for 6 h at about 130 oC, 2 bar. The pulp was

94 neutralized by water until pH 7, blended by a mixer for 10 min then screened by using a printing

95 screen with T61 mesh. The wet pulp was then solar dried.

96

97 2.2. Measuring chemical compositions

98 The cellulose, lignin, and hemicellulose composition of WHF tested. Cellulose and

99 hemicellulose content was measured using the Technical Association of the Pulp and Paper

100 Industry (TAPPI) standard T9M-54, and the lignin content was tested by using TAPPI T13M-54.

101 For the sake of conciseness, these procedures are not explained in this paper.

102

103 2.3. Preparation of biocomposites

104 The tapioca powder used for this research was a commercial product with name Cap Pak

105 Tani, bought in a local market. The content of water, amylose, and amylopectin in the powder

106 was 19, 15, and 85%, respectively. Tapioca starch bioplastic (TSB) was processed by mixing 1g

107 WHF with 140 ml of water. 10 g tapioca powder and 3 ml of glycerol were added to the mixture

108 which was then mixed with a (Daihan HG 15D-Set A) homogenizer for 5 min at 5000 rpm. The

109 solution was gelatinized using a magnetic stirrer at 150oC, 500 rpm for 18 min. Finally, the

110 gelatinizing biocomposite was poured into a rectangular glass mold and placed in a Universal

111 drying oven (Memmert UN-55) at 50 oC for 20 h. The studied samples were released from the

112 mold after 24 h at room temperature.


113

114 2.4. Thermogravimetric Analysis

115 Thermal properties of 7 mg samples were measured by using TGA/DSC1 equipment from

116 Mettler Toledo. The heating rate was 10 C per min from 25 C to 600 C.

117

118 2.5. Moisture absorption

119 Films with a dimension of 1 x 1.5 cm were dried in an oven (Universal Oven Memmert

120 UN-55) until constant weight then stored in a closed chamber with 99% relative humidity (RH)

121 for 10 h. The samples were taken out and weighed on a precision balance to the nearest 0.1 mg

122 (Kenko). Percentage of moisture absorption in the sample was calculated from the difference

123 between dried weight and wet weight divided by dried weight [23].

124

125 2.6. Tensile testing

126 A Com-Ten 95T Series testing machine was used for measuring the tensile strength of all

127 samples at 3 mm/min and room temperature 14 days after samples were released from the mold.

128 Five repeats for each fiber percentage were conducted using ASTM D638 type I Standard test [26].

129 Maximum force on the sample was divided by average cross section area to obtain tensile strength.

130 In order to obtain average cross section area accurately, thickness and width were measured at ten

131 different location points using a dial indicator micrometer to 1m accuracy. Tensile modulus was

132 automatically derived by the tensile equipment as the slope of the linear region of the stress-strain

133 curve. The strain of a sample was measured as the fracture strain.

134

135 2.7. XRD testing


136 XRD testing was performed by using PANalytical Xpert PRO at 25 oC, 40 kV and 30 mA.

137 The samples were scanned from 2 = 3 o to 40o. Crystal size (L) was calculated using Eq. 1:

138 L = k. / (B cos 2) (1)

139 where the values of k and were 0.89, 1.54, respectively. B was measured from full width at half

140 maximum of XRD pattern peak (FWHM) multiplied by /180. The percentage of crystallinity

141 index (CI) was measured by using Eq. 2 [27]:

142 CI (%)= (I200- Iam)/ I200 100 (2)

143 where I200 is the intensity of the peak corresponding to cellulose I, and Iam is the intensity of the

144 peak of the amorphous fraction.

145

146 2.8. FTIR Characterization

147 FTIR characterization was performed by using The PerkinElmer Frontier. The aim of

148 FTIR characterization was to determine the functional group of the TSB and the TSB/WHF

149 composite. The sheet film from the dried samples was scanned at the frequency range of 4000

150 600 cm-1.

151

152 2.9. SEM observation

153 Observation of the fracture surface of all tested samples was performed by using SEM

154 (Hitachi 3400 N series) at 15 kV.

155

156 3. Results and discussion

157 3.1. Chemical compositions of WHF


158 Chemical compositions of natural fibers depend on growing location, age, temperature, and

159 humidity [23]. The WHF studied came from local sources, and its chemical composition is

160 reported in Table 1. The composition of the WHF pulp was changed after processing. Before

161 treatment, the WHF pulp had high lignin (4.1%) and hemicellulose (20.6%) content which

162 decreased after digesting and blending for 10 min to 3.9% and 3.5% respectively. Meanwhile, the

163 cellulose content became more dominant in the treated WHF sample being 67%.

164

165 3.2. Thermogravimetric Analysis (TG) and derivative of TGA (DTG)

166 Thermal degradation and decomposition of the present samples were observed to measure

167 the change of thermal properties due to differences in chemical composition. Fig. 1a shows the TG

168 curve in thermal weight loss, and Fig. 1b is the DTG curve of samples as a function of rising

169 temperature. TG curve shows the three main sections that have also been observed in previous

170 studies [10,11]. Up until 100 oC, the sample displays a slight loss of weight due to evaporation of

171 water [3,28]. This small amount of dehydration up until about 100 0C was also confirmed by the

172 DTG curve (Fig. 1b). As the temperature reaches the range of 250-3400C, the TG curve indicates

173 maximum decomposition of the sample probably attributed to oxidation of the partially

174 decomposed starch as well as ber degradation [29,30]. Above 300 0C, the highest rate of

175 decomposition is observed for the TSB. The rate of mass loss for the biocomposite samples

176 containing WHF was lower than for the pure TSB. This was also observed in a previous study

177 [31]. The maximum thermal resistance was measured on the 10% WHF biocomposite owning the

178 highest CI of all tested samples as evidenced in the section dealing with XRD diffraction in this

179 present paper. Previous work reported that the greater crystalline structure led to a high resistance

180 toward heat and an increase in maximum temperature for thermal degradation [32]. This
181 phenomenon was similar to a previous report [33]. Another factor that could also improve the

182 thermal stability of the biocomposites is good interface bonding between TSB and WHF that led

183 to strong hydrogen bonding, thus reducing the mass loss in the sample [32,34,35]. This is

184 consistent with FTIR curve (Fig. 2) that has shown shifting the wave number of biocomposite as

185 evidence of good interface bonding.

186

187 3.3. FTIR

188 Fig. 2 exhibits the FTIR spectra of the studied samples along with an untreated WHF

189 control. This is similar to the FTIR spectrum published previously [23]. After processing in the

190 digester and blending 10 min the functional group of the WHF had a changed wave number

191 position. Untreated WHF exhibits a broad spectrum and O-H stretching at wave number 3344.5

192 cm-1. After boiling with 25% NaOH and blending 10 min, the O-H functional group shifted to

193 3283.8 cm-1. Previous research reported that decrease in the wave number of O-H stretching can

194 be attributed to lower force from electron delocalization [36]. C-H stretching data corresponds to

195 the cellulose and hemicellulose component [37].

196 After the treatment, these wave numbers were shifted from 2910.4 cm-1 to 2920.9 cm-1.

197 The absorption area at 1253.9 cm-1 corresponds to a C-O-C functional group. After chemical and

198 mechanical treatment, the C-O-C functional group disappeared (Fig. 2b). Peaks in the wavenumber

199 from 1037.6 cm-1 to 1017.3 cm-1 corresponding to the C-O functional group related to the cellulose

200 component shifted after chemical and mechanical treatment. The continuing presence of C-O

201 functional groups proves that the cellulose component was not eliminated by alkali treatment [37].

202 Meanwhile, TSB has O-H functional group at wavenumber 3293 cm-1 (Fig. 2c). After

203 addition of 1, 3, 5, 10 % WHF to TSB, the O-H band shows a significant shifting. For example,
204 the O-H functional group of 1% WHF biocomposite shifted to 3295 cm-1 as shown in Fig. 2d. This

205 shift is strong evidence for the formation of hydrogen bonds between WHF and the TSB matrix

206 [38][39]. The O-H functional groups were reduced as the amount of WHF fiber increased leading

207 to a decrease in the ability of the O-H functional groups to bind with water resulting in a more

208 hydrophobic biocomposite.

209

210 3.4. X-ray diffraction

211 X-ray diffraction was used to measure changes in the crystal. Fig. 3 shows XRD patterns

212 for TSB (a), WHF (b), and WHF-TSB composites (c-d), meanwhile Table 2 shows crystallinity

213 index (CI) and crystal system of all tested samples. The XRD pattern of WHF exhibits a sharp

214 peak at a high scattering angle that indicates a high crystallinity index (CI) [32]. The CI of WHF

215 peaked at 59.56% (Fig. 3). TSB shows a broad diffraction pattern and a peak at a low scattering angle

216 which can be attributed to high amorphous content and small crystal size [7]. CI of TSB was

217 10.9%. This explains why increases in WHF content in the TSB results in improvement of the CI

218 in the resulting biocomposite as shown in Table 2. The highest CI was measured in the 10% WHF

219 biocomposite. A high CI in a biocomposite is correlated with improved tensile properties so it is

220 expected that addition of WHF results in an enhancement of the TSB based biocomposite tensile

221 properties [32,40].

222

223 3.5. Moisture absorption

224 Fig. 4 shows average moisture absorption (MA) of samples with different fiber content as

225 a function of time. There were five repeats for each fiber content tested. At the beginning, MA for

226 all samples is similar and high due to the large difference in relative humidity between the sample
227 and humid environment in the chamber. After time in the chamber, the rate of MA decreases

228 toward a saturation point. It was noted that achievement of the saturation point for TSB took longer

229 than for the biocomposite. Furthermore, MA of the biocomposite was lower than TSB and tended

230 to decrease as fiber content increased. This confirms that WHF used in the present study was less

231 hygroscopic than the starch due to the higher degree of molecular order resulting in improved

232 barrier properties [41]. At 10 h in the humid chamber, the lowest MA value was 26.8% for the

233 10% fibers biocomposite and the highest 32.6% for the TSB. Increasing WHF content in the TSB

234 resulted in a more hydrophobic biocomposite presumably due to strong hydrogen bonding between

235 WHF and TSB resulting in lower diffusivity of water molecules into the sample [8,42] This is

236 reinforced by observation of the SEM fracture surface (Fig.7d).

237

238 3.6. Mechanical Properties

239 Fig. 5 shows the tensile strength (TS) versus strain curve of individual samples with

240 differing WHF content. As predicted from the CI results, pure TSB displayed a low TS and tensile

241 modulus (TM), and the highest fracture strain. As fiber content increased, TS and TM values

242 increased and fracture strain decreased indicating that fiber strengthens the biocomposite but

243 causes to become more brittle. Average values of TS, TM, and fracture strain versus the various

244 WHF content in TSB are displayed with error bars in Fig. 6a, 6b, and 6c respectively. These

245 confirm that both the TS and TM improved as fiber content increased but fracture strain decreased.

246 Average highest values for TS and TM of the 10% WHF biocomposite were 6.7 and 210.9

247 MPa respectively. While this biocomposite was the strongest it also had the lowest fracture strain

248 of 3.7%. Hence, addition of 10% WHF increased TS by 549%, and TM by 973% in comparison

249 to TSB. Similar improvement of mechanical properties with increases in fiber content has been
250 observed in previous studies [11,43,44]. In case of the present data, improvement of these

251 mechanical properties may be due to the fact that the WHF was well distributed in the TSB so

252 reinforced the biocomposite against the applied external tensile load. Good interface bonding

253 between WHF and TSB as shown in Fig. 7d may also contribute to increases in TS and TM values

254 of the biocomposites.

255

256 3.7. SEM photograph

257 The SEM photograph of pure TSB (Fig. 7a) shows a smooth fracture surface. In comparison, the

258 SEM photographs of biocomposites with various WHF content ( Fig. 7b-c) show rougher fracture

259 surfaces due to the fiber content. As shown in Fig. 7b-c, the fibers in the TSB matrix were

260 uniformly distributed, and there was no agglomeration of the fibers observed. Agglomeration

261 could reduce the mechanical properties of the biocomposite. The higher amount of WHF in TSB

262 provides the higher possibility of fibers reinforcing the biocomposite to withstand external load,

263 creating increases in the mechanical properties of the sample. Furthermore, the enlarged fracture

264 surface in Fig. 7d shows no gap between fibers and the matrix indicating good interfacial bonding.

265 This is to be expected due to the similar polysaccharide structures of cellulose and starch [9].

266 Strong adhesion bonding effectively transfers applied stress from the matrix to the WHF improving

267 mechanical properties of the biocomposite.

268

269 4. Conclusions

270 The present study has highlighted that WHF, a major environmental hazard, could be put

271 to good use in biocomposite production. Chemical and mechanical treatment reduced

272 hemicellulose and lignin content in WHF while increasing cellulose content. 1-10% fractions of
273 WHF in TSB resulted in an improved material as evidenced by SEM photographs which indicated

274 uniform distribution and strong good interfacial bonding. As the amount of WHF in the TSB

275 matrix increased so did tensile properties, and tensile modulus, thermal and moisture resistance.

276 However, the strain of the biocomposite decreased. 10% WHF biocomposite delivered the best

277 performance (improvement of 549% TS, 973% TM), thermal, and moisture resistance and has

278 potential for commercial applications in the future.

279

280 Acknowledgements

281 Acknowledgment is addressed to Department of Mechanical Engineering, Andalas

282 University for supporting this research in year of 2016.

283

284 References:

285
286 [1] B.N. Avrous L, Biocomposites based on plasticized starch: thermal and mechanical
287 behaviour, Carbohydr. Polym. 56 (2004) 111122.
288 [2] A. Dufresne, Plasticized Starch/Tunicin Whiskers Nanocomposites. 1. Structural Analysis,
289 Macromolecules. 33 (2000) 83448353.
290 [3] E. de M. Teixeira, D. Pasquini, A.A.S. Curvelo, E. Corradini, M.N. Belgacem, A.
291 Dufresne, Cassava bagasse cellulose nanofibrils reinforced thermoplastic cassava starch,
292 Carbohydr. Polym. (2009).
293 [4] X. Cao, Y. Chen, P.R. Chang, A.D. Muir, G. Falk, Starch-based nanocomposites
294 reinforced with flax cellulose nanocrystals, Express Polym. Lett. 2 (2008) 502510.
295 [5] Y. Lu, L. Weng, X. Cao, Morphological, thermal and mechanical properties of ramie
296 crystallitesreinforced plasticized starch biocomposites, Carbohydr. Polym. 63 (2006)
297 198204.
298 [6] K. Majdzadeh-Ardakani, A.H. Navarchian, F. Sadeghi, Optimization of mechanical
299 properties of thermoplastic starch/clay nanocomposites, Carbohydr. Polym. 79 (2010)
300 547554.
301 [7] Q. Sun, T. Xi, Y. Li, L. Xiong, Characterization of Corn Starch Films Reinforced with
302 CaCO3 Nanoparticles, PLoS One. 9 (2014) 16.
303 [8] A. Dufresne, Cellulose Microfibrils from Potato Tuber Cells: Processing and
304 Characterization of Starch Cellulose Microfibril Composites, Polymer (Guildf). 76
305 (2000) 20802092.
306 [9] P.R. Chang, R. Jian, P. Zheng, J. Yu, X. Ma, Preparation and properties of glycerol
307 plasticized-starch (GPS)/cellulose nanoparticle (CN) composites, Carbohydr. Polym. 79
308 (2010) 301305.
309 [10] W. Chen, H. Yu, Y. Liu, P. Chen, M. Zhang, Y. Hai, Individualization of cellulose
310 nanofibers from wood using high-intensity ultrasonication combined with chemical
311 pretreatments, Carbohydr. Polym. 83 (2011) 18041811.
312 [11] A. Kaushik, M. Singh, G. Verma, Green nanocomposites based on thermoplastic starch
313 and steam exploded cellulose nanofibrils from wheat straw, Carbohydr. Polym. 82 (2010)
314 337345.
315 [12] C. Miao, W.Y. Hamad, Cellulose reinforced polymer composites and nanocomposites: A
316 critical review, Cellulose. 20 (2013) 22212262.
317 [13] A. Bhattacharya, P. Kumar, Water hyacinth as a potential biofuel crop, Electron. J.
318 Enviromental, Agricltural, Food Chem. 9 (2010) 112122.
319 [14] M. Thiripura Sundari, A. Ramesh, Isolation and characterization of cellulose nanofibers
320 from the aquatic weed water hyacinth - Eichhornia crassipes, Carbohydr. Polym. 87
321 (2012) 17011705.
322 [15] M. Ahmed, S. Azizi, F. Alloin, A. Dufresne, Review of Recent Research into Cellulosic
323 Whiskers, Their Properties and Their Application in Nanocomposite Field,
324 Biomacromolecules. 6 (2005) 612626.
325 [16] Q. Cheng, S. Wang, Q. Han, Novel process for isolating fibrils from cellulose fibers by
326 high-intensity ultrasonication. II. fibril characterization, J. Appl. Polym. Sci. 115 (2010)
327 27562762.
328 [17] A. Chakraborty, M. Sain, Cellulose microfibrils: A novel method of preparation using
329 high shear refining and cryocrushing, 59 (2005) 102107.
330 [18] F. Fahma, S. Iwamoto, N. Hori, T. Iwata, A. Takemura, Isolation, preparation, and
331 characterization of nanofibers from oil palm empty-fruit-bunch (OPEFB), Cellulose. 17
332 (2010) 977985.
333 [19] Y. Chen, C. Liu, P.R. Chang, X. Cao, D.P. Anderson, Bionanocomposites based on pea
334 starch and cellulose nanowhiskers hydrolyzed from pea hull fibre: Effect of hydrolysis
335 time, Carbohydr. Polym. 76 (2009) 607615.
336 [20] M. Rahimi, R. Behrooz, Effect of Cellulose Characteristic and Hydrolyze Conditions on
337 Morphology and Size of Nanocrystal Cellulose Extracted from Wheat Straw, Int. J.
338 Polym. Mater. 60 (2011) 529541.
339 [21] B. Khan, M. Bilal Khan Niazi, G. Samin, Z. Jahan, Thermoplastic Starch: A Possible
340 Biodegradable Food Packaging MaterialA Review, J. Food Process Eng. 40 (2017).
341 [22] A.G. Supri, H. Ismail, The Effect of Isophorone Diisocyanate-Polyhydroxyl Groups
342 Modified Water Hyacinth Fibers (Eichhornia Crassiper) on Properties of Low Density
343 Polyethylene/Acrylonitrile Butadiene Styrene (LDPE/ABS) Composites, Polym. Plast.
344 Technol. Eng. 50 (2011) 113120.
345 [23] H. Abral, D. Kadriadi, A. Rodianus, P. Mastariyanto, Ilhamdi, S. Arief, S.M. Sapuan,
346 M.R. Ishak, Mechanical properties of water hyacinth fibers - polyester composites before
347 and after immersion in water, Mater. Des. 58 (2014) 125129.
348 [24] A.G. Supri, S.J. Tan, H. Ismail, P.L. Teh, Effect of Poly(methyl Methacrylate) Modified
349 Water Hyacinth Fiber on Properties of Low Density Polyethylene/Natural Rubber/Water
350 Hyacinth Fiber Composites, Polym. Plast. Technol. Eng. 50 (2011) 898906.
351 [25] H. Abral, H. Putra, S.M. Sapuan, M.R. Ishak, Effect of Alkalization on Mechanical
352 Properties of Water Hyacinth Fibers-Unsaturated Polyester Composites, Polym. Plast.
353 Technol. Eng. 52 (2013) 446451.
354 [26] ASTM-D638, Standard test method for tensile properties of plastic, Am. Soc. Test. Mater.
355 (2002).
356 [27] L. Segal, J.J. Creely, A.E. Martin, M. Conrad, Empirical Method for Estimating the
357 Degree of Crystallinity of Native Cellulose Using the X-Ray Diffractometer, Text. Res. J.
358 (1958) 786794.
359 [28] M.G. Lomel-ramrez, S.G. Kestur, R. Manrquez-gonzlez, S. Iwakiri, G. Bolzon, D.
360 Muniz, Bio-composites of cassava starch-green coconut fiber: Part II Structure and
361 properties, Carbohydr. Polym. 102 (2014) 576583.
362 [29] L. Ribba, A. Dufresne, M.I. Aranguren, N.L. Garc, S. Goyanes, Physico-Mechanical
363 Properties of Biodegradable Starch Nanocomposites, 294 (2009) 169177.
364 [30] S. Karimi, A. Dufresne, Biodegradable starch-based composites: effect of micro and
365 nanoreinforcements on composite properties, J Mater Sci. 49 (2014) 45134521.
366 [31] M. Jonoobi, A. Khazaeian, P.M. Tahir, S.S. Azry, K. Oksman, Characteristics of cellulose
367 nanofibers isolated from rubberwood and empty fruit bunches of oil palm using chemo-
368 mechanical process, 18 (2011) 10851095.
369 [32] M. Jonoobi, J. Harun, A. Shakeri, M. Misra, K. Oksman, Chemical composition,
370 crystallinity, and thermal degradation of bleached and unbleached kenaf bast, Bio Resour.
371 4 (2009) 626639.
372 [33] S. Lee, D.J. Mohan, I. Kang, G. Doh, S. Lee, S.O. Han, Nanocellulose Reinforced PVA
373 Composite Films: Effects of Acid Treatment and Filler Loading, Fibers Polym. 10 (2009)
374 7782.
375 [34] H.M. Ng, L.T. Sin, T.T. Tee, S.T. Bee, D. Hui, C.Y. Low, A.R. Rahmat, Extraction of
376 cellulose nanocrystals from plant sources for application as reinforcing agent in polymers,
377 Compos. Part B Eng. 75 (2015) 176200.
378 [35] M. Kumar, S. Mohanty, S.K. Nayak, M. Rahail Parvaiz, Effect of glycidyl methacrylate
379 (GMA) on the thermal, mechanical and morphological property of biodegradable
380 PLA/PBAT blend and its nanocomposites, Bioresour. Technol. 101 (2010) 84068415.
381 [36] K.M. Dean, M.D. Do, E. Petinakis, L. Yu, Key interactions in biodegradable
382 thermoplastic starch / poly ( vinyl alcohol )/ montmorillonite micro- and nanocomposites,
383 Compos. Sci. Technol. 68 (2008) 14531462.
384 [37] A. Alemdar, M. Sain, Isolation and characterization of nanofibers from agricultural
385 residues Wheat straw and soy hulls, Bioresour. Technol. 99 (2008) 16641671.
386 [38] K. Kaewtatip, J. Thongmee, Studies on the structure and properties of thermoplastic
387 starch/luffa fiber composites, Mater. Des. 40 (2012) 314318.
388 [39] H. Abral, G.J. Putra, M. Asrofi, J. Park, H. Kim, Effect of Vibration Duration of High
389 Ultrasound Applied to Bio- composite While Gelatinized on its Properties, Ultrason. -
390 Sonochemistry. (2017).
391 [40] J. Sahari, S.M. Sapuan, E.S. Zainudin, M.A. Maleque, Mechanical and thermal properties
392 of environmentally friendly composites derived from sugar palm tree, Mater. Des. 49
393 (2013) 285289.
394 [41] A. Kaushik, M. Singh, G. Verma, Green nanocomposites based on thermoplastic starch
395 and steam exploded cellulose nanofibrils from wheat straw, Carbohydr. Polym. 82 (2010)
396 337345.
397 [42] M.S. Sreekala, K. Goda, P. V Devi, Sorption characteristics of water, oil and diesel in
398 cellulose nanofiber reinforced corn starch resin/ramie fabric composites, Compos.
399 Interfaces. 15 (2008) 281299.
400 [43] Y. Dong, A. Ghataura, H. Takagi, H.J. Haroosh, A.N. Nakagaito, K.T. Lau, Polylactic
401 acid (PLA) biocomposites reinforced with coir fibres: Evaluation of mechanical
402 performance and multifunctional properties, Compos. Part A Appl. Sci. Manuf. 63 (2014)
403 7684.
404 [44] A.N. Frone, D.M. Panaitescu, D. Donescu, C.I. Spataru, C. Radovici, R. Trusca, R.
405 Somoghi, Composites With Cellulose Nanofibers Obtained By Ultrasonication,
406 Bioresources. 6 (2011) 487512.
407
(a)

(b)

Fig. 1. Curve for thermogravimetric TG (a), and derivate DTG (b) for all tested samples.
Fig. 2. FTIR spectra for the studied samples of (a) untreated 100% WHF (b) treated 100% WHF

(c) TSB (d) 1% WHF (e) 3% WHF (f) 5% WHF (g) 10% WHF in TSB matrix.
Fig. 3. XRD patterns of six studied samples

Fig. 4. Performance of moisture absorption for all studied samples


Fig. 5. Tensile strength of individual sample as a function of WHF in TSB matrix
(a) (b)

Fig. 6. Effect WHF fraction in TSB matrix on mechanical properties of (a) tensile strength (b)

tensile modulus (c) fracture strain.


Fig. 7. SEM photograph of (a) TSB matrix (b) 1% WHF in TSB matrix (c) 10% WHF in TSB matrix

(d) good adhesion bonding of WHF in TSB matrix


Table 1. Chemical compositions in the untreated and treated WHF pulp

Lignin (%) Hemicellulose (%) Cellulose (%)

Untreated 4.1 20.6 42.8

Treated WHF Pulp 3.9 3.5 67.0

Table 2. Crystallinity index, crystal system, and compound of all tested samples

Sample L () d () Crystallinity Crystal Compound

index (%) system

WHF 57.25 3.93 59.56 Monoclinic Cellulose-I

TSB 68.99 4.03 10.90 Monoclinic Cellulose

1% WHF in TSB 60.69 4.00 11.98 Anorthic Cellulose-I

3% WHF in TSB 60.18 4.00 14.26 Monoclinic Cellulose-I

5% WHF in TSB 59.92 3.98 12.08 Monoclinic Cellulose-I

10% WHF in TSB 59.43 3.95 17.36 Monoclinic Cellulose-I

You might also like