You are on page 1of 26

Accepted Manuscript

Title: Polymer blend of PLA/PHBV based bionanocomposites


reinforced with nanocrystalline cellulose for potential
application as packaging material

Author: Y.K. Dasan A.H. Bhat A. Faiz

PII: S0144-8617(16)31277-2
DOI: http://dx.doi.org/doi:10.1016/j.carbpol.2016.11.012
Reference: CARP 11723

To appear in:

Received date: 10-8-2016


Revised date: 26-10-2016
Accepted date: 3-11-2016

Please cite this article as: Dasan, YK., Bhat, AH., & Faiz, A., Polymer
blend of PLA/PHBV based bionanocomposites reinforced with nanocrystalline
cellulose for potential application as packaging material.Carbohydrate Polymers
http://dx.doi.org/10.1016/j.carbpol.2016.11.012

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.
Polymer blend of PLA/PHBV based
bionanocomposites reinforced with nanocrystalline
cellulose for potential application as packaging
material

Y.K. Dasan1, a), A.H.Bhat2, a), A. Faiz3, b)

1,2
Department of Fundamental and Applied Sciences
, 3Department of Chemical Engineering,
Universiti Teknologi PETRONAS, 32610, Bandar Seri Iskandar, Perak,
Malaysia

Corresponding authors: 1, a) yaleenikannadasan@gmail.com,


2, a)
aamir.bhat@petronas.com.my,
b)
faizahmad@petronas.com.my

1
Highlights
 Nanocrystalline cellulose extracted from oil palm empty fruit bunch fiber through

Sulphuric acid hydrolysis process.

 Surface functionalization of nanocrystalline cellulose using TEMPO-mediated oxidation.

 Nanocrystalline cellulose reinforced PLA/PHBV bionanocomposites films were prepared

by solution casting techniques.

 Improvement of oxygen barrier was observed in nanocrystalline cellulose reinforced

bionanocomposite films.

 Nanocrystalline cellulose increased the mechanical properties of bionanocomposite film.

2
Abstract. The current research discusses the development of poly (lactic acid) (PLA) and poly-(3-hydroxybutyrate-

co-3-hydroxyvalerate) (PHBV) reinforced nanocrystalline cellulose bionanocomposites. The nanocrystalline

cellulose was derived from waste oil palm empty fruit bunch fiber by acid hydrolysis process. The resulting

nanocrystalline cellulose suspension was then surface functionalized by TEMPO-mediated oxidation and solvent

exchange process. Furthermore, the PLA/PHBV/nanocrystalline cellulose bionanocomposites were produced by

solvent casting method. The effect of the addition of nanocrystalline cellulose on structural, morphology,

mechanical and barrier properties of bionanocomposites was investigated. The results revealed that the developed

bionanocomposites showed improved mechanical properties and decrease in oxygen permeability rate. Therefore,

the developed bio-based composite incorporated with an optimal composition of nanocrystalline cellulose exhibits

properties as compared to the polymer blend.

Keywords. Nanocrystalline cellulose; bionanocomposites; solution casting; flexural; morphology; permeability

3
1.0 INTRODUCTION
In recent years, the potentials of bionanocomposites in the development of plastic materials have attracted

increasing investment. The use of natural fibers in composite making process has reduced the production

cost while maintaining the properties (Kalia et al., 2011; Verma, Gope, Maheshwari, & Sharma). In

addition, the use thermoplastic biopolymers have added more value to this bionanocomposite based

products as it is an environmentally friendly material. Bionanocomposite has gained tremendous attention

due to its renowned properties such as biodegradable, biocompatible and it gives zero impact to the

environment.

Biopolymers are great in demand as they reduce the dependency on the petroleum-based polymer, reduce

the accumulation of persistent plastic waste and lower the CO 2 emission in the environment (Avérous &

Pollet, 2012). Poly (lactic acid) (PLA) and Poly-(3-hydroxybutyrate-co-3-hydroxyvalerate) (PHBV) are the

most commonly used biopolymers in the market. PLA produced by chemical synthesis from bio-derived

monomers while PHBV produced by a microorganism or genetically modified bacteria (Jamshidian,

Tehrany, Imran, Jacquot, & Desobry, 2010). Blending polymer is a simple technique to enhance the

property of pure polymer include providing new materials with desired properties at the low price, quick

formulation changes, plant flexibility and high productivity and reduction of the number of grades that need

to be manufactured and stored (Ploypetchara, Suppakul, Atong, & Pechyen, 2014).

In order to produce fully renewable and biodegradable nanocomposites, both the polymer matrix and the

nano reinforcement have to be derived from renewable resources. Generally, synthetic fibers such as

carbon, glass, and aramid have been applied as a fillers or reinforcement to improve the desired properties

of the polymer or simply reduce the cost (Saba, Tahir, & Jawaid, 2014). However, the severe health and

ecological problems created by these synthetic fibers during the manufacturing of their corresponding

composites have impelled the necessity to look for new alternatives which could replace the traditional

filler materials. This renewed the interest of researchers in natural fibers including orange peels, grape

pomace, coffee grounds, and turmeric waste which could be used as a reinforcements or fillers in the

composites (Iyer, Zhang, & Torkelson, 2016; Ramamoorthy, Skrifvars, & Persson, 2015). Cellulose is one

of the major component of natural fibers that are generally used as an organic filler material for the

composite preparation due to their high levels of strength and stiffness per unit weight and it is the building

material of long fibrous cells (Ibrahim & El-Zawawy, 2015; Jonoobi et al., 2015). The use of natural

4
cellulose is becoming vital due to its availability, low cost, low density, low abrasiveness, nontoxicity,

good mechanical property, biocompatibility, and biodegradability (dos Santos, Iulianelli, & Tavares, 2016;

Kalia et al., 2011; Thakur, Thakur, & Prasanth, 2014). Wood is the most commercially used natural

resource containing cellulose. However, there are many other non-wood plant fibers used a source of

cellulose such as hemp, flax, jute, ramie and cotton. Besides that, bioresidue from industries, agricultural

waste and cellulose rich municipal solid waste materials (MSW) such as oil palm empty fruit bunch fiber

(OPEFB), corrugated cardboard (CB), and waste papers depict another cellulose source possessing great

potential for cellulose derivatives (Iyer, Flores, & Torkelson, 2015; Jonoobi et al., 2015). Furthermore, the

incorporation of nanofillers such as cellulose nanocrystals is believed to be an efficient strategy to tailor the

properties of the polymeric material in the production of the nanocomposites (Martínez-Sanz et al., 2014).

What makes the nanofillers unique is their high aspect ratio which plays a structural role in which they act

as reinforcement to improve the mechanical and barrier properties of the matrix (Othman, 2014). Cellulose

nanocrystals are usually produced by applying an acid hydrolysis process in which it leads to the

preferential digestion of cellulose amorphous region (Martínez-Sanz et al., 2014). Sulfuric acid (H2SO4) is

the most extensively used strong acid for hydrolysis of cellulose fibers. Sulfuric acid reacts with the surface

hydroxyl groups via an esterification process allowing the grafting of anionic sulfate ester groups which is

the major reason for using it as a hydrolyzing agent. In addition, the presence of these negatively charged

groups promotes the formation of a negative electrostatic layer covering the nanocrystals and aid to their

dispersion in water (Dufresne, 2013). There are other acids besides H2SO4 that can be used for the

hydrolysis of cellulose fibers into nanocrystalline cellulose, for example, hydrochloric acid solution,

phosphoric acid solution, hydrobromic acid solution, and a mixture of acetic acid and nitric acid solution in

the ratio of 10:1. However, sulfuric acid gives the highest crystallinity index followed by phosphoric acid.

Nanocrystalline cellulose prepared by hydrochloric acid and the mixed acetic and nitric acid solution

possess the lowest crystallinity index due to their higher tendency to promote the breakage of the hydrogen

bonds in crystalline regions of cellulose. Furthermore, both the hydrochloric acid and the mixed acetic acid

and nitric acid solution have better capability to swell cellulose, thereby facilitating the breakage of inter-

and intra-molecular hydrogen bonds in the crystalline regions. The absence of surface groups on the NCC

5
prepared by hydrochloric acid and or hydrobromic acid leads to an unstable colloidal dispersion as

compared to the one prepared by sulfuric acid hydrolysis (Börjesson & Westman, 2015).

While nanocrystalline cellulose has potentials as a reinforcement element, low compatibility, and poor

interfacial adhesion between the hydrophilic natural fiber nanofiller and hydrophobic polymer matrix are

the major disadvantages of bionanocomposites. Moreover, the lack of adhesion between the nanofiller and

polymer matrix leads to the formation of strongly bonded nanoparticles aggregates during the

bionanocomposite preparation, thus prompt early failure at the interface of the final composite material

(Iyer, Flores, et al., 2015; Iyer, Schueneman, & Torkelson, 2015; Šupová, Martynková, & Barabaszová,

2011). Hence many studies had been conducted on the improvement of the interface adhesion between the

filler and polymer matrix. The surface of cellulose nanofillers can be modified and tuned either (a) through

physical interactions or adsorption of molecules or macromolecules onto their surface or (b) via chemical

approach to achieve covalent bonds between cellulosic substrates and the grafting agent. Cellulose

nanofillers exhibit a high surface area of 50-70 m2/g which greatly increase the quantity of surface

hydroxyl groups that are available for surface modifications and also it changes the typical conditions of

grafting. TEMPO-mediated oxidation is a common method for modifying selectively the surface of native

cellulose under aqueous and mild conditions. Carboxyl acid groups were introduced on the surface of

cellulose nanoparticles through TEMPO oxidation method. Besides that, esterification of cellulose

nanocrystals has also been used in literature to modify the native cellulose surface. A heterogeneous

catalytic method was successfully conducted by Tingaut et al with the help of solvent exchange process

that was used. to modify the surface of cellulose nanocrystals suspension from water to DMF (Tingaut,

Zimmermann, & Lopez-Suevos, 2010). However, the most commonly used surface modification technique

is solvent exchange process to obtain a non-flocculated dispersion of NCC in an appropriate organic

medium by grafting hydrophobic chains at the surface of NCC (Siqueira, Bras, & Dufresne, 2010).

Furthermore, solvent exchange method of nanocrystalline cellulose could improve the dispersion of

nanofillers in organic solvents and in the polymer matrix, hence enhancing the mechanical and thermal

properties of nanocomposites (H.-Y. Yu, Huang, Chen, & Chang, 2014).

However, to the best of our knowledge, in the past investigations, there is a rare concern on the material

properties of PLA/PHBV/NCC based bionanocomposites. Therefore, in this work the innovative

6
combination of PLA, PHBV blend reinforced with NCC were prepared by solution casting method to

examine their material properties. The main objective of this work is to evaluate the effect of NCC loading

percentage on the morphological, mechanical and structural properties of the prepared bionanocomposites

intended for packaging applications.

2.0 EXPERIMENTAL

2.1 Materials

Oil palm empty fruit bunch (OPEFB) fiber was obtained from the FELCRA Nasaruddin Oil palm mill,

Perak. Commercially available Poly (lactic acid) (PLA) granules and Poly-(3-hydroxybutyrate-co-3-

hydroxyvalerate) (PHBV) with 12 % HV resins obtained from GoodFellow Corp. as a primary and

secondary polymer matrix respectively. Chemicals used in this study were all laboratory grades purchased

from Global Plus Sdn. Bhd and Benua Sains Sdn. Bhd: Sulphuric acid (97.5 %), Hydrochloric acid (37 %),

Acetic acid (99.7 %), Sodium hydroxide (98 %), Benzene (99.0 %), Ethanol (99.5 %), Sodium

Hypochlorite (with Chlorine 10.25 %), Sodium bromide (99 %), Hydrochloric acid (37 %), 2,2,6,6-

tetramethylpiperidine-1-oxyl (TEMPO) (98 %), Acetone (97%), Dimethylformamide (99.8 %) and

Chloroform (99 %).

2.2 Isolation and modification of nanocrystalline cellulose

Alkali treatment of OPEFB was carried out using 200 mL beaker. Initially, 100 mL of 2M NaOH was

added to physically-treated OPEFB sample at 12.5 % (w/v) solid loading. The conditions for pretreatment

were adopted from Sudiyani et al. where the mixture was heated at 80 °C for 120 minutes (Sudiyani et al.,

2013). After that, the samples were filtered to separate the insoluble solid fiber from soluble fractions. The

insoluble fibers were washed with water until it reaches constant neutral pH value of (5-7) and dried at

105°C until it reached a constant weight. The alkali pretreated fiber was then subjected to bleaching

operation using Sodium hypochlorite (1.7 wt %) and in the presence of acetate buffer (8-10 drops). The

ratio of fiber to liquor was (1:25 g/mL). The bleaching pretreatment was performed at 80 °C for 1 hour and

this process was repeated for four times. Finally, the resulting cellulose fiber was washed with distilled

water until it reaches the constant pH value of (5-7).

7
Sulfuric acid hydrolysis reaction was conducted on the bleached cellulose fiber to produce nanocrystalline

cellulose (NCC). Five grams of cellulose were added into 100 mL of Sulfuric acid at a concentration of 45

wt%. The acid hydrolysis reaction was performed at 45 °C under continuous mechanical stirring for 60

minutes. Finally, the reaction was quenched with 10-fold of ice water (4 °C) and centrifuged at 10,000 rpm

to remove excess of acids. This centrifugation step was repeated until it reaches a constant pH value of (5-

7). The precipitate was dialyzed against distilled water for 3 days to remove non-reactive sulfate groups,

salts, and soluble sugars. The resulted suspension was refrigerated at 4 °C for characterization process. The

yield was calculated as a percentage of the initial weight of NCC after the hydrolysis. The final suspension

named as nanocrystalline cellulose (NCC).

One gram of NCC sample was suspended in distilled water (75 mL) containing 2,2,6,6-

tetramethylpiperidine-1-oxyl (TEMPO) (0.0125 g) and Sodium bromide (0.125 g). The reaction was

initiated by the addition of 0.206 mL of 13 % Sodium hypochlorite. The pH value of the suspension was

kept at (10-10.5) by adjusting with 2 wt % of Sodium hydroxide dropwise. After about 30-45 minutes when

there is no more change in the pH value, the oxidation reaction was terminated by the addition of 5 mL of

Ethanol and let it stir for 10 minutes. Thereafter, the suspension was neutralized using 0.5 M Hydrochloric

acid. Finally, the water insoluble product was washed with deionized water by successive centrifugation

(5,000 rpm for 10 min) until it reaches constant neutral pH value of (5-7). The solid content of the

supernatant is measured by drying at 100 ± 5 ˚C for 3 hours to determine the yield (%) of cellulose

nanocrystals based on the dry weight of the oxidized cellulose (Cao, Ding, Yu, & Al-Deyab, 2012;

Filpponen & Argyropoulos, 2010; Habibi, Chanzy, & Vignon, 2006; Missoum, Belgacem, & Bras, 2013).

Solvent exchanging method was used to prepare a stable suspension of carboxylated nanocrystalline

cellulose in Chloroform for polymer composite preparation. Firstly, 100 mL of Acetone was added to 1

wt% of carboxylated NCC suspension (50 mL). The mixture was then agitated for 1 hour and centrifuged

to remove the supernatant. Again the sediment was redispersed in 100 mL of Acetone and stirred for 1

hour. After three successive mixing with Acetone followed by centrifugation, the carboxylated NCC

sediment was dispersed in 100 mL of DMF and agitated for about 2 hours until obtaining a stable

suspension. Subsequently, the suspension was centrifuged to remove the supernatant. Finally, the sediment

8
was dispersed in 100 mL of Chloroform and stirred for 1 hour to obtain a stable suspension of carboxylated

NCC in Chloroform (Dufresne, 2010; Habibi, Lucia, & Rojas, 2010).

2.3 PLA/PHBV blending and bionanocomposites preparation

Two polymers were mixed using solvent casting technique to prepare a blend of PLA/PHBV with a

composition ratio of 70:30. The PLA/PHBV blend was prepared by dissolving PLA/PHBV pellets in

Chloroform at a concentration of 20 wt%. The PLA/PHBV/Chloroform mixtures are continuously stirred at

40 °C until the pellets were fully dissolved. The PLA/PHBV solution was immediately cast on the clean

glass plates and left for the solvent to evaporate at ambient temperature for 48 hours. The thickness of the

cast solution was standardized approximately to 100 µm based on previously published work and noted as

pure PLA/PHBV (Abdelwahab et al., 2012; L. Yu, Dean, & Li, 2006; Zembouai et al., 2013; Zhao, Cui,

Wang, Turng, & Peng, 2013).

The bionanocomposite film was prepared by mixing 20 wt% suspension of PLA/PHBV blend solution with

varying loading percentage of modified NCC and was stirred for 8 hours. The suspension was then

sonicated for 5 minutes before casting on a clean glass plate to generate a composite film with a thickness

of 100 µm after removal of solvent (Haafiz et al., 2013; D. Liu, Yuan, Bhattacharyya, & Easteal, 2010;

Tao, Yan, & Jie, 2009). The polymer blend and bionanocomposites was designated as in Table 1. Besides

that, the overall processing techniques of nanocrystalline cellulose and bionanocomposite preperations are

shown schematically in Fig. 1.

TABLE 1. Polymer blend (PB) and bionanocomposites (BC) formulations


Samples Sample Code PLA/PHBV NCC (wt%)

Polymer blend PB 70/30 -

Bionanocomposite 1 BC1 70/30 0.25

Bionanocomposite 2 BC2 70/30 0.50

Bionanocomposite 3 BC3 70/30 0.75

Bionanocomposite 4 BC4 70/30 1.00

Bionanocomposite 5 BC5 70/30 1.50

Bionanocomposite 6 BC6 70/30 2.00

9
Fig 1. Schematic diagram of nanocrystalline cellulose and bionanocomposites preparation a) Isolation of NCC; b)
Surface modification of NCC; c) Development of bionanocomposites

2.4 Characterizations

2.4.1 Fourier transformed infrared spectroscopy (FTIR)


Perkin-Elmer infrared spectrometer was used to obtain spectra for raw OPEFB, nanocrystalline cellulose

(NCC). The KBr pellet method was used in taking the IR spectra. Samples were mixed with KBr

(Sample/KBr ratio 1/100) to prepare pastilles. FTIR spectra analysis was performed within the wave

number range of 400-4000 cm-1.

2.4.2 X-ray diffraction (XRD)

X-ray diffraction was used to determine the crystallinity of raw OPEFB, nanocrystalline cellulose (NCC)

and PLA/PHBV/NVV bionanocomposites. The diffractograms were recorded using Bruker D8 Advance

Power diffractometer with Cu Kα radiation (λ=1.54Å) for a 2θ range from 10° to 60°. The crystalline-to-

amorphous ratio of materials was determined using Equation (1), proposed by Segal et al (1959).and

Computer aided curve resolving technique Equation (2) which was developed by Segal et al., and for

nanocellulose and bionanocomposites respectively.

I002 −Iam
Cr.I (%) = (1)
I002

10
where Cr.I is the crystallinity Index, I002 is a maximum Intensity of diffraction from 002 planes at 2θ=22°,

and Iam is an intensity of the background scattered measured at 2θ=15°.

𝐴𝑐
𝑋 (%) = × 100 % (2)
𝐴𝑐 + 𝐴𝑎

where Aa is the experimental integrated intensity of amorphous phase and Ac is the experimental integrated
intensity of crystalline phase.

2.4.3 Atomic force microscope (AFM)


The morphology and size distribution of the sample was studied by using Nano Navi (E-sweep) Universal

Scanning Probe Microscopy (USPM) with scanning area of (2.5 x 2.5) nm. Samples were prepared by

placing a drop of diluted NCC suspension (0.001 %) onto a clean glass slide (covered with aluminum foil)

and dried at ambient conditions.

2.4.4 Field emission scanning electron microscope (FESEM)


The fracture surface of composites was studied to identify the nature of the failure mechanism and

particularly to examine the level of fiber matrix interaction. Field Emission Scanning Electron Microscopy

(FESEM model Zeiss Supra VP55) with an acceleration voltage of 20 kV was used to evaluate the surface

topography and tensile fracture surface of neat PLA/PHBV blend and PLA/PHBV/NCC blend composite

film. The specimen was coated with gold to avoid charging (Oksman, Mathew, Långström, Nyström, &

Joseph, 2009).

2.4.5 Transmission Electron Microscope (TEM)


The shape and size of NCC were analyzed with Transmission Electron Microscopy (TEM). Initially, NCC

samples were ultrasonically dispersed in anhydrous ethanol to form an NCC suspension. A droplet of NCC

suspension was deposited on a carbon microgrid (400 meshes) and allowed to dry. The grid was then

negatively stained with Uranyl acetate (around 2 wt. %) for 3 min. Subsequently, TEM images were

obtained by using a LIBRA 120 Energy Filter Transmission Electron Microscope (EFTEM) as in at an

accelerating voltage of 80 kV. Moreover, to study the morphology of PLA/PHBV/NCC composite film a

thin section were cut from the film with a diamond knife on a Reichert-Jung ultramicrotome. The trimmed

section was placed on a carbon-coated copper grid and then stained with 2 wt% uranyl acetate for 3

minutes.

11
2.4.6 Flexural properties
The flexural testing was performed according to ASTM D790 standard for flexural testing on Universal

Testing Machine AMATEK INC (Model: LLOYD LR). First, the bionanocomposite samples were

prepared through solution casting method. Subsequently, the samples specimens for the flexural test were

prepared in a rectangular shape 127 x 12.7 x 3.2 mm using CARVER Inc. compression molding machine.

2.4.7 Gas permeability


The oxygen transmission of the composite film was measured using 4-channel low-pressure flat sheet

permeation cell unit. The samples were prepared in a square shape with 5.5 cm diameter. The oxygen

permeability value was calculated by the multiplication of Oxygen Transmission Rate (OTR) at steady state

by the average film thickness divided by the partial pressure differences between the two sides of the film.

Prior to the testing, samples were equilibrated at 25 °C and 55% of relative humidity for at least 2 days in

desiccators.

3.0 RESULTS AND DISCUSSION


3.1 Properties of nanocrystalline cellulose
Chemical groups on the surface on NCC were characterized using FTIR spectra, as shown Fig.2a. The

broad bands in the region of 3340 cm-1 to 3500 cm-1 attributes to the O-H stretching vibrations and the

peaks at 2900 cm-1 and 2905 cm-1 correspond to C-H stretching vibrations (Lu & Hsieh, 2010; Pasquini,

Teixeira, Curvelo, Belgacem, & Dufresne, 2010). The peak at region of 1638 cm-1 indicates absorbed water

(H2O) in the cellulose powder, while the peak at 1206 cm-1 represent sulfate groups of the NCC from the

esterification reactions (D. Chen, D. Lawton, M. Thompson, & Q. Liu, 2012; Nazir, Wahjoedi, Yussof, &

Abdullah, 2013; Neto, Silvério, Dantas, & Pasquini, 2013). Furthermore, the NCC spectra shows the

disappearance of C=O stretching vibration at 1722.16 cm-1and 1533 cm-1 which corresponds to the acetyl

and uronic ester groups of hemicelluloses as well as the ester linkage of the carboxyl groups of lignin and

to the C=C vibration in lignin respectively (Dan Chen et al., 2012; Johar, Ahmad, & Dufresne, 2012).

Finally, the majority of peaks that present in the region of 800-1500 cm-1for both raw and NCC sample

represents unique fingerprint region for cellulose(D. Chen, D. Lawton, M. R. Thompson, & Q. Liu, 2012).

The X-ray diffractograms in Fig. 2b shows the diffraction patterns of raw OPEFB and nanocrystalline

cellulose (NCC) with diffraction peaks around 2θ = 15°, 22°, and 34°, corresponding to the (101), (002),

and (040) crystallographic planes of typical cellulose I structure (Y. Liu et al., 2014; Nguyen et al., 2013).

12
The calculated crystallinity index (CI) value for NCC is 82.22 % as compared to CI of 35.29 % for raw

OPEFB. The increase in crystallinity index of NCC attributed to the hydrolysis of the amorphous region in

raw OPEFB fibers. Furthermore, the diffraction peaks of NCC at 15° (101), and 22° (002) became more

intense as compared to raw fibers indicating higher perfection of crystal lattice (Fahma, Iwamoto, Hori,

Iwata, & Takemura, 2010; Y. Liu et al., 2014; Mohamad, Eichhorn, Hassan, & Jawaid, 2013).

Fig. 2c and 1d present the TEM and AFM micrographs of the extracted nanocrystalline cellulose. It can be

observed from the images that NCC appears as a rod-like structure with some agglomerated network

(Zhou, Fu, Zheng, & Zhan). The established strong hydrogen bond and high specific area of NCC is one of

the reasons for the appearance of laterally aggregated elementary crystallites in TEM and AFM images

(Neto et al., 2013). Furthermore, the average diameter distributions of nanocrystalline cellulose are 63 nm

and 8.75 nm for AFM and TEM samples respectively as compared to 300 μm of raw OPEFB. It was

expected that controlled acid hydrolysis could cleave amorphous region of cellulosic fibers by keeping the

crystalline domains intact. Therefore, acid hydrolysis treatments tend to reduce the size of the fibers size

from micron to the nanometer scale (Johar et al., 2012). The diameter distribution of NCC’s is significantly

different from the one obtained through AFM analysis. This is due to the AFM tip broadening effect which

results in an overestimation of particles dimension (Tsukruk & Singamaneni, 2012). Furthermore, there are

few works which reported diameter values of nanocrystalline cellulose with a big difference between TEM

and AFM analysis (Mohamed, Salleh, Jaafar, Asri, & Ismail, 2015).

13
Fig. 2. Characterization of nanocrystalline cellulose (NCC): (a) FTIR spectra of raw OPEFB and NCC; (b) X-ray diffraction
patterns for raw OPEFB and NCC; (c) TEM micrograph of extracted NCC; (d) AFM image of extracted NCC

3.2 Properties of PLA/PHBV/nanocrystalline cellulose composite


3.2.1 XRD analysis
X-ray diffraction analysis was used to determine the degree of crystallinity and crystalline structure of

polymer blend and their composites. As shown in Fig 3 for PLA/PHBV blend (PB), the three prominent

peaks are located at 2θ= 13.5°, 16.8°, 19° and 22.3°, corresponding to the (020), (110), (010) and (111)

planes of PHBV respectively. Furthermore, PB has weak reflection peaks at 2θ= 15° representing

(110/200) plane of PLA. In general, all the samples showed diffraction patterns similar to PLA/PHBV

blend (Abdelwahab et al., 2012; Arrieta et al., 2014). However, there is a small shift in the angles of

bionanocomposites samples due to the presence of NCCs which alters the crystalline structure exhibited by

PB at 2θ= 22.3° (111). It can be observed that there are two new diffraction peaks evolved in

bionanocomposite samples at 2θ= 22.5° and 34°, corresponding to (200) and (040) planes respectively.

These two peaks are recognized as typical diffraction patterns of cellulose I structure (Suryanegara,

Nakagaito, & Yano, 2010). The intensity of peak for bionanocomposites (BC) samples was almost similar

to those of PB. However, as the NCC composition reaches 1.5 g and 2.0 g the recorded peak intensity has

slightest decreased which indicates low crystallinity of material (Table 2). The changes in crystallinity

14
index (CI) of PB and BCs were recorded in Table 2. The calculated values of CI showed a significant

increase of crystallinity in BC1 and BC2 by 10 % as compared to the PB (50.67 %) due to the high

crystalline structure of NCC. However, as the loading percentage of NCC increases, there is a gradual

decrease in the CI values of BCs. Excess in nanocrystalline cellulose loading percentage causes the

molecular chains of polymer matrices to pull the NCC aggregates into a closely packed form, thus leads to

the agglomeration of the NCCs at higher concentration. Furthermore, the CI values of BC5 and BC6 was

lower than the neat polymer blend (PB) is due to the agglomeration, thereby not contributing to the

enhancement in the percentage of crystallinity of the bionanocomposites (Vengatesan & Mittal, 2016).

Fig. 3. XRD analysis for polymer blend (PB) and bionanocomposites (BC)

15
Table 2: Crystallinity data for polymer blend (PB) and bionanocomposites (BC) film

Samples Crystallinity Index (CI) (%)


PB 50.67
BC1 71.69
BC2 71.11
BC3 55.29
BC4 55.13
BC5 44.29
BC6 34.78

3.2.2 Transmission electron microscopy (TEM) analysis


Fig. 4 shows the morphology of bionanocomposites film analyzed by Transmission electron microscopy

(TEM) at nanoscale dimension. The areas where the staining appeared to be more concentrated (some are

indicated with red arrows) represent an agglomeration of nanocrystalline cellulose (NCC). On the contrary,

the low contrast between the NCC and polymer matrices suggests the uniform distribution of NCC within

the polymer matrix (represented by black arrows). A reasonable contrast between the nanocrystalline

cellulose and the surroundings was given by the staining process to constitute a continuous background

(Mohanty, Nayak, Kaith, & Kalia, 2015). The TEM micrographs of bionanocomposites (BC2 to BC6)

shows that the majority of the NCCs were present in the flakes form due to agglomeration and the number

of NCC agglomerates increased as the NCC loading percentages increase. From Fig. 4 TEM analysis, it

was observed that the NCCs in the polymer matrix have particles diameter range from 34 – 87 nm.

However, BC1 shows the most uniform distribution of NCCs as compared to the other bionanocomposite

films. Furthermore, the NCCs that are visible in BC1 are more needles like shapes which indicate single

crystals. This indicates that NCC has been less agglomerated within the matrix material in case of BC1 film

as compared to other bionanocomposite films.

16
Fig. 4. TEM micrographs of polymer blend (PB) and bionanocomposites (BC)

3.2.3 Field emission scanning electron microscope (FESEM) analysis


Polymer blend (PB) and bionanocomposites (BCs) were examined using FESEM micrographs of the tensile

fractured surface (Fig. 5) revealed relatively rougher surfaces in the nanocrystalline cellulose reinforced

bionanocomposites as compared to the neat polymer blend. The aggregated NCC structure observed on the

BC micrographs (indicated with red arrow) does not appear to be debonded from the matrix, which

indicates good fiber-matrix interfacial adhesion. However, it can be seen clearly from the micrographs of

BC that the increase in NCC content produces modifications in the matrix surface with pull out breakage of

fibers. This pull out indicates that stress builds up exceeding the fiber-matrix bond strength. Moreover, the

BC2 to BC6 bionanocomposites samples shows more brittle failure and the appearance of voids on the

fractured surface (Qamhia, Sabo, & Elhajjar, 2013). Furthermore, results from Fig 5. show that

Nanocrystalline cellulose (NCC) is well compatible with PLA/PHBV blend at low loading percentage of

NCC (< 0.75 wt %), taking into account the fact that the NCCs are well embedded within the polymer

blend matrix and there is no gross phase separation or void formation at the interface (Yuwawech,

Wootthikanokkhan, & Tanpichai, 2015).

17
Fig. 5. FESEM micrographs of polymer blend (PB) and bionanocomposites (BC)

3.2.4 Flexural properties analysis

The flexural properties of PB and BCs are demonstrated in Fig. 6a, although the flexural strength (MPa)

and flexural modulus (GPa) of bionanocomposites (BCs) were seen to increase initially, yet the graph

demonstrates a gradual decrease in their flexural properties when compared to that of the polymer blend

(PB). According to the data in Fig. 6a, the flexural strength and modulus of PB are 24.85 MPa and 1.395

GPa, respectively. However, for bionanocomposites (BC1 to BC6) those values have significantly

decreased compared to PB due to the addition of nanocrystalline cellulose (NCC). The maximum flexural

18
strength (MPa) and flexural modulus (GPa) was recorded by BC1 which is 104.02 % and 27.74 %,

respectively greater than PB properties. This is due to the optimal dispersion of NCC within the polymer

blend (PB) matrix which indicates the efficient stress transfer between polymer blend and NCCs which

improves resistance to bending (Iyer, Schueneman, et al., 2015). On the other hand, this increase in

flexural properties was attributed to the mechanical interlocking made possible by the nanocrystalline

cellulose (NCC) deposits within the blend composition, which improved the interfacial properties between

the fiber and matrix (Hossain, Felfel, Rudd, Thielemans, & Ahmed, 2014). It is also found that the values

of flexural strength and modulus decreases as the NCCs loading percentage increases due to the

agglomeration of NCCs and stress transfer get blocked (Jacob, Thomas, & Varughese, 2004; Khalil, Fizree,

Bhat, Jawaid, & Abdullah, 2013; Shinoj, Visvanathan, Panigrahi, & Kochubabu, 2011; Suryanegara et al.,

2010). The lowest flexural properties recorded for BC6 which has highest loading percentage of NCC (2.0

wt %). The BC6 had a flexural strength (MPa) and flexural modulus (GPa) of 38.53 % and 1.322 %,

respectively, which was lowest as compared to the values of PB and other BCs. Furthermore, the results of

flexural strength and modulus are strongly related to crystallinity percentage of the composites. The results

demonstrate that the high crystallinity of sample promotes the resistance of crack propagation. Therefore,

bionanocomposites sample with high crystallinity index (CI) exhibits highly disordered crystallites without

boundary crystals displayed the best flexural property results (T. Yu, Wu, Chang, Wang, & Rwei, 2012).

3.2.5 Gas permeability analysis


Fig. 6b shows the plot of oxygen permeability (OP) data that is obtained for the polymer blend (PB) and

bionanocomposite (BCs) films. The PB found to have OP of about 0.2999 (cm3 (STP).cm) /

(cm2.day.cmHg). However, the OP values have decreased in BCs with reinforcement of NCC in it. As can

be seen from Fig. 6b, the OP value of BC1, BC2, and BC3 has reduced by 27.64 %, 13.67 %, and 5.17 %

respectively. The presence of nanocrystalline cellulose reinforcements in semi-crystalline polymers is

believed to enhance the tortuosity in the composite materials resulting slower diffusion process, thus lowers

permeability of oxygen. Besides that, OP properties can be improved by high aspect ratio and good

dispersion of filler materials in the polymer matrix as it serves as a trap to preserve the active oxygen

rummages in the polymer matrix while reducing the rate of oxygen transmission. (Azeredo, Miranda,

Ribeiro, Rosa, & Nascimento, 2012; Jamal, Anuar, Razak, & Bahri, 2010; Khan et al., 2010; Siró &

Plackett, 2010). The maximum reduction of OP was recorded by BC1 while the OP values of other BCs

19
shows increasing trend as the filler composition increases. As the NCC filler composition increases the OP

values of BC4, BC5, BC6 gets higher than that of neat PB films. The decrease in OP values is due to the

phase separation of poor particle NCC distribution and aggregation which resulted in the reduction of

crystallinity of BCs film refers to the excess addition of NCC (Savadekar, Karande, Vigneshwaran, Kadam,

& Mhaske, 2014). Besides that, the reduction in crystallinity index that obtained through XRD analysis

could be another reason for high oxygen permeance in BC4-BC6. This is because the reduction in

crystallinity indicates that the material is more amorphous, hence more permeable for oxygen molecules

(Jamal et al., 2010). Thus, this confirms that optimal NCC loading percentage is important in order to

achieve remarkable oxygen permeability barriers. The results show NCC loading percentage below 1wt %

is the best to enhance barrier properties of NCC reinforced composites.

Fig. 6. a) Typical flexural properties of polymer blend (PB) and bionanocomposites (BC); Inserted image show the three point
bending that was conducted to determine the flexural properties, b) Oxygen permeability properties of polymer blend (PB) and
bionanocomposites (BC)

4.0 CONCLUSION

Novel environmental friendly bionanocomposite were successfully prepared by incorporating nanocrystalline

cellulose in PLA/PHBV polymer blend. The fabricated bionanocomposites with 0.25 wt% of NCC content has

shown improved characteristics as compared to the neat polymer blend. The addition of an optimal amount of

nanocrystalline cellulose has improved the interfacial adhesion between PLA, PHBV, and nanofiller, thus enhanced

the mechanical, morphological and barrier properties. However, an excess amount of nanocrystalline cellulose have

shown negative effects on developed bionancomposites due to the aggregation of NCCs in polymer blend and then

chain movement restriction respectively. An effort has been made to develop environmental friendly

20
bionanocomposites based on nanocrystalline cellulose which was extracted at the lowest concentration possible.

Therefore, this study proposes that nanocrystalline cellulose is a potentially attractive reinforcing agent for

thermoplastics based composites.

ACKNOWLEDGMENT

The authors would like to thank Universiti Teknologi PETRONAS for their financial support under project grant of

YUTP –FRG, 2015.

REFERENCES

Abdelwahab, M. A., Flynn, A., Chiou, B.-S., Imam, S., Orts, W., & Chiellini, E. (2012).
Thermal, mechanical and morphological characterization of plasticized PLA–PHB
blends. Polymer Degradation and Stability, 97(9), 1822-1828.
Arrieta, M., Fortunati, E., Dominici, F., Rayón, E., López, J., & Kenny, J. (2014).
Multifunctional PLA–PHB/cellulose nanocrystal films: Processing, structural and thermal
properties. Carbohydrate Polymers, 107, 16-24.
Avérous, L., & Pollet, E. (2012). Biodegradable Polymers. In L. Avérous & E. Pollet (Eds.),
Environmental Silicate Nano-Biocomposites (pp. 13-39). London: Springer London.
Azeredo, H. M., Miranda, K. W., Ribeiro, H. L., Rosa, M. F., & Nascimento, D. M. (2012).
Nanoreinforced alginate–acerola puree coatings on acerola fruits. Journal of Food
Engineering, 113(4), 505-510.
Börjesson, M., & Westman, G. (2015). Crystalline Nanocellulose — Preparation, Modification,
and Properties, Cellulose. In D. M. P. (Ed.) (Ed.), Fundamental Aspects and Current
Trends: InTech.
Cao, X., Ding, B., Yu, J., & Al-Deyab, S. S. (2012). Cellulose nanowhiskers extracted from
TEMPO-oxidized jute fibers. Carbohydrate Polymers, 90(2), 1075-1080.
Chen, D., Lawton, D., Thompson, M., & Liu, Q. (2012). Biocomposites reinforced with cellulose
nanocrystals derived from potato peel waste. Carbohydrate Polymers, 90(1), 709-716.
Chen, D., Lawton, D., Thompson, M. R., & Liu, Q. (2012). Biocomposites reinforced with
cellulose nanocrystals derived from potato peel waste. Carbohydrate Polymers, 90(1),
709-716. doi:http://dx.doi.org/10.1016/j.carbpol.2012.06.002
dos Santos, F. A., Iulianelli, G. C., & Tavares, M. I. B. (2016). The Use of Cellulose Nanofillers
in Obtaining Polymer Nanocomposites: Properties, Processing, and Applications.
Materials Sciences and Applications, 7(05), 257.
Dufresne, A. (2010). Processing of Polymer Nanocomposites Reinforced with Polysaccharide
Nanocrystals. Molecules, 15(6), 4111-4128.
Dufresne, A. (2013). Nanocellulose: a new ageless bionanomaterial. Materials Today, 16(6),
220-227.
Fahma, F., Iwamoto, S., Hori, N., Iwata, T., & Takemura, A. (2010). Isolation, preparation, and
characterization of nanofibers from oil palm empty-fruit-bunch (OPEFB). Cellulose,
17(5), 977-985.
Filpponen, I., & Argyropoulos, D. S. (2010). Regular linking of cellulose nanocrystals via click
chemistry: Synthesis and formation of cellulose nanoplatelet gels. Biomacromolecules,
11(4), 1060-1066.

21
Haafiz, M. M., Hassan, A., Zakaria, Z., Inuwa, I., Islam, M., & Jawaid, M. (2013). Properties of
polylactic acid composites reinforced with oil palm biomass microcrystalline cellulose.
Carbohydrate Polymers, 98(1), 139-145.
Habibi, Y., Chanzy, H., & Vignon, M. R. (2006). TEMPO-mediated surface oxidation of
cellulose whiskers. Cellulose, 13(6), 679-687. doi:10.1007/s10570-006-9075-y
Habibi, Y., Lucia, L. A., & Rojas, O. J. (2010). Cellulose nanocrystals: Chemistry, self-
assembly, and applications. Chemical Reviews, 110(6), 3479-3500.
Hossain, K. M. Z., Felfel, R. M., Rudd, C. D., Thielemans, W., & Ahmed, I. (2014). The effect
of cellulose nanowhiskers on the flexural properties of self-reinforced polylactic acid
composites. Reactive and functional polymers, 85, 193-200.
doi:http://dx.doi.org/10.1016/j.reactfunctpolym.2014.09.012
Ibrahim, M. M., & El-Zawawy, W. K. (2015). Extraction of cellulose nanofibers from cotton
linter and their composites Handbook of Polymer Nanocomposites. Processing,
Performance and Application (pp. 145-164): Springer.
Iyer, K. A., Flores, A. M., & Torkelson, J. M. (2015). Comparison of polyolefin biocomposites
prepared with waste cardboard, microcrystalline cellulose, and cellulose nanocrystals via
solid-state shear pulverization. Polymer, 75, 78-87.
doi:http://dx.doi.org/10.1016/j.polymer.2015.08.029
Iyer, K. A., Schueneman, G. T., & Torkelson, J. M. (2015). Cellulose nanocrystal/polyolefin
biocomposites prepared by solid-state shear pulverization: Superior dispersion leading to
synergistic property enhancements. Polymer, 56, 464-475.
doi:http://dx.doi.org/10.1016/j.polymer.2014.11.017
Iyer, K. A., Zhang, L., & Torkelson, J. M. (2016). Direct Use of Natural Antioxidant-rich Agro-
wastes as Thermal Stabilizer for Polymer: Processing and Recycling. ACS Sustainable
Chemistry & Engineering, 4(3), 881-889. doi:10.1021/acssuschemeng.5b00945
Jacob, M., Thomas, S., & Varughese, K. T. (2004). Mechanical properties of sisal/oil palm
hybrid fiber reinforced natural rubber composites. Composites Science and Technology,
64(7), 955-965.
Jamal, N. A., Anuar, H., Razak, A., & Bahri, S. (2010). A linear relationship between the tensile,
thermal and gas barrier properties of mape modified rubber toughened nanocomposites.
IIUM Engineering Journal, 11(2), 225-239.
Jamshidian, M., Tehrany, E. A., Imran, M., Jacquot, M., & Desobry, S. (2010). Poly‐Lactic
Acid: production, applications, nanocomposites, and release studies. Comprehensive
Reviews in Food Science and Food Safety, 9(5), 552-571.
Johar, N., Ahmad, I., & Dufresne, A. (2012). Extraction, preparation and characterization of
cellulose fibres and nanocrystals from rice husk. Industrial Crops and Products, 37(1),
93-99.
Jonoobi, M., Oladi, R., Davoudpour, Y., Oksman, K., Dufresne, A., Hamzeh, Y., & Davoodi, R.
(2015). Different preparation methods and properties of nanostructured cellulose from
various natural resources and residues: a review. Cellulose, 22(2), 935-969.
Kalia, S., Dufresne, A., Cherian, B. M., Kaith, B., Avérous, L., Njuguna, J., & Nassiopoulos, E.
(2011). Cellulose-based bio-and nanocomposites: a review. International Journal of
Polymer Science, 2011.
Khalil, H. A., Fizree, H., Bhat, A. H., Jawaid, M., & Abdullah, C. K. (2013). Development and
characterization of epoxy nanocomposites based on nano-structured oil palm ash.
Composites Part B: Engineering, 53, 324-333.

22
Khan, R. A., Salmieri, S., Dussault, D., Uribe-Calderon, J., Kamal, M. R., Safrany, A., &
Lacroix, M. (2010). Production and properties of nanocellulose-reinforced
methylcellulose-based biodegradable films. Journal of agricultural and food chemistry,
58(13), 7878-7885.
Liu, D., Yuan, X., Bhattacharyya, D., & Easteal, A. (2010). Characterisation of solution cast
cellulose nanofibre—reinforced poly (lactic acid). Express Polymer Letters, 4(1), 26-31.
Liu, Y., Wang, H., Yu, G., Yu, Q., Li, B., & Mu, X. (2014). A novel approach for the
preparation of nanocrystalline cellulose by using phosphotungstic acid. Carbohydrate
Polymers, 110, 415-422.
Lu, P., & Hsieh, Y.-L. (2010). Preparation and properties of cellulose nanocrystals: rods,
spheres, and network. Carbohydrate Polymers, 82(2), 329-336.
Martínez-Sanz, M., Villano, M., Oliveira, C., Albuquerque, M. G., Majone, M., Reis, M., . . .
Lagaron, J. M. (2014). Characterization of polyhydroxyalkanoates synthesized from
microbial mixed cultures and of their nanobiocomposites with bacterial cellulose
nanowhiskers. New Biotechnology, 31(4), 364-376.
Missoum, K., Belgacem, M., & Bras, J. (2013). Nanofibrillated Cellulose Surface Modification:
A Review. Materials, 6(5), 1745-1766. doi:10.3390/ma6051745
Mohamad, H. M. K., Eichhorn, S. J., Hassan, A., & Jawaid, M. (2013). Isolation and
characterization of microcrystalline cellulose from oil palm biomass residue. Carbohydr
Polym, 93(2), 628-634. doi:10.1016/j.carbpol.2013.01.035
Mohamed, M., Salleh, W., Jaafar, J., Asri, S., & Ismail, A. (2015). Physicochemical properties of
“green” nanocrystalline cellulose isolated from recycled newspaper. RSC Advances,
5(38), 29842-29849.
Mohanty, S., Nayak, S. K., Kaith, B., & Kalia, S. (2015). Polymer Nanocomposites Based on
Inorganic and Organic Nanomaterials: John Wiley & Sons.
Nazir, M. S., Wahjoedi, B. A., Yussof, A. W., & Abdullah, M. A. (2013). Eco-friendly
extraction and characterization of cellulose from oil palm empty fruit bunches.
BioResources, 8(2), 2161-2172.
Neto, W. P. F., Silvério, H. A., Dantas, N. O., & Pasquini, D. (2013). Extraction and
characterization of cellulose nanocrystals from agro-industrial residue–Soy hulls.
Industrial Crops and Products, 42, 480-488.
Nguyen, H. D., Mai, T. T. T., Nguyen, N. B., Dang, T. D., Le, M. L. P., & Dang, T. T. (2013). A
novel method for preparing microfibrillated cellulose from bamboo fibers. Advances in
Natural Sciences: Nanoscience and Nanotechnology, 4(1), 015016.
Oksman, K., Mathew, A. P., Långström, R., Nyström, B., & Joseph, K. (2009). The influence of
fibre microstructure on fibre breakage and mechanical properties of natural fibre
reinforced polypropylene. Composites Science and Technology, 69(11), 1847-1853.
Othman, S. H. (2014). Bio-nanocomposite materials for food packaging applications: types of
biopolymer and nano-sized filler. Agriculture and Agricultural Science Procedia, 2, 296-
303.
Pasquini, D., Teixeira, E. d. M., Curvelo, A. A. d. S., Belgacem, M. N., & Dufresne, A. (2010).
Extraction of cellulose whiskers from cassava bagasse and their applications as
reinforcing agent in natural rubber. Industrial Crops and Products, 32(3), 486-490.
doi:10.1016/j.indcrop.2010.06.022

23
Ploypetchara, N., Suppakul, P., Atong, D., & Pechyen, C. (2014). Blend of polypropylene/poly
(lactic acid) for medical packaging application: physicochemical, thermal, mechanical,
and barrier properties. Energy Procedia, 56, 201-210.
Qamhia, I. I., Sabo, R. C., & Elhajjar, R. F. (2013). Static and dynamic characterization of
cellulose nanofibril scaffold-based composites. BioResources, 9(1), 381-392.
Ramamoorthy, S. K., Skrifvars, M., & Persson, A. (2015). A review of natural fibers used in
biocomposites: plant, animal and regenerated cellulose fibers. Polymer Reviews, 55(1),
107-162.
Saba, N., Tahir, P. M., & Jawaid, M. (2014). A review on potentiality of nano filler/natural fiber
filled polymer hybrid composites. Polymers, 6(8), 2247-2273.
Savadekar, N., Karande, V., Vigneshwaran, N., Kadam, P., & Mhaske, S. (2014). Preparation of
cotton linter nanowhiskers by high-pressure homogenization process and its application
in thermoplastic starch. Applied Nanoscience, 5(3), 281-290.
Shinoj, S., Visvanathan, R., Panigrahi, S., & Kochubabu, M. (2011). Oil palm fiber (OPF) and its
composites: A review. Industrial Crops and Products, 33(1), 7-22.
Siqueira, G., Bras, J., & Dufresne, A. (2010). Cellulosic bionanocomposites: a review of
preparation, properties and applications. Polymers, 2(4), 728-765.
Siró, I., & Plackett, D. (2010). Microfibrillated cellulose and new nanocomposite materials: a
review. Cellulose, 17(3), 459-494.
Sudiyani, Y., Styarini, D., Triwahyuni, E., Sudiyarmanto, Sembiring, K. C., Aristiawan, Y., . . .
Han, M. H. (2013). Utilization of Biomass Waste Empty Fruit Bunch Fiber of Palm Oil
for Bioethanol Production Using Pilot–Scale Unit. Energy Procedia, 32, 31-38.
doi:10.1016/j.egypro.2013.05.005
Šupová, M., Martynková, G. S., & Barabaszová, K. (2011). Effect of nanofillers dispersion in
polymer matrices: a review. Science of Advanced Materials, 3(1), 1-25.
Suryanegara, L., Nakagaito, A. N., & Yano, H. (2010). Thermo-mechanical properties of
microfibrillated cellulose-reinforced partially crystallized PLA composites. Cellulose,
17(4), 771-778.
Tao, Y., Yan, L., & Jie, R. (2009). Preparation and properties of short natural fiber reinforced
poly (lactic acid) composites. Transactions of Nonferrous Metals Society of China, 19,
s651-s655.
Thakur, M. K., Thakur, V. K., & Prasanth, R. (2014). Nanocellulose‐Based Polymer
Nanocomposites: An Introduction. Nanocellulose Polymer Nanocomposites:
Fundamentals and Applications, 1-15.
Tingaut, P., Zimmermann, T., & Lopez-Suevos, F. (2010). Synthesis and Characterization of
Bionanocomposites with Tunable Properties from Poly(lactic acid) and Acetylated
Microfibrillated Cellulose. Biomacromolecules, 11(2), 454-464. doi:10.1021/bm901186u
Tsukruk, V. V., & Singamaneni, S. (2012). Scanning Probe Microscopy of Soft Matter:
Fundamentals and Practices: Wiley.
Vengatesan, M. R., & Mittal, V. (2016). Nanoparticle-and Nanofiber-Based Polymer
Nanocomposites: An Overview. Spherical and Fibrous Filler Composites, 1, 1.
Verma, D., Gope, P., Maheshwari, M., & Sharma, R. Bagasse fiber composites-A review.
Yu, H.-Y., Huang, J., Chen, Y., & Chang, P. R. (2014). Preparation of Polysaccharide
Nanocrystal-Based Nanocomposites. Polysaccharide-Based Nanocrystals: Chemistry and
Applications, 109-164.

24
Yu, L., Dean, K., & Li, L. (2006). Polymer blends and composites from renewable resources.
Progress in Polymer Science, 31(6), 576-602.
Yu, T., Wu, C., Chang, C., Wang, C., & Rwei, S. (2012). Effects of crystalline morphologies on
the mechanical properties o. Express Polymer Letters, 6(4).
Yuwawech, K., Wootthikanokkhan, J., & Tanpichai, S. (2015). Effects of Two Different
Cellulose Nanofiber Types on Properties of Poly(vinyl alcohol) Composite Films.
Journal of Nanomaterials, 2015, 10. doi:10.1155/2015/908689
Zembouai, I., Kaci, M., Bruzaud, S., Benhamida, A., Corre, Y.-M., & Grohens, Y. (2013). A
study of morphological, thermal, rheological and barrier properties of Poly (3-
hydroxybutyrate-Co-3-Hydroxyvalerate)/polylactide blends prepared by melt mixing.
Polymer Testing, 32(5), 842-851.
Zhao, H., Cui, Z., Wang, X., Turng, L.-S., & Peng, X. (2013). Processing and characterization of
solid and microcellular poly (lactic acid)/polyhydroxybutyrate-valerate (PLA/PHBV)
blends and PLA/PHBV/clay nanocomposites. Composites Part B: Engineering, 51, 79-
91.
Zhou, Y., Fu, S., Zheng, L., & Zhan, H. Effect of nanocellulose isolation techniques on the
formation of reinforced poly (vinyl alcohol) nanocomposite films.

25

You might also like