You are on page 1of 12

Carbohydrate Polymers 129 (2015) 156–167

Contents lists available at ScienceDirect

Carbohydrate Polymers
journal homepage: www.elsevier.com/locate/carbpol

Bio-nanocomposite films reinforced with cellulose nanocrystals:


Rheology of film-forming solutions, transparency, water vapor barrier
and tensile properties of films
Nassima El Miri a,b , Karima Abdelouahdi c , Abdellatif Barakat d , Mohamed Zahouily a,b ,
Aziz Fihri b , Abderrahim Solhy e,∗ , Mounir El Achaby e,∗
a
Laboratoire de Matériaux, Catalyse & Valorisation des Ressources Naturelles, FST Mohammedia, Université Hassan II de Casablanca, Rue Tarik Bnou Ziad,
Mers Sultan, 9167 Casablanca, Morocco
b
MAScIR Foundation, Rabat Design, Rue Mohamed El Jazouli, Madinat Al Irfane, 10100 Rabat, Morocco
c
Division UATRS, Centre National pour la Recherche Scientifique et Technique (CNRST), Angle Allal Fassi/FAR, Hay Riad, 10100 Rabat, Morocco
d
INRA, UMR 1208 Ingénierie des Agropolymères et Technologies Emergentes (IATE), 2, place Pierre Viala, 34060 Montpellier Cedex 1, France
e
Center For Advanced Materials, Université Mohammed VI Polytechnique, Lot 660-Hay Moulay Rachid, 43150 Ben Guerir, Morocco

a r t i c l e i n f o a b s t r a c t

Article history: This study was aimed to develop bio-nanocomposite films of carboxymethyl cellulose (CMC)/starch (ST)
Received 29 January 2015 polysaccharide matrix reinforced with cellulose nanocrystals (CNC) using the solution casting method.
Received in revised form 22 April 2015 The CNC were extracted at the nanometric scale from sugarcane bagasse via sulfuric acid hydrolysis
Accepted 26 April 2015
and used as reinforcing phase to produce CMC/ST-CNC bio-nanocomposite films at different CNC load-
Available online 30 April 2015
ing levels (0.5–5.0 wt%). Steady shear viscosity and dynamic viscoelastic measurements of film-forming
solution (FFS) of neat CMC, CMC/ST blend and CMC/ST-CNC bio-nanocomposites were evaluated. Vis-
Keywords:
cosity measurements revealed that a transition from Newtonian behavior to shear thinning occurred
Bio-nanocomposite
Cellulose nanocrystals when CNC were added. The dynamic tests confirmed that all FFS have a viscoelastic behavior with an
Rheology entanglement network structure, induced by the hydrogen bonding. In regard to the cast film quality, the
Solvent casting rheological data showed that all FFS were suitable for casting of films at ambient temperature. The effect
Optical transparency of CNC addition on the optical transparency, water vapor permeability (WVP) and tensile properties of
Tensile properties bio-nanocomposite films was studied. It was found that bio-nanocomposite films remain transparent
due to CNC dispersion at the nanoscale. The WVP was significantly reduced and the elastic modulus
and tensile strength were increased gradually with the addition of CNC. Herein, the steps to form new
eco-friendly bio-nanocomposite films were described by taking advantage of the combination of CMC, ST
and CNC. The as-produced films exhibited good optical transparency, reduced WVP and enhanced tensile
properties, which are the main properties required for packaging applications.
© 2015 Elsevier Ltd. All rights reserved.

1. Introduction applications, especially in the field of biomaterials and packaging


applications (Rhim, Park, & Ha, 2013). In this context, biopolymers
The use of biopolymers for sustainable development and envi- are considered potential replacements for conventional plastic
ronmental preservation has become ubiquitous in the last 20 years, materials; nevertheless, some of their properties must be improved
due to their several advantages including non-toxicity, biodegrad- to position them as potential candidates that can be competed with
ability, wide availability and biocompatibility, especially when fossil derivatives (Kanmani & Rhim, 2014), especially, mechanical
compared to their synthetic counterparts (Rodríguez-González and water vapor permeability properties, which are the properties
et al., 2012). Biopolymers such as proteins and carbohydrates are required for packaging applications (Rhim et al., 2013). How-
widely used as materials for both conventional and innovative ever, the blending of biopolymers and/or adding of nanofillers
represents an effective way to improve the properties of biopoly-
mers, and therefore, broadens their fields of application (Reddy,
∗ Corresponding authors. Tel.: +212 662010620. Vivekanandhan, Misra, Bhati, & Mohanty, 2013; Rhim et al., 2013).
E-mail addresses: abderrahim.solhy@um6p.ma (A. Solhy), Biopolymer blending is one of the most effective methods to
Mounir.elachaby@um6p.ma (M. El Achaby). create new biomaterials with desired properties. Films produced

http://dx.doi.org/10.1016/j.carbpol.2015.04.051
0144-8617/© 2015 Elsevier Ltd. All rights reserved.
N. El Miri et al. / Carbohydrate Polymers 129 (2015) 156–167 157

from the blending of biopolymers usually exhibit modified prop- bio-nanocomposite materials with high mechanical, optical, ther-
erties as compared to films made from an individual component mal, and barrier properties (Alves, dos Reis, Menezes, Pereira, &
(Ghanbarzadeh, Almasi, & Entezami, 2010; Sionkowska, 2011). Pereira, 2015; Azizi-Samir, Alloin, Sanchez, & Dufresne, 2004; Chen
Recently, there has been a noticeable increase in interest in the et al., 2009; El Miri et al., 2015; Fortunati et al., 2015, 2013; Kamal
use of biodegradable films made from modified biopolymers or & Khoshkava, 2015; Mariano, El Kissi, & Dufresne, 2014), which
biopolymer based blends for packaging applications. Because they are the main properties required for packaging applications (Rhim
offer many advantages over synthetic or non-biodegradable films et al., 2013). The adhesion properties and the large-number func-
for coating applications on foodstuff (BenBettaïeb, Karbowiak, tional groups that are present in the CNC’s surface and biopolymer
Bornaz, & Debeaufort, 2015; Ghanbarzadeh et al., 2010; Silva- matrices can be exploited to improve the interfacial interactions
Pereira, Teixeira, Pereira-Júnior, & Stefani, 2015; Van den Broek, between the CNC and biopolymer matrices (El Miri et al., 2015).
Knoop, Kappen, & Boeriu, 2015). This study examines bio-nanocomposite films based on CNC-
On the other hand, the bio-nanocomposite technology with low filled CMC/ST biopolymer blend. The CNC were extracted from
loading of nanofillers (5 wt%) has already been proven as an effec- sugarcane bagasse (SCB) via sulfuric acid hydrolysis process and
tive way to produce new biomaterials with specific properties and successfully characterized. Then, the as-extracted CNC were dis-
high performances for packaging applications (Reddy et al., 2013; persed into CMC/ST (70/30, w/w) biopolymer blend in order to
Rhim et al., 2013). The inclusion of well-dispersed nanofillers into fabricate CMC/ST-CNC bio-nanocomposite films using the solvent
a polymeric matrix may result in improving the physicochemi- casting method, which is the frequently used method for processing
cal properties of the resulted bio-nanocomposites, especially the of biopolymer blend and bio-nanocomposite films for packag-
mechanical, optical, thermal and barrier properties. ing applications. CMC and ST are biopolymers that have hydroxyl
Among the most studied biopolymers for packaging applications and carboxyl groups in their macromolecular chains; these groups
are the starch (ST) and cellulose derivatives, such as carboxymethyl could be useful to link the functional groups of CNC with CMC and ST
cellulose (CMC) (Rhim et al., 2013). Despite their good film-forming macromolecules. The rheological properties of film-forming solu-
abilities, films made from CMC or ST have poor mechanical and tions (FFS) of CMC/ST blend and CMC/ST-CNC bio-nanocomposites
water vapor permeability properties, but these can be enhanced were studied and their relation with the casting process was
by mixing them together to produce biodegradable blended films evaluated. Structural, optical transparency, WVP and mechanical
with modified properties (Ghanbarzadeh et al., 2010; Peressini, properties of the prepared bio-nanocomposite films were evalu-
Bravin, Lapasin, Rizzotti & Sensidoni, 2003; Tongdeesoontorn, ated and studied in this report. The high performances of these
Mauer, Wongruong, Sriburi, & Rachtanapun, 2011) or by adding bio-nanocomposite films are expected to have potential in bioma-
nanofillers to produce CMC or ST based bio-nanocomposite films terials or packaging applications.
with improved properties (Chen, Liu, Chang, Cao, & Anderson, 2009;
Oun & Rhim, 2015; Savadekar, Karande, Vigneshwaran, Kadam, &
2. Materials and experimental details
Mhaske, 2014; Yadollahi, Namazi, & Barkhordari, 2014).
The CMC biopolymer contains hydroxyl and carbonyl groups
2.1. Materials
and it is potentially miscible with ST biopolymer, due to
the formation of hydrogen bonds (Ghanbarzadeh et al., 2010;
The SCB was provided by a SUNABEL-COSUMAR Group, a
Rodríguez-González et al., 2012; Tongdeesoontorn et al., 2011).
company located in the region of Gharb-Loukkos, Morocco. The
Indeed, the combination of their properties allows the forma-
moisture content of the raw SCB was about 7 wt%. The car-
tion of a CMC/CST blend that may lead to the preparation of
boxymethyl cellulose sodium salt powder (CAS 9004-32-4) was
a new biocompatible and a homogeneous blend matrix for bio-
purchased from Alfa Aesar and corn starch (CAS 9005-25-8) was
nanocomposite development (Almasi, Ghanbarzadeh, & Entezami,
purchased from Sigma–Aldrich. Analytical grade chemicals used
2010; Rodríguez-González et al., 2012). Ghanbarzadeh et al.
for CNC extraction were purchased from Sigma–Aldrich and used
(2010) studied CMC/ST blend films, and the effect of CMC con-
without further purification.
tent (0–20 wt%) on the physicochemical properties of the resulted
CMC/ST blend films was evaluated. It was found that the CMC/ST
blend films showed better mechanical and barrier properties than 2.2. Isolation of CNC
neat ST film.
Accordingly, the current research is focused on the use CNC were extracted from SCB through three well-known steps:
of CMC/ST blend as a biopolymer matrix for developing bio- alkali and bleaching treatments and sulfuric acid hydrolysis,
nanocomposite films with improved water vapor permeability according to the procedure reported by Siqueira et al. (2013).
(WVP) and mechanical properties. Recently, Almasi et al. (2010) Briefly, the raw SCB fibers were washed with distilled water for 1 h
incorporated nanoclay within a CMC/ST blend to produce high per- at 60 ◦ C under mechanical stirring. Then, the prewashed SCB fibers
formance nanoclay filled CMC/ST bio-nanocomposite films with were treated three times with 4 wt% NaOH solution at 80 ◦ C for 2 h
improved properties. In the same context, Rodríguez-González under stirring and bleached with a solution made by equal parts
et al. (2012) developed bio-nanocomposites of CMC/ST blend of acetate buffer, aqueous chlorite (1.7 wt% in water) and distilled
reinforced with graphene oxide and keratin-grafted graphene water. The bleaching treatment was performed four times at 80 ◦ C,
oxide. From their results, the thermomechanical properties of the under mechanical stirring, during 4 h for each one. Acid hydrol-
blends were drastically enhanced with addition of modified and ysis was achieved at 55 ◦ C with 64 wt% sulfuric acid for 30 min
unmodified graphene oxide. Thus, with an aim to transform the under mechanical stirring. The suspension was diluted with ice
biopolymers into materials with high structural and mechanical cubes to stop the reaction and was washed by successive centrifu-
properties that can be competitive with the fossil derivatives, the gations at 12,000 rpm at 15 ◦ C for 30 min at each step and dialyzed
researches on biopolymer blend based bio-nanocomposites have against distillated water until neutral pH. Afterward, the obtained
gained considerable interest in the recent years. CNC suspension was sonicated (Branson sonifier model 450) for
Recently, the use of elongated cellulose nanocrystals (CNC) for 5 min in an ice bath to avoid overheating, which can cause desul-
bio-nanocomposites development has attracted more and more fation of the sulfate groups on the surface of cellulose. Then, the
attention in the field of nanotechnology. It has been widely demon- obtained sonicated suspension was freeze-dried (Christ Alpha 2-4
strated that the incorporation of CNC into biopolymer can result in LD Freeze Dryer) to remove water. The freeze drying process was
158 N. El Miri et al. / Carbohydrate Polymers 129 (2015) 156–167

used to obtain the CNC in solid form in order to measure their mass 2.4.2. Rheological measurements of FFS
fractions with accuracy for bio-nanocomposites preparation. Steady state flow and dynamic viscoelasticity measurements of
nanocomposites’ FFS were performed at 20 ◦ C, using a rotational
Physica MCR500 rheomter equipped with concentric cylinder
2.3. Preparation of bio-nanocomposite films geometry (CC27). The temperature was regulated by a PaarPhys-
ica circulating bath and a controlled Peltier system (TEZ 150P-C).
The CMC/ST-CNC bio-nanocomposite films were prepared by Firstly, steady shear flow measurement from 0.01–100 s−1 was per-
using the solvent casting method (Table 1). Firstly, the CMC solu- formed and flow curves of all FFS were obtained. Secondly, strain
tion was prepared by mixing the desired amount of CMC powder sweep measurement was performed, at a frequency of 0.75 Hz,
with 350 mL of distilled water and kept under vigorous stirring for in order to determine the appropriate strain of linear-viscoelastic
2 h at ambient temperature. Afterwards, the ST solution was sepa- region of the studied FFS. The storage or elastic modulus (G ) and
rately prepared by dissolving the desired amount of ST in 65 mL of loss or viscous modulus (G ) as a function of strain were mea-
distilled water and kept under stirring for 30 min at 90 ◦ C. During sured. Finally, frequency sweep was carried out over a range of
the heating process, 0.5 g of glycerol was added to the ST solution 0.01–100 Hz at a strain of 1%. Dynamic G and G moduli as a func-
as plasticizer. The CMC/ST mass ratio was fixed at 70/30 in accor- tion of frequency were measured.
dance to the work published by Rodríguez-González et al. (2012).
The aqueous suspension of CNC was prepared separately by mix-
2.4.3. FTIR analysis of films
ing the desired amount of freeze dried CNC with 65 mL of distilled
Transform Infrared Spectroscopy (FTIR) characterizations of bio-
water, followed by ultrasonic treatment (Branson sonifier model
nanocomposite films were performed on an ABB Bomem FTLA 2000
450) for 30 min in an ice bath in order to homogenize the disper-
spectrometer equipped with a Golden Gate single attenuated total
sion of CNC in water. The obtained ST solution and CNC suspension
reflection (ATR) cell. The experiments were carried out in the range
were then mixed and stirred for 30 min at ambient temperature.
from 4000 to 400 cm−1 with a 4 cm−1 resolution and an accumula-
Next, the resulting ST/CNC mixture was added to the pre-prepared
tion of 32 scans. The FTIR spectra were taken in the transmittance
CMC solution and stirred for 1 h at 25 ◦ C. Finally, the CMC/ST-CNC
mode.
mixture was casted onto Petri dishes and the water evaporated
at ambient temperature for 3 days, and the obtained films were
dried for 3 h at 80 ◦ C to completely remove the water. Neat CMC 2.4.4. Morphology of films
and CMC/ST blend films were also prepared according to the same Morphology of bio-nanocomposite films was evaluated at their
process mentioned above, without the addition of CNC. The load- cryo-fractured surfaces, using scanning electron microscope (SEM)
ing levels of CNC in CMC/ST-CNC bio-nanocomposite films were (FEI, Quanta 200-ESEM) operated at 20 kV. The samples were frozen
fixed at 0.5, 2.5 and 5.0 wt%. The films were coded as CMC, CMC/ST, in liquid nitrogen and cryo-fractured before being coated by a thin
CMC/ST-CNC-0.5, CMC/ST-CNC-2.5 and CMC/ST-CNC-5.0. The num- conductive carbon layer to help improve SEM observations.
ber in each code name indicates the weight fraction of CNC in each
formulation. Film codes and details of the formulations of each film 2.4.5. UV–vis spectroscopy analysis of films
are shown in Table 1. The UV–vis spectroscopy analysis of bio-nanocomposite films
was carried out using a PerkinElmer LAMBDA 1050 UV/vis/NIR
spectrophotometer. The film samples were placed directly in the
2.4. Characterization techniques spectrophotometer test cell and the air was used as reference. The
optical absorbance/transmittance of the films were measured in
2.4.1. Characterization of CNC the wavelength range of 200–800 nm.
Atomic Force Microscopy (AFM) measurements of CNC were
carried out using a Veeco Dimension ICON. The tapping mode was
used to capture height images at a scan rate of 1.5 Hz. Samples 2.4.6. Tensile properties of films
for AFM observations were prepared by depositing a droplet of Tensile tests were performed using an Instron 8821S tensiome-
a diluted CNC suspension onto freshly cleaved mica sheets, after ter according to our previous work (El Achaby et al., 2014). The
being sonicated for 30 min, and allowing the solvent to dry in air. tensile specimens were cut in rectangular shapes with dimensions
The zeta potential measurement of CNC suspension was carried out of 80 mm in length and 10 mm in width. The gauge length was fixed
using a Malvern Zetasizer Nano ZS instrument. For this measure- at 30 mm and the cross head speed was 5 mm/min at room tem-
ment, capillary cell was used and the CNC suspension was diluted perature. All tests were carried out on a minimum of five samples
and sonicated for 30 min before being analyzed. X-ray diffraction and the reported results are average values.
(XRD) characterizations of freeze-dried CNC were carried out using
a Bruker D8 Discover using the Cu K␣ radiation ( = 1.54184 nm) in 2.4.7. Water vapor permeability (WVP) of films
the 2 range of 2–60◦ , while the voltage and current were held at The determination of the WVP of the CMC/ST-CNC bio-
45 kV and 100 mA. nanocomposite films was carried out according to the standard
method E96-E95 (ASTM method 1995) with some modifications
(Ghanbarzadeh et al., 2010). Glass bottles, with diameter of 20 mm
Table 1 and depth of 45 mm, were used to accomplish the test. After posi-
Sample codes and the composition of each formulation of films. tioning 4 g of anhydrous CaCl2 inside each glass bottle and thus
for maintaining a relative humidity (RH) of 0%. The glass bottles
Film code CMC ST CNC Total mass
(g)/H2 O (g)/H2 O (g)/H2 O (g)/H2 O (mL) were covered with the elaborated films, weighted and inserted in
(mL) (mL) (ml) climatic chamber with RH of 50% and temperature of 32 ◦ C. The
CMC 5.000/350 0.000/65 0.000/65 5/500 (1%, w/v)
weighting process was made each hour over the duration of 7 h. The
CMC/ST 3.500/350 1.500/65 0.000/65 5/500 (1%, w/v) changes in the weight of the glass bottle were recorded as a function
CMC/ST-CNC-0.5 3.482/350 1.492/65 0.025/65 5/500 (1%, w/v) of time. Slopes were calculated by linear regression (weight change
CMC/ST-CNC-0.5 3.412/350 1.463/65 0.125/65 5/500 (1%, w/v) versus time). The water vapor transmission rate (WVTR) was deter-
CMC/ST-CNC-0.5 3.325/350 1.425/65 0.250/65 5/500 (1%, w/v)
mined as the slope of the straight line (g/h) divided by the area of
N. El Miri et al. / Carbohydrate Polymers 129 (2015) 156–167 159

the glass bottle mouth (m2 ). Finally, the WVP (g m/m2 h Pa) was alkali-labile linkages (ether and ester linkages) between lignin
calculated as follows: monomers or between lignin and polysaccharides may have been
WVTR WVTR broken, which helped in the partial defibrillation of fibers. This
WVP = X= X treatment yielded about 45% yellow colored fibers (Fig. 1b), which
p S(R1 − R2)
was calculated by the weight of the treated fibers divided by the ini-
where X is the thickness of the film (m), S is the saturation vapor tial weight of SCB. The next step was the bleaching treatment, with
pressure (Pa) at the test temperature (32 ◦ C), R1 and R2 are the the aim to purify the cellulosic fibers from any residual lignin or
relative humidity in the climatic chamber and in the glass bottle, other impurities obtained after the alkaline treatment. This bleach-
respectively. ing treatment can result in the total defibrillation of fibers into
individual microfibrils with smaller diameter. The bleached fibers
3. Results and discussions may have contained pure cellulose, as shown by a clearly white
color (Fig. 1c), thus confirming that the residual non-cellulosic ele-
3.1. Extraction and characterization of CNC ments were totally removed under the bleaching treatment. The
yield of this bleached product was about 33% with respect to the
The CNC used as nano-reinforcing agents were extracted from initial amount of SCB. This yield is close to the values reported
SCB via alkali and bleaching treatments followed by acid hydrol- in the literature for treated rice straw (36%) (Lu & Hsieh, 2012)
ysis. Visually, the initial ground SCB fibers had a yellow-brown and treated pineapple leaf (40%) (Dos Santos et al., 2013). The
color (Fig. 1a). Beginning with the alkali treatment of the raw third and final step was the acid hydrolysis, which was carried out
SCB fibers, the main focus of this step was to eliminate the whole under appropriate conditions that allowed the removal of amor-
non-cellulosic components (lignin, hemicelluloses and pectin) (Dos phous domains from the bleached cellulose. The acid hydrolysis
Santos et al., 2013). With the subsequent treatment of alkali, some and removal of amorphous domains was performed by cleaving

Fig. 1. Digital images of (a) raw SCB, (b) alkali-treated SCB, (c) bleached SCB and (d) freeze dried CNC samples; and (e) CNC suspension (7.6 mg/ml) and (f and g) high and
low scale AFM images of CNC and (h) XRD pattern of CNC.
160 N. El Miri et al. / Carbohydrate Polymers 129 (2015) 156–167

cellulose microfibrils into bundles of CNC with nanometric dimen- peaks observed at 14–16◦ and 34.5◦ are also characteristic of cel-
sions, becoming more or less individualized by sonication with the lulose I (Flauzino-Neto et al., 2013). The crystallinity index (CrI)
ability to be obtained in a clearly white powdered form via a freeze- of CNC was found to be 65% according to calculations using the
drying process (Fig. 1d). The yield of CNC was 15% with respect to equation CrI = ((I0 0 2 − Iamorph )/I0 0 2 ) × 100, where I0 0 2 and Iamorph
the initial amount of dried SCB fibers, and 45% with respect to the are the peak intensities of crystalline and amorphous cellulose,
amount of bleached cellulose. Comparatively, this yield is similar respectively (Flauzino-Neto et al., 2013; Teixeira et al., 2011).
to the CNC derived from corncob (46–50%) (Silvério, Flauzino-Neto,
Dantas, & Pasquini, 2013), higher than that obtained from rice straw 3.2. Stability of film-forming solutions
(13–18%) (Lu & Hsieh, 2012) and smaller than that obtained for CNC
extracted from pineapple leaf (55–77%) (Dos Santos et al., 2013), Owing to the hydrophilic nature of CMC, ST and CNC, their
and SCB (50–58%) (Teixeira et al., 2011). mixture in water can be easily achieved in controlled conditions,
The stability of the CNC suspension is crucial in the preparation enabling the formation of a homogeneous and stable aqueous mix-
of nanocomposites and can be derived from the zeta potential of the ture. Fig. 2a shows the photographs of neat CMC, CMC/ST blend
CNC suspension. It is well-known that the stability of CNC depends and CMC/ST-CNC bio-nanocomposites FFS. The photographs were
strongly on its aspect ratio, surface functionalization, and the ability recorded at room temperature 15 days after preparation of the
of the solvent and surface groups to counterbalance the attrac- mixtures, showing high stability and good homogeneity. This is
tive hydrogen-bond interactions exerted by the abundant hydroxyl attributed to the good compatibility and the strong interactions
groups (Espinosa, Kuhnt, Foster, & Weder, 2013). Fig. 1e shows a between CMC and ST in the blend and between CMC, ST and CNC
photograph illustrating the CNC aqueous suspension stability, in in the bio-nanocomposites. CMC and ST are miscible polymers via
which a white gel appearance is observed. To note, this suspension hydrogen bonding interactions, because the hydroxyl groups of ST
was prepared by redispersing of freeze dried CNC in water via ultra- are able to form hydrogen bonds with the hydroxyl and carboxyl
sonic treatment. Some suggested reasons for good stability of CNC groups of the CMC macromolecules (Ghanbarzadeh et al., 2010;
suspension is the exclusion of apolar components, the insertion of Rodríguez-González et al., 2012; Tongdeesoontorn et al., 2011).
polar sulfate groups, and the exposition of OH groups from the Very interestingly, the sulfuric acid hydrolyzed CNC exhibited free
cellulose structure. For our CNC suspension, the zeta potential had hydroxyl and inserted anionic sulfate groups on their surfaces,
a mean value of −61.9 mV. The suspension of CNC was considered which can also strongly interact with the hydroxyl groups and the
stable because the absolute value was higher than 25 mV (Morais carboxyl groups of CMC and ST. In this way, when CNC were incor-
et al., 2013). porated into CMC/ST polymeric blend, a fine dispersion of CNC can
The morphology and size of the as-obtained CNC was char- be achieved.
acterized by AFM analysis; the micrographs are shown in Fig. 1f
and g. To note, the AFM is an appropriate method to characterize 3.3. Steady shear viscosity of FFS
CNC because it provides accurate information about the diame-
ter, length and shape of the CNC (Flauzino-Neto, Silvério, Dantas, Fig. 2b shows the dependence of the steady shear viscosity
& Pasquini, 2013). The AFM images showed that the as-isolated () on the shear rate ()˙ for FFS of neat CMC, CMC/ST blend and
CNC had needle-like shape, confirming that their extraction from CMC/ST-CNC bio-nanocomposites at different CNC loading levels
the treated SCB was successful. These images showed individ- (0.5–5.0 wt%). As seen in Fig. 2b, neat CMC FFS exhibited a Newto-
ual CNC and some aggregates through transverse height profiles. nian behavior at low shear rates, which was characterized by the
The appearance of laterally aggregated CNC in AFM images was shear rate-independent viscosity. At high shear rates, the most pro-
expected due to the high specific area and strong hydrogen bonds nounced shear-thinning behavior was also observed for the neat
established between the CNC. Additionally, AFM images showed CMC FFS. This result indicates that as shear rate increases, the inter-
that CNC are uniform in diameter and irregular in length. In general, molecular junctions are disrupted at a rate faster than their rate
the exact dimensions of CNC depend on the nature of the original of reformation, resulting in a decrease in the junction density and
raw material and hydrolysis conditions or pretreatments (Eichhorn hence a drop in the viscosity. The same behavior was previously
et al., 2010). Herein, the average diameter of the as-extracted CNC observed for neat CMC FFS (Florjancic, Zupancic, & Zumer, 2002).
was 5 ± 1.1 nm and the average length was 275 ± 73 nm, giving rise These results indicate that the CMC FFS behaves as an entangle-
to an average aspect ratio of about 55. Comparatively, the mea- ment network system (or a concentrated solution) (Florjancic et al.,
sured aspect ratio was similar to nano-sized CNC derived from 2002; Kulicke, Kull, Kull, & Thielking, 1996). The CMC/ST blend
other sources such as SCB (32–64) (Teixeira et al., 2011), Corconob FFS showed the same behavior observed for neat CMC (Newto-
(52–63) (Silvério et al., 2013), sisal (43–60) (Garcia de Rodriguez, nian behavior at low shear rate and shear-thinning behavior at high
Thielemans, & Dufresne, 2006) and pineapple leaf (50–60) (Dos shear rate) with a slight increase in the shear viscosity at low shear
Santos et al., 2013); and higher than that measured for CNC from rate values; this is due to interfacial interactions between macro-
rice straw (9–10.5) (Lu & Hsieh, 2012), cotton (9.5–12.5) (Espinosa molecular chains of CMC and ST via the hydrogen bonding. The
et al., 2013) or cotton linters (20–24) (Morais et al., 2013), coconut decrease of the viscosity at high shear rates for CMC/ST blend may
husks (35–44) (Rosa et al., 2010), ramie (12) (De Menezes, Siqueira, be explained by the disruption of the formed interactions between
Curvelo, & Dufresne, 2009) and Luffa cylindrica fibers (46) (Siqueira CMC and CMC or CMC and ST macromolecular chains during the
et al., 2013). Contrastingly, it was smaller than that of CNC from rotational measurements.
tunicate (70) (Azizi-Samir et al., 2004) and (148) (Sacui et al., 2014) For FFS of CMC/ST-CNC bio-nanocomposites, it was observed
and bacteria (60–94) (Olsson et al., 2010; Sacui et al., 2014). Con- that the steady shear viscosity of CMC/ST blend was significantly
sequently, it has been reported in the literature that the efficient influenced by the addition of CNC. At the low shear rate region, the
reinforcement effect of CNC in nanocomposite materials can be viscosity increased gradually with the increase of CNC loading; this
guaranteed when CNC with an aspect ratio around 50 are incor- is due to the flow-impeding effect induced by the presence of CNC
porated into a polymer matrix (Eichhorn et al., 2010). (Wu, Wang, Zhang, & Zhou, 2012). Furthermore, the presence of
Fig. 1h shows the XRD pattern of CNC. The XRD patterns show CNC restrained the shear flow of the CMC and ST macromolecules
that the major intensity peak is located at a 2 value of around because the CNC’s average size is larger than that of the chain seg-
22.6◦ , which is related to the crystalline structure of cellulose I ments of the CMC and ST biopolymers, resulting in an increase of
(Teixeira et al., 2011). Additionally, it has been reported that the shear viscosity (Wu et al., 2012).
N. El Miri et al. / Carbohydrate Polymers 129 (2015) 156–167 161

Fig. 2. (a) Digital images, (b) shear viscosity versus shear rate and (c) zero shear viscosity of FFS based on (A) CMC, (B) CMC/ST, (C) CMC/ST-CNC-0.5, (D) CMC/ST-CNC-2.5 and
(E) CMC/ST-CNC-5.0.

As a result, the FFS of CMC/ST-CNC-0.5 bio-nanocomposite In relation to the casting film quality, FFS based on neat CMC,
showed a Newtonian behavior at very low shear rate; but the shear CMC/ST blend and CMC/ST-CNC bio-nanocomposites were homo-
thinning behavior became more pronounced than that observed geneous and without bubbles. Excellent stability without phase
for FFS of neat CMC and CMC/ST blend, because this trend is separation was observed in all FFS studied. It has been reported
still controlled by the CMC behavior (Florjancic et al., 2002). The that a high-viscosity FFS of biopolymers would make it difficult
FFS of CMC/ST-CNC-2.5 and CMC/ST-CNC-5.0 bio-nanocomposites to remove air bubbles and hinder casting in thin layers (Cuq,
showed only the shear thinning behavior in the full shear rate Aymard, Cuq, & Guilbert, 1995). According to Cuq et al. (1995), it
range, indicating that a network structure is formed by CNC within is very difficult to cast a consistent film from high-viscosity FFS
the bio-nanocomposite FFS, which is constructed by strong hydro- and the process needs a dispersing machine (spreader). Neverthe-
gen interactions and consequently causes the viscosity to increase less, the FFS must be characterized by a relatively low viscosity
(Zhou, Wang, & Wu, 2012). A similar trend was reported for poly in order to avoid retention of air bubbles. Silva-Weiss, Bifani, Ihl,
(ethylene oxide) solutions containing different contents of CNC, in Sobral, and Gómez-Guillén (2014) studied the viscosity of FFS
which the addition of CNC causes the transition from Newtonian based on CMC/ST and CMC/modified ST and observed a shear thin-
behavior observed in polymer solution, to shear thinning behavior ning behavior for these FFS. It was found that, at a shear rate of
in the full range of shear rate (Zhou et al., 2012). This transition was 2 s−1 , the apparent viscosity of CMC, CMC/ST and CMC/modified
attributed to the formation of a network structure of CNC within ST was 8.49, 7.02 and 4.77 Pa s−1 , respectively; these values were
the polymer solution (Bercea & Navard, 2000; Liu, Chen, Yue, Chen, found to be suitable for the casting process at ambient tempera-
& Wu, 2011; Zhou et al., 2012). ture. Moraes, Carvalho, Bittante, Solorza-Feria, and Sobral (2009)
The zero shear viscosity 0 was estimated by extrapolation demonstrated that the low viscosity with Newtonian behavior
along the low shear rate plateau, and the 0 values determined of FFS based on gelatin/polyvinyl alcohol facilitated the cast-
for the neat CMC, CMC/ST blend and CMC/ST-CNC FFS were plot- ing process. In this present study, the viscosity values observed
ted in Fig. 2c. The 0 showed a significant increase when CNC for all studied FFS were higher than those observed by Moraes
were added. Consequently, this 0 enhancement suggests that the et al. (2009) and smaller than that observed by Silva-Weiss et al.
CMC, ST and CNC are perfectly compatible phases. Therefore, a net- (2014). Consequently, the observed shear viscosity for all FFS
work structure was assumed to have been formed in CMC/ST-CNC was suitable for the casting process since uniform layers of FFS
bio-nanocomposite solutions (Zhou et al., 2012). The formation of were spread on the plate, covering the entire surface at room
this network could be related to the special nanometric rod-like temperature. The dried films were smooth, homogeneous, trans-
morphology of the CNC as well as their functionalized surfaces, pro- parent and easily peeled off from the plate surface. Additionally,
viding a well homogeneous dispersion of CNC within the polymer these films were optically transparent, as confirmed by UV–vis
matrix, and thus leading to a high contact area at the CNC-polymer absorbance/transmittance measurements and visual observations
chains interface. (see below).
162 N. El Miri et al. / Carbohydrate Polymers 129 (2015) 156–167

Fig. 3. (a and b) Storage (G ) and loss (G ) moduli versus strain and (c and d) G and G versus frequency of FFS based on neat CMC, CMC/ST blend and CMC/ST-CNC
bio-nanocomposites.

3.4. Viscoelastic behavior of FFS the increase of CNC loading from 0.5 to 5.0 wt%. A  crit of 17.5%
was observed for CMC/ST-CNC-5.0 bio-nanocomposite FFS, indi-
To help select the strain appropriate for the linear viscoelas- cating that the structure of the bio-nanocomposite FFS containing
tic behavior of the studied FFS, dynamic strain sweeps were a high CNC loading level started to destroy rapidly in compar-
first collected for all FFS at 0.75 Hz and 20 ◦ C. Fig. 3a and b ison with that of neat CMC and CMC/ST FFS. The decreasing of
shows the strain dependence of dynamic storage (G ) and loss  crit with the addition of CNC was recently observed by Kamal
(G ) moduli for neat CMC, CMC/ST and CMC/ST-CNC FFS. The and Khoshkava (2015) in CNC filled PLA bio-nanocomposite sys-
behavior illustrated in these Figures is typical of many viscoelas- tems. It is also noted that below  crit , all the studied FFS exhibited
tic materials: the dynamic moduli remain constant in the linear G higher than G , indicating that the viscous behavior is pre-
range and decreases above a critical shear strain ( crit ) (Kamal & dominant over the elastic one in the linear viscoelastic range
Khoshkava, 2015; Silva-Weiss et al., 2014). From Fig. 3a and b, (at low frequency) (Table 2). Consequently, to avoid structural
the  crit , G and G at  crit are extracted and presented in Table 2. breakdown, a strain of 1% was chosen as an appropriate strain to
For the neat CMC and CMC/ST blend FFS, both moduli (G and study the linear viscoelastic behaviors in all studied FFS. Dynamic
G ) started to decrease when the strain exceeded  crit = 63.94%, frequency sweeps with an applied strain level of 1% were then
indicating that the material structure starts to destroy. In CMC/ST- collected. The storage modulus (G ) and loss modulus (G ) were
CNC FFS, G and G increased monotonously with the increase measured as a function of frequency in the linear viscoelastic
of CNC loading, and the  crit decreased more dramatically with region.

Table 2
Viscoelastic parameters of FFS of neat CMC, CMC/ST blend and CMC/ST-CNC bio-nanocomposites.

Film code Strain sweep Frequency sweep

 crit (%) G at  crit (Pa) G at  crit (Pa) G at f = 0.1 Hz (Pa) G at f = 0.1 Hz (Pa) G = G (Pa) f at G = G (Hz)

CMC 63.94 0.70 1.07 0.01 0.13 1.35 1.39


CMC/ST 50.22 0.76 1.14 0.02 0.15 1.60 1.53
CMC/ST-CNC-0.5 31.36 0.93 1.58 0.03 0.21 2.38 1.58
CMC/ST-CNC-2.5 20.03 1.27 1.83 0.10 0.32 2.42 1.48
CMC-ST-CNC-5.0 17.52 1.99 2.64 0.24 0.59 2.91 1.33

 crit : critical strain from G curves; f: frequency.


N. El Miri et al. / Carbohydrate Polymers 129 (2015) 156–167 163

Dynamic frequency sweep results can be used to characterize


or classify polymer based solutions (Clark & Ross-Murphy, 1987).
The four most traditional classifications are: a dilute solution, an
entanglement network system, a weak gel and a strong gel. A dilute
solution shows loss modulus (G ) larger than storage modulus (G )
over the entire frequency range, yet the moduli approach each other
at higher frequencies (Clark & Ross-Murphy, 1987). The entangle-
ment network system is characterized by the intersecting of both
moduli (G = G ) at the middle of the frequency range, indicating a
clear tendency for more solid like behavior at higher frequencies
(Clark & Ross-Murphy, 1987; Silva-Weiss et al., 2014; Xu, Liu, &
Zhang, 2006).
According to the G and G curves presented in Fig. 3c and
d, it is clear that all FFS show predominantly viscous behavior
at low frequencies (G < G ) and elastic behavior at high frequen-
cies (G > G ) (Table 2). According to the classifications mentioned
above, all FFS can be classified as an entanglement network system,
with the intersection of G and G (G = G ) at a frequency of about
1.3 Hz. The intersection points and the corresponding frequency
are summarized in Table 2. These results strongly align with the
results of Florjancic et al. (2002), Kulicke et al. (1996), Peressini
et al. (2003) and Silva-Weiss et al. (2014). At low frequencies,
there was enough time for entanglement to occur and be disrupted
during the oscillation period, which allowed viscous flow to pre-
dominate over the elastic behavior (G < G ). With an increase in
frequency, the time became insufficient for disruption of the entan-
glement; and the dispersion behaved as a cross-linked network
with the elastic response predominating over viscous flow (G > G )
(Ghasemlou, Khodaiyan, & Oromiehie, 2011; Peressini et al., 2003).
Silva-Weiss et al. (2014) studied the dynamic rheological proper-
ties of a series of polysaccharides FFS. They showed that neat CMC
and CMC/ST exhibited an entanglement network system caused
by physical cross-linking through polymer–polymer interactions.
The same trend was previously observed in film-forming disper-
sions of ST-methylcellulose (Peressini et al., 2003). It has been
reported that the elastic behavior is a function of the number of
effective chains participating in the formation of a network struc-
ture (Peressini et al., 2003). Herein, the increase in G values with
the addition of ST to CMC solution, or with the dispersion of CNC Fig. 4. SEM micrographs of (a) CMC/ST, (b) CMC/ST-CNC-0.5, (c) CMC/ST-CNC-2.5
into CMC/ST solution, suggested the presence of a greater num- and (d) CMC/ST-CNC-5.0 films and (e) FTIR spectra of neat CMC, CMC/ST and CMC/ST-
CNC films.
ber of interacting chains than the neat CMC solution. In the FFS of
neat CMC and CMC/ST blend, the entangled network was caused
by inter-chain interactions between CMC-CMC and CMC-ST. How- directly correlated with its effectiveness in improving the proper-
ever, in the FFS of CMC/ST-CNC bio-nanocomposite, the dispersion ties of bio-nanocomposite films.
of CNC can cause additional interfacial interactions between their Fig. 4e shows the FTIR spectra of neat CMC, CMC/ST blend and
functionalized surfaces and the macromolecular chains of CMC and CMC/ST-CNC bio-nanocomposite films. For neat CMC film spec-
ST. Alternatively, the increase of G and G with the increase of fre- trum, the band at 1055 cm−1 can be associated to C O stretching
quency when CNC were added, suggests that the CNC were well vibration of ether groups, while the bands at 1434 and 1604 cm−1
dispersed at the nanometric scale within the CMC/ST blend solu- can be attributed to the asymmetric and symmetric modes of
tion. stretching vibration of carboxylate groups (C O), respectively, and
the broad band centered at 3338 cm−1 is attributed to the stretching
of OH groups and intermolecular and intramolecular hydrogen
3.5. FTIR and SEM analysis of films bonds (Layek, Kundu, & Nandi, 2013; Luna-Martínez et al., 2011).
Owing to the chemical similarities between CMC and ST and
The SEM photographs of cryo-fractured sections of CMC/ST the relatively low content (5.0 wt%) of CNC within the CMC/ST
blend and CMC/ST-CNC bio-nanocomposite films are shown in blend matrix, the FTIR spectra of CMC/ST blend and CMC/ST-CNC
Fig. 4a–d. It was observed that the cross-section of unloaded bio-nanocomposite films showed similar characteristics of neat
CMC/ST film was smooth and without any cracks or pores (Fig. 4a), CMC film. By adding ST biopolymer and CNC nanofillers to CMC
which revealed a better film homogenization of CMC/ST blend. film, the band observed at 3338 cm−1 in neat CMC spectrum ( OH
Comparatively, the cross-section of the CMC/ST film was not group) was shifted to lower frequencies (3293 cm−1 in CMC/ST
strongly affected by the addition of CNC (Fig. 4b–d). In the bio- and 3299–3287 cm−1 in CMC/ST-CNC). In the same way, the band
nanocomposite films, polymers chains encapsulated the CNC. This observed at 1604 cm−1 in neat CMC spectrum ( C O groups) was
indicates that CNC were dispersed uniformly within the CMC/ST shifted to lower frequencies (1592 cm−1 in CMC/ST and 1587 cm−1
blend, and thus formed a stronger interaction and adhesion on the in CMC/ST-CNC). These results were attributed to the formation of
interfaces between CNC and macromolecular chains of CMC/ST. The hydrogen bonds between hydroxyl groups in ST and hydroxyl and
excellent dispersion of CNC within the polymeric blend matrix was carboxyl groups in CMC, and the abundant hydroxyl groups that
164 N. El Miri et al. / Carbohydrate Polymers 129 (2015) 156–167

are present in the surface of CNC. These trends have been previ- to scatter and/or absorb the light and to increase the absorbance
ously observed in CMC/ST blend (Tongdeesoontorn et al., 2011), level of the films. However, no scattering and/or absorbance dif-
CNC-filled ST bio-nanocomposite films (Chen et al., 2009) and ference were found between the CMC/ST blend and CMC/ST-CNC
nanoclay-filled CMC/ST blend bio-nanocomposites (Almasi et al., bio-nanocomposites, indicating that the CNC were well dispersed
2010). at the nanometric scale within the CMC/ST polymeric blend (El
Achaby et al., 2012).
3.6. Transparency properties of films Fig. 5c shows the optical transparency of the studied films,
which was performed by acquiring transmittance spectra in the
The optical absorbance, transparency and local dispersion of wavelength region of 200–800 nm. It is clear that all studied films
CNC for the films of neat CMC, CMC/ST blend and CMC/ST-CNC exhibited optical transmittance between 95% and 80% under the
bio-nanocomposites were investigated using UV–vis spectroscopy. visible light region. It is well documented that the neat CMC
The UV–vis absorbance spectra of such films are presented in film exhibits a high transparency level (>90%) in the visible light
Fig. 5a. The absorbance at an arbitrarily chosen wavelength of region (Luna-Martínez et al., 2011; Yadollahi et al., 2014). This
600 nm was also extracted and plotted for all studied films (Fig. 5b). transparency level of CMC was not largely affected by its blend-
The absorbance of all films was dominated by the Raleigh scat- ing with ST, and the resulting CMC/ST blend film exhibited the
tering component (the 1/4 tail), characterized by a decrease of same transparency level with a very slight decrease of optical
the absorbance with the increase of the wavelength (El Achaby, transmittance. Importantly, the CMC/ST-CNC bio-nanocomposite
Arrakhiz, Vaudreuil, Essassi, & Qaiss, 2012). The absorbance level of films maintained the transparency properties of the CMC/ST
neat CMC film was slightly affected by its blending with ST, because blend matrix, confirming that CNC were well dispersed in bio-
the neat ST film showed a relatively high level in its absorbance nanocomposite films (Fortunati et al., 2015, 2013). It also confirmed
spectra (not shown) with respect to neat CMC film, resulting in that CNC have a good compatibility with the CMC/ST polymeric
increased absorbance for the CMC/ST blend film, in agreement with blend matrix to avoid CNC aggregation, thus reducing the amount
the observation of Silva-Weiss et al. (2014) for the CMC/ST blend of light scattering and favoring the transmittance of visible light
film. Furthermore, when CNC were dispersed into CMC/ST blend, all through the bio-nanocomposite films. Indeed, a good transparency
the investigated CMC/ST-CNC bio-nanocomposite films maintained for bio-nanocomposite films is an important property required for
the same absorbance level of the CMC/ST blend film, as can be seen packaging applications (Rhim et al., 2013).
in Fig. 5b for a wavelength of 600 nm, in which the absorbance The results obtained from UV–vis absorbance/transmittance
level is not largely affected by the addition of CNC. The average were also supported by the visual observations of the bio-
size of CNC was about 275 nm in length, as observed by AFM anal- nanocomposite films (Fig. 5d). The images illustrated in this figure
ysis (Fig. 1); its presence in bio-nanocomposite films is responsible clearly show that the transparent level of CMC/ST blend film was

Fig. 5. (a) UV–vis absorbance spectra, (b) plot of absorbance at 600 nm, (c) UV–vis transmittance and (d) digitals images of neat CMC, CMC/ST, CMC/ST-CNC films.
N. El Miri et al. / Carbohydrate Polymers 129 (2015) 156–167 165

140
15.0
13.5 120
WVP (10 .g.m/m².h.Pa)

12.0 CMC/ST-CNC-5.0

10.5 100 CMC/ST-CNC-2.5

Stress (MPa)
9.0 CMC/ST-CNC-0.5
80
CMC
7.5
-7

CMC/ST
6.0 60
4.5 40
3.0
1.5 20
0.0
0
A B C D E
0 5 10 15 20 25 30 35 40
Fig. 6. WVP of (A) neat CMC, (B) CMC/ST blend, (C) CMC/ST-CNC-0.5, (D) CMC/ST- Strain (%)
CNC-2.5 and (E) CMC/ST-CNC-5.0.
Fig. 7. Typical stress-strain curves of neat CMC, CMC/ST blend and CMC/ST-CNC
bio-nanocomposite films.
not affected by the addition of CNC due to its dispersion at nano-
metric scale (Fortunati et al., 2013, 2015). and the polymer matrix are the key points needed to achieve poly-
mer nanocomposite films with enhanced final properties (Habibi,
3.7. WVP of films Lucia, & Rojas, 2010). Herein, the tensile properties of the neat
CMC, CMC/ST blend and CMC/ST-CNC films were investigated by
In packaging applications, film should to avoid or at least uni-axial tensile tests and the obtained stress-strain curves are pre-
decrease moisture transfer between food and surrounding atmo- sented in Fig. 7. From the stress-strain curves, the elastic modulus
sphere; the WVP should be as low as possible. The WVP of neat (Young’s modulus), tensile strength and elongation at break of the
CMC, CMC/ST blend and CMC/ST-CNC bio-nanocomposite films films were extracted and presented in Table 3.
are shown in Fig. 6. The WVP of the neat CMC film was found According to these results, it is clear that the selected tensile
to be 13.6 × 10−7 g m/m2 h Pa and it was influenced by blending properties of CMC were affected by its blending with ST and then by
with the ST biopolymer and then by adding of CNC. This WVP the addition of different contents of CNC. The neat CMC film exhib-
value is smaller than that observed by Oun and Rhim (2015) ited an elastic modulus of 1066.16 MPa, a tensile stress of 76.23 MPa
(50.4 ×10−7 g m/m2 h Pa) and comparable to that observed by and an elongation at break of 22.99%. These obtained values are the
Yadollahi et al. (2014) (15.2 × 10−7 g m/m2 h Pa) for neat CMC film. characteristics of high molecular weight CMC biopolymer (Layek
It is well known that the polysaccharides are highly sensitive to et al., 2013). When CMC was blended with ST, the elastic modulus
moisture and exhibit poor water vapor barrier properties due to and tensile strength were decreased from 1066.16 to 847.23 MPa
their hydrophilic character (BenBettaïeb et al., 2015). Importantly, and from 76.23 to 66.82 MPa, respectively. The elongation at break
when CMC is blended with ST, the WVP decreased significantly to of CMC (22.99%) was not affected in the CMC/ST blend (23.88%)
6.05 × 10−7 g m/m2 h Pa, this is might be explained by the interac- (Table 3). This variation in the tensile properties of CMC was not
tion between both biopolymers via hydrogen bonding, which lead surprising because of the lower tensile properties of ST biopolymer
to decrease of the number of free hydroxyl groups that give the with respect to the CMC biopolymer. These results agree with those
hydrophilic character to biopolymers, and thus decrease of WVP of observed by Ghanbarzadeh et al. (2010) and Tongdeesoontorn et al.
the CMC/ST blend. This result is in agreement with results previ- (2011) in term of the variation of tensile properties of CMC/ST com-
ously reported for polymer blends which are studied for packaging pared to those of neat CMC.
applications (BenBettaïeb et al., 2015; Ghanbarzadeh et al., 2010). With respect to the CMC/ST blend film, the elastic modulus
On the other hand, the addition of CNC into CMC/ST blend showed a and tensile strength of the CMC/ST-CNC bio-nanocomposite films
positive effect on the WVP of the resulted bio-nanocomposite films. increased gradually with increasing of CNC content from 0.5 to
The WVP of bio-nanocomposite films decreased to 4.22 × 10−7 , 5.0 wt%. When 5.0 wt% CNC was added (CMC/ST-CNC-5.0), the elas-
4.01 × 10−7 and 5.24 × 10−7 g m/m2 h Pa when 0.5, 2.5 and 5.0 wt% tic modulus and tensile strength were increased from 847.23 to
CNC were added, respectively. 1650.19 MPa (94.77%), and from 66.82 to 110.83 MPa (65.86%),
From these results, the decrease in the WVP was attributed respectively (Table 3). Such improvements confirmed that CMC/ST-
to the impermeable CNC which were well dispersed/distributed CNC bio-nanocomposite films are mechanically strong materials.
through the blend matrix to form a tortuous path for water vapor The elongation at break of CMC/ST was not largely affected by the
transmission and to increase the effective diffusion path length addition of CNC; a decrease from 23.88% for unloaded CMC/ST to
(Bai, Sun, Xu, Dong, & Liu, 2015; Kanmani & Rhim, 2014; Yadollahi 19.01% for CMC/ST-CNC-5.0 occurred (Table 3), indicating that the
et al., 2014). This trend has been recently reported by Oun and
Rhim (2015) in cellulose nanofibril (CNF) reinforced CMC and by
Table 3
Savadekar et al. (2014) in CNC reinforced ST bio-nanocomposite Elastic modulus, tensile strength and elongation at break of neat CMC, CMC/ST blend
films. and CMC/ST-CNC bio-nanocomposite films.

Film code Elastic Tensile Elongation at


3.8. Tensile properties of films modulus (MPa) strength (MPa) break (%)

CMC 1066.16 ± 53.30 76.23 ± 3.81 22.99 ± 0.91


The relatively large aspect ratio and high elastic modulus of CMC/ST 847.23 ± 42.35 66.82 ± 3.34 23.88 ± 0.95
CNC should have a significant reinforcing impact on the mechanical CMC/ST-CNC-0.5 1225.41 ± 61.25 83.56 ± 4.17 21.27 ± 0.85
properties of the polymer. Homogenous dispersion/distribution of CMC/ST-CNC-2.5 1375.34 ± 68.80 99.06 ± 4.95 20.88 ± 0.83
CMC/ST-CNC-5.0 1650.19 ± 82.51 110.83 ± 5.54 19.01 ± 0.76
CNC along with the favorable interfacial interactions between CNC
166 N. El Miri et al. / Carbohydrate Polymers 129 (2015) 156–167

CMC/ST-CNC bio-nanocomposite films are mechanically flexible References


materials. For CMC/ST based bio-nanocomposites, the reinforcing
ability of CNC is comparable to that observed for nanoclay Almasi, H., Ghanbarzadeh, B., & Entezami, A. A. (2010). Physicochemical properties
of starch–CMC–nanoclay biodegradable films. International Journal of Biological
nanofillers at the same loading level (Almasi et al., 2010). Macromolecules, 46, 1–5.
The main reason for the improvement of the mechanical prop- Alves, J. S., dos Reis, K. C., Menezes, E. G. T., Pereira, F. V., & Pereira, J. (2015). Effect of
erties is the strong interfacial interactions between the CMC/ST cellulose nanocrystals and gelatin in corn starch plasticized films. Carbohydrate
Polymers, 115, 215–222.
blend matrix and CNC, which was due to the high contact area Azizi-Samir, M. A. S., Alloin, F., Sanchez, J. Y., & Dufresne, A. (2004). Cellulose
exposed on the CNC and their functionalized surfaces. During the nanocrystals reinforced poly (oxyethylene). Polymer, 45, 4149–4157.
processing and drying of the bio-nanocomposite films, the origi- Bai, H., Sun, Y., Xu, J., Dong, W., & Liu, X. (2015). Rheological and structural character-
ization of HA/PVA-SbQ composites film-forming solutions and resulting films as
nal hydrogen bonds formed between the macromolecular chains of affected by UV irradiation time. Carbohydrate Polymers, 115, 422–431.
CMC were replaced by new hydrogen bonds that formed between BenBettaïeb, N., Karbowiak, T., Bornaz, S., & Debeaufort, F. (2015). Spectroscopic
the hydroxyl and carboxyl groups in CMC chains, the hydroxyl analyses of the influence of electron beam irradiation doses on mechanical,
transport properties and microstructure of chitosan-fish gelatin blend films.
groups in ST chains and the hydroxyl groups in CNC. The existence
Food Hydrocolloids, 46, 37–51.
of these new hydrogen bonds improved the tensile properties of Bercea, M., & Navard, P. (2000). Shear dynamics of aqueous suspensions of cellulose
the resulted CMC/ST-CNC bio-nanocomposite films. whiskers. Macromolecules, 33, 6011–6016.
Chen, Y., Liu, C., Chang, P. R., Cao, X., & Anderson, D. P. (2009). Bionanocomposites
based on pea starch and cellulose nanowhiskers hydrolyzed from pea hull fibre:
Effect of hydrolysis time. Carbohydrate Polymers, 76, 607–615.
4. Conclusions Clark, A. H., & Ross-Murphy, S. B. (1987). Structural and mechanical properties of
biopolymer gels. Advances in Polymer Science, 83, 57–192.
Cuq, B., Aymard, C., Cuq, J. L., & Guilbert, S. (1995). Edible packaging films based on
By using the solvent casting method, new eco-friendly bio- fish myofibrillar proteins: Formulation and functional Properties. Journal of Food
nanocomposite film materials that were based on carboxymethyl Science, 160, 1369–1374.
cellulose (CMC)/starch (ST) biopolymer blend filled with cellulose De Menezes, A. J., Siqueira, G., Curvelo, A. A. S., & Dufresne, A. (2009). Extrusion and
characterization of functionalized cellulose whiskers reinforced polyethylene
nanocrystals (CNC) have been developed and studied through this
nanocomposites. Polymer, 50, 4552–4563.
paper. In particular, the rheology of film-forming solution (FFS) and Dos Santos, R. M., Neto, W. P. F., Silvério, H. A., Martins, D. F., Dantas, N. O., & Pasquini,
the structural, transparency, water vapor permeability and ten- D. (2013). Cellulose nanocrystals from pineapple leaf, a new approach for the
reuse of this agro-waste. Industrial Crops and Products, 50, 707–714.
sile properties of resulting solid cast films were examined. It was
Eichhorn, S. J., Dufresne, A., Aranguren, M., Marcovich, N. E., Capadona, J. R., Rowan, S.
found that the steady shear viscosity showed that the dispersion J., et al. (2010). Review: Current international research into cellulose nanofibres
of CNC turned CMC and CMC/ST FFS from a Newtonian behavior and nanocomposites. Journal of Materials Science, 45, 1–33.
into a shear thinning behavior, especially at low shear rate val- El Achaby, M., Arrakhiz, F. E., Vaudreuil, S., Essassi, E., & Qaiss, A. (2012). Piezoelectric
␤ polymorph formation and properties enhancement in graphene oxide–PVDF
ues, and the zero shear viscosity increased with the increase of the nanocomposite films. Journal of Applied Polymer Science, 258, 7668–7677.
CNC content. Dynamic tests showed that all the studied FFS had a El Achaby, M., Essamlali, Y., El Miri, N., Snik, A., Abdelouahdi, K., Fihri, A., et al.
viscoelastic behavior with an entanglement network structure, as (2014). Graphene oxide reinforced chitosan/polyvinylpyrrolidone polymer bio-
nanocomposites. Journal of Applied Polymer Science, 131, 41042.
characterized by the intersecting of G and G at the middle of the El Miri, N., Abdelouahdi, K., Zahouily, M., Fihri, A., Barakat, A., Solhy, A., et al. (2015).
frequency range. In relation to the cast film quality, it was found Bio-nanocomposite films based on cellulose nanocrystals filled polyvinyl alco-
that the measured rheological properties were suitable for casting hol/chitosan polymer blend. Journal of Applied Polymer Science, 132, 42004.
Espinosa, S. C., Kuhnt, T., Foster, E. J., & Weder, C. (2013). Isolation of thermally
all the FFS to form solid films with high quality and smooth surface. stable cellulose nanocrystals by phosphoric acid hydrolysis. Biomacromolecules,
Due to the homogeneous dispersion of CNC into CMC/ST blend, the 14, 1223–1230.
as-produced CMC/ST-CNC bio-nanocomposite films showed a good Flauzino-Neto, W. P. F., Silvério, H. A., Dantas, N. O., & Pasquini, D. (2013). Extraction
and characterization of cellulose nanocrystals from agro-industrial residue–soy
optical transparency (80–95%). At the same time, the presence of
hulls. Industrial Crops and Products, 42, 480–488.
CNC in bio-nanocomposite films reduced significantly the water Florjancic, U., Zupancic, A., & Zumer, M. (2002). Rheological characterization of
vapor permeability of the neat CMC and the CMC/ST blend. Ten- aqueous polysaccharide mixtures undergoing shear. Chemical and Biochemical
Engineering Quarterly, 16, 105–118.
sile tests showed that the elastic modulus and tensile strength of
Fortunati, E., Luzi, F., Puglia, D., Petrucci, R., Kenny, J. M., & Torre, L. (2015). Processing
CMC/ST blend based bio-nanocomposite films increased by 94.77% of PLA nanocomposites with cellulose nanocrystals extracted from Posidoniao-
and 65.86%, respectively, when 5.0 wt% CNC were added. The ceanica waste: Innovative reuse of coastal plant. Industrial Crops and Products,
enhancement of the properties of bio-nanocomposite films was 67, 439–447.
Fortunati, E., Puglia, D., Luzi, F., Santulli, C., Kenny, J. M., & Torre, L. (2013). Binary PVA
due to strong interfacial interactions generated from the hydro- bio-nanocomposites containing cellulose nanocrystals extracted from different
gen bonding between the functional groups that are present in natural sources: Part I. Carbohydrate Polymers, 97, 825–836.
the mixed components (CMC, ST and CNC). The as-produced bio- Garcia de Rodriguez, N. L., Thielemans, W., & Dufresne, A. (2006). Sisal cellulose
whiskers reinforced polyvinyl acetate nanocomposites. Cellulose, 13, 261–270.
nanocomposite films exhibited good optical transparency, reduced Ghanbarzadeh, B., Almasi, H., & Entezami, A. A. (2010). Physical properties of edi-
WVP and enhanced tensile properties, which are the main proper- ble modified starch/carboxymethyl cellulose films. Innovative Food Science and
ties required for packaging applications. Emerging Technologies, 11, 697–702.
Ghasemlou, M., Khodaiyan, F., & Oromiehie, A. (2011). Rheological and structural
characterisation of film-forming solutions and biodegradable edible film made
from kefiran as affected by various plasticizer types. International Journal of
Acknowledgments Biological Macromolecules, 49, 814–821.
Habibi, Y., Lucia, L. A., & Rojas, O. J. (2010). Cellulose nanocrystals: Chemistry, self-
assembly and applications. Chemical Reviews, 110, 3479–3500.
The financial assistance of the Office Chérifien des Phosphates Kamal, M. R., & Khoshkava, V. (2015). Effect of cellulose nanocrystals (CNC) on
(OCP Group) in the Moroccan Kingdom toward this research is rheological and mechanical properties and crystallization behavior of PLA/CNC
hereby acknowledged. This work was supported also by grant nanocomposites. Carbohydrate Polymers, 123, 105–114.
Kanmani, P., & Rhim, J. W. (2014). Properties and characterization of bio-
from the OCP Foundation. The authors would like to acknowl- nanocomposite films prepared with various biopolymers and ZnO nanoparticles.
edge SUNABEL-COSUMAR Company in Morocco for providing us Carbohydrate Polymers, 106, 190–199.
the raw sugarcane bagasse. This work was performed as part of a Kulicke, W. M., Kull, A. H., Kull, W., & Thielking, H. (1996). Characterization of aque-
ous carboxymethylcellulose solutions in terms of their molecular structure and
collaboration between the Mohammed VI Polytechnic University its influence on rheological behaviour. Polymer, 37, 2723–2731.
and INRA-Montpellier. Financial support of the INRA-Montpellier Layek, R. K., Kundu, A., & Nandi, A. K. (2013). High-performance nanocomposites of
is acknowledged. We equally thank all administrative and tech- sodium carboxymethyl cellulose and graphene oxide. Macromolecular Materials
and Engineering, 298, 1166–1175.
nical support teams of the Université Mohammed VI Polytechnique,
Liu, D., Chen, X., Yue, Y., Chen, M., & Wu, Q. (2011). Structure and rheology of
especially Mrs Khadija Ait Hadouch et Mrs Loubnae Choukard. nanocrystalline cellulose. Carbohydrate Polymers, 84, 316–322.
N. El Miri et al. / Carbohydrate Polymers 129 (2015) 156–167 167

Lu, P., & Hsieh, Y. L. (2012). Preparation and characterization of cellulose nanocrystals Savadekar, N. R., Karande, V. S., Vigneshwaran, N., Kadam, P. G., & Mhaske, S. T.
from rice straw. Carbohydrate Polymers, 87, 564–573. (2014). Preparation of cotton linter nanowhiskers by high-pressure homoge-
Luna-Martínez, J. F., Hernández-Uresti, D. B., Reyes-Melo, M. E., Guerrero-Salazar, nization process and its application in thermoplastic starch. Applied Nanoscience,
C. A., González-González, V. A., & Sepúlveda-Guzmán, S. (2011). Synthesis and 5, 281–290.
optical characterization of ZnS–sodium carboxymethyl cellulose nanocompos- Silva-Pereira, M. C., Teixeira, J. A., Pereira-Júnior, V. A., & Stefani, R. (2015). Chi-
ite films. Carbohydrate Polymers, 84, 566–570. tosan/corn starch blend films with extract from Brassica oleraceae (red cabbage)
Mariano, M., El Kissi, N., & Dufresne, A. (2014). Cellulose nanocrystals and related as a visual indicator of fish deterioration. Food Science and Technology, 61,
nanocomposites: Review of some properties and challenges. Journal of Polymer 258–262.
Science Part B: Polymer Physics, 52, 791–806. Silva-Weiss, A., Bifani, V., Ihl, M., Sobral, P. J. A., & Gómez-Guillén, M. C. (2014).
Moraes, I. C. F., Carvalho, R. A., Bittante, A. M. Q. B., Solorza-Feria, J., & Sobral, P. Polyphenol-rich extract from murta leaves on rheological properties of film-
J. A. (2009). Film forming solutions based on gelatin and poly (vinyl alcohol) forming solutions based on different hydrocolloid blends. Journal of Food
blends: Thermal and rheological characterizations. Journal of Food Engineering, Engineering, 140, 28–38.
95, 588–596. Silvério, H. A., Flauzino-Neto, W. P., Dantas, N. O., & Pasquini, D. (2013). Extrac-
Morais, J. P. S., Rosa, M. D. F., Filho, M. D. M. D. S., Nascimento, L. D., Nascimento, D. tion and characterization of cellulose nanocrystals from corncob for application
M. D., & Cassales, A. R. (2013). Extraction and characterization of nanocellulose as reinforcing agent in nanocomposites. Industrial Crops and Products, 44,
structures from raw cotton linter. Carbohydrate Polymers, 91, 229–235. 427–436.
Olsson, R. T., Kramer, R. H., Lopez-Rubio, A., Torres-Giner, S., Ocio, M. J., & Lagaron, J. Sionkowska, A. (2011). Current research on the blends of natural and synthetic poly-
M. (2010). Extraction of microfibrils from bacterial cellulose networks for elec- mers as new biomaterials: Review. Progress in Polymer Science, 36, 1254–1276.
trospinning of anisotropic biohybrid fiber yarns. Macromolecules, 43, 4201–4209. Siqueira, G., Bras, J., Follain, N., Belbekhouche, S., Marais, S., & Dufresne, A. (2013).
Oun, A. A., & Rhim, J.-W. (2015). Preparation and characterization of sodium Thermal and mechanical properties of bio-nanocomposites reinforced by Luffa
carboxymethyl cellulose/cotton linter cellulose nanofibril composite films. Car- cylindrica cellulose nanocrystals. Carbohydrate Polymers, 91, 711–717.
bohydrate Polymers, 127, 101–109. Teixeira, E. d. M., Bondancia, T. J., Teodoro, K. B. R., Corrêa, A. C., Marconcini, J. M., &
Peressini, D., Bravin, B., Lapasin, R., Rizzotti, C., & Sensidoni, A. (2003). Mattoso, L. H. C. (2011). Sugarcane bagasse whiskers: Extraction and character-
Starch–methylcellulose based edible films: Rheological properties of film- izations. Industrial Crops and Products, 33, 63–66.
forming dispersions. Journal of Food Engineering, 59, 25–32. Tongdeesoontorn, W., Mauer, L. J., Wongruong, S., Sriburi, P., & Rachtanapun, P.
Reddy, M. M., Vivekanandhan, S., Misra, M., Bhati, S. K., & Mohanty, A. K. (2013). (2011). Effect of carboxymethyl cellulose concentration on physical properties
Biobased plastics and bionanocomposites: Current status and future opportu- of biodegradable cassava starch-based films. Chemistry Central Journal, 5, 6–13.
nities. Progress in Polymer Science, 38, 1653–1689. Van den Broek, L. A. M., Knoop, R. J. I., Kappen, F. H. J., & Boeriu, C. G. (2015).
Rhim, J. W., Park, H. M., & Ha, C. S. (2013). Bio-nanocomposites for food packaging Chitosan films and blends for packaging material. Carbohydrate Polymers, 116,
applications. Progress in Polymer Science, 38, 1629–1652. 237–242.
Rodríguez-González, C., Martínez-Hernández, A. L., Castaño, V. M., Kharissova, O. V., Wu, D., Wang, J., Zhang, M., & Zhou, W. (2012). Rheology of carbon nanotubes-
Ruoff, R. S., & Velasco-Santos, C. (2012). Polysaccharide nanocomposites rein- filled poly(vinylidene fluoride) composites. Industrial & Engineering Chemistry
forced with graphene oxide and keratin-grafted graphene oxide. Industrial & Research, 51, 6705–6713.
Engineering Chemistry Research, 51, 3619–3629. Xu, X., Liu, W., & Zhang, L. (2006). Rheological behavior of Aeromonas gum in aqueous
Rosa, M. F., Medeiros, E. S., Malmonge, J. A., Gregorski, K. S., Wood, D. F., Mat- solutions. Food Hydrocolloids, 20, 723–729.
toso, L. H. C., et al. (2010). Cellulose nanowhiskers from coconut husk fibers: Yadollahi, M., Namazi, H., & Barkhordari, S. (2014). Preparation and properties
Effect of preparation conditions on their thermal and morphological behavior. of carboxymethyl cellulose/layered double hydroxide bionanocomposite films.
Carbohydrate Polymers, 81, 83–92. Carbohydrate Polymers, 108, 83–90.
Sacui, I. A., Nieuwendaal, R. C., Burnett, D. J., Stranick, S. J., Jorfi, M., Weder, C., et al. Zhou, C., Wang, Q., & Wu, Q. (2012). UV-initiated crosslinking of electro-
(2014). Comparison of the properties of cellulose nanocrystals and cellulose spunpoly(ethylene oxide) nanofibers with pentaerythritoltriacrylate: Effect of
nanofibrils isolated from bacteria, tunicate, and wood processed using acid, irradiation time and incorporated cellulose nanocrystals. Carbohydrate Polymers,
enzymatic, mechanical, and oxidative methods. ACS Applied Material Interfaces, 87, 1779–1786.
6, 6127–6138.

You might also like