Вы находитесь на странице: 1из 12

Industrial Crops & Products 112 (2018) 499–510

Contents lists available at ScienceDirect

Industrial Crops & Products


journal homepage: www.elsevier.com/locate/indcrop

Alfa fibers as viable sustainable source for cellulose nanocrystals extraction: T


Application for improving the tensile properties of biopolymer
nanocomposite films

Mounir El Achabya, , Zineb Kassaba,b, Abdelatif Barakatc, Adil Aboulkasd
a
Materials Science and Nanoengineering Department, Mohammed VI Polytechnic University, Lot 660–Hay Moulay Rachid, 43150, Benguerir, Morocco
b
Laboratoire d'Ingénierie et Matériaux (LIMAT), Faculté des Sciences Ben M'sik, Université Hassan II de Casablanca, B.P.7955, Casablanca, Morocco
c
IATE, CIRAD, Montpellier SupAgro, INRA, Université de Montpelier, 34060, Montpellier, France
d
Laboratoire des procédés chimiques et matériaux appliqués (LPCMA), Faculté polydisciplinaire de Béni-Mellal, Université Sultan Moulay Slimane, BP 592, 23000 Béni-
Mellal, Morocco

A R T I C L E I N F O A B S T R A C T

Keywords: Due to its renewability, availability and high cellulose content (≈45%), Alfa fibers (Stipa tenacissima) have been
Lignocellulosic fibers identified as a sustainable source for cellulose microfibers (CMF) and cellulose nanocrystals (CNC) production.
Cellulose nanocrystals Subjecting raw Alfa fibers to alkali, bleaching and sulfuric acid hydrolysis treatments allowed producing CMF
Biopolymer nanocomposites and CNC with high yields. The fluorescence microscopy confirmed that CMF, with average diameter of 10 μm,
Tensile properties
were successfully obtained after bleaching treatments. TEM and AFM showed that the CNC exhibit needle-like
shape with an average diameter and length of 5 ± 3 nm and 330 ± 30 nm, respectively, giving rise to an aspect
ratio of about 66. XPS measurement confirmed the presence of sulfate groups on the surface of CNC with 2.04
sulfate groups per 100 anhydroglucose units, confirming the negatively charged surface of CNC, with zeta po-
tential value of − 47.39 mV. XRD studies showed that CMF and CNC exhibit cellulose I structure with crys-
tallinity index of 71% and 90%, respectively. FTIR and TGA analyses were used to identify the chemical com-
position and thermal stability changes during different chemical treatments, suggesting that all non-cellulosic
compounds were removed after alkali and bleaching treatments. The obtained CNC were dispersed into three
different biopolymer matrices, e.g. chitosan, alginate, and k-carrageenan, at various CNC loadings (1, 3, 5 and
8 wt%), to evaluate their ability to enhance the tensile properties of biopolymers and, at the same time, to
produce new biopolymer-based nanocomposite films. It was found that the tensile properties of the as-produced
nanocomposite films were largely improved with addition of CNC, resulting in mechanically strong and flexible
ecofriendly nanocomposite films.

1. Introduction been devoted to use lignocellulosic fibers as renewable raw materials


for the production of cellulose, lignin and hemicellulose, as high value-
Natural lignocellulosic fibers are the most abundant renewable raw added materials or polymers (Bian et al., 2012; Nadji et al., 2009;
materials on earth, and mainly consists of cellulose, hemicelluloses and Trache et al., 2016). However, it is well known that lignocellulosic fi-
lignin as major constituents, and pectin, waxes, proteins, lipids, ash, bers exhibit a complex hierarchal structure in which the constituents
pigments and extractive compounds can be found as minor constituents are linked with each other, making it difficult to fractionate these ori-
(Malherbe and Cloete, 2002). In general, the composition and propri- ginal fibers into value added materials (Rangan et al., 2017).
eties of lignocellulosic fibers vary depending on the origin and the type From lignocellulosic materials, the pure cellulose fibers can be
of fibers, species of plant, and the environment in which the original prepared using physical, mechanical and/or chemical routes (Trache
plant is growing (Trache et al., 2016). Most of the lignocellulosic fibers et al., 2016). In this case, the emphasis is to remove lignin and hemi-
such as cotton, flax, hemp, jute, kenaf, sisal, ramie, curaua, pineapple, cellulose and focus on getting a maximum yield of pure cellulose, which
bamboo, coir can be isolated, treated and functionalized before used in is a hierarchically structured material that contains both highly crys-
various applications such as textiles (Yu, 2015) and polymer composites talline domains (ordered cellulose chains stabilized by inter and in-
(Pickering et al., 2016). More importantly, considerable efforts have tramolecular hydrogen bonding), and amorphous domains (disordered


Corresponding author.
E-mail address: mounir.elachaby@um6p.ma (M. El Achaby).

https://doi.org/10.1016/j.indcrop.2017.12.049
Received 29 July 2017; Received in revised form 25 November 2017; Accepted 19 December 2017
Available online 04 January 2018
0926-6690/ © 2017 Elsevier B.V. All rights reserved.
M. El Achaby et al. Industrial Crops & Products 112 (2018) 499–510

cellulose chains). Pure cellulose fibers can be chemically or mechani- alginate (ALg) and k-Carrageenan (k-CA) have been used extensively for
cally disintegrated yielding in nanosized cellulose called nanocelluloses the preparation of biopolymer-based nanocomposite films because of
(NC), cellulose nanocrystals (CNC) or cellulose nanowhiskers (CNC), their excellent film forming properties (Huq et al., 2012; Khan et al.,
having very interesting properties such as high aspect ratio (10–100), 2012; Sánchez-García et al., 2010).
large specific surface area (∼250–500 m2/g), low density (∼1.6 g/ CH is a linear biopolymer and deacetylated derivative of chitin. It is
cm3), high crystallinity (up to 90%), and high tensile strength composed of β-(1,4)-2-acetamido-2-deoxy-D-glucose and β-(1,4)-2-
(7.5 GPa), and very high elastic modulus (approximately 100–140 GPa) amino-2-deoxy-D-glucose units (El Achaby et al., 2014). This biopo-
(Bandera et al., 2014; Buffiere et al., 2017; Grishkewich et al., 2017; lymer is one of the most interesting polymers due to its advanced ap-
Tan et al., 2015). Recently, mechanical treatments such as high shear plications that range from pharmaceuticals to materials science (El Miri
homogenization, ball-milling or grinding processes have shown to be et al., 2015b). ALg is a water soluble biopolymer derived from brown
efficient in producing nanometer or micrometer sized cellulose parti- seaweed such as Laminaria hyperborea, Macrocystis pirifera, Laminaria
cles, with relatively high aspect ratio (Bandera et al., 2014; Buffiere digitata, and Ascophyllum nodosum (Remminghorst and Rehm, 2006).
et al., 2017; Phanthong et al., 2017; Sofla et al., 2016). For example, it ALg is composed of β-D-mannuronic acid and a-L-glucuronic acid, which
has been reported that cellulose whiskers with average aspect ratio of are linked together in varying proportions by 1–4-linkages
52 can be extracted from microcrystalline cellulose by using high shear (Remminghorst and Rehm, 2006). Owing to its good film forming
homogenization process (Bandera et al., 2014). They demonstrated that ability, ALg biopolymer has been used in various industrial applications
these mechanically obtained cellulose whiskers performed a similar such as food additives and gelling agents, emulsifier, food packaging
reinforcing potential to commercially available whiskers. Regarding to (Carneiro-da-Cunha et al., 2010; Shankar et al., 2016). k-CA is also a
the chemical treatment, the sulfuric acid hydrolysis represents the most water soluble biopolymer extracted from various species of red algae
effective process for obtaining CNC, because it produces CNC with (Rhodophyta). It is composed of long linear chains of D-galactose and D-
sulfate-modified surface, high crystallinity and good colloidal stability anhydrogalactose with anionic sulfate groups (Liu et al., 2015). This
in water (Habibi et al., 2010). However, it should be noted that the biopolymer has good gel and film forming properties and has been used
morphology, the surface charge density (sulfur content), and the in various applications particularly as a food additive or as a stabilizer
properties of acid hydrolyzed-CNC depend on the source of the original (Dafe et al., 2017).
cellulose and on the hydrolysis conditions, such as time, temperature, The objective of this work was to exploit raw Alfa fibers, harvested
agitation and acid concentration (Martins et al., 2015; Sapkota et al., from the oriental region of eastern Morocco, for the production of pure
2017a; Trache et al., 2017; Yao et al., 2017). cellulose microfibers and cellulose nanocrystals (CNC) using alkaliza-
“Alfa” fibers are considered as the most important lignocellulosic tion, bleaching and sulfuric acid hydrolysis treatments. Indeed, the
fibers feedstocks. “Alfa” is the Arabic name of Stipa tenacissima plant, valorization of the Alfa fibers for the production of cellulose nano-
which is a fast growing perennial plant that thrives in dry regions of crystals using such processes should contribute to the economic viabi-
North Africa (mainly in the Maghreb) and southeast Spain (Trache lity and the sustainability of this largely available lignocellulosic bio-
et al., 2014). Its fibers rich leaves can reach 1000 mm in length, with mass. After their successful characterization, the as-extracted CNC were
spherical shape. In North African countries, it covers a large area esti- dispersed into three different biopolymer matrices, e.g. chitosan (CH),
mated to about 3,186,000 ha in Morocco, 4,000,000 ha in Algeria, alginate (AL), and k-carrageenan (k-CA), to evaluate their ability to
400,000 ha in Tunisia and 350,000 ha in Libya (Marrakchi et al., 2012; reinforce of biopolymers with different surface functionalities and
Paiva et al., 2007). Alfa is considered as one of the most interesting properties and, at the same time, to produce new biopolymer-based
annual plants for the production of fibers for papermaking (Marrakchi nanocomposite films. Herein, the adhesion properties and the functio-
et al., 2011) and polymer composites development (Brahim and Cheikh, nalized surface of sulfuric acid hydrolyzed CNC can be widely exploited
2007; Maafi et al., 2010; Omri et al., 2016). to enhance the interfacial interactions between the dispersed CNC and
Alfa fibers are composed mainly of cellulose (44–48 wt%), hemi- the biopolymer matrices (CH, ALG and k-CA).
celluloses (27–22 wt%) and lignin (12–18 wt%) (Belkhir et al., 2013;
Hanana et al., 2015), making them a promising substrate for the pro- 2. Materials and experimental details
duction of carbohydrate derivatives with good performances. Recently,
Alfa fibers have been used for the production of microcrystalline and 2.1. Materials
nanofibrillated cellulose (Besbes et al., 2011; Trache et al., 2014) and
lignin-based biopolymers (Nadji et al., 2009). Importantly, the high The raw Alfa fibers (AF) used in this work was collected from the
content of cellulose in Alfa fibers make them a very interesting bio- oriental region of eastern Morocco. The as-received AF (up to 1000 mm
sourced raw material for the production of CNC with high reinforcing fiber long) (Fig. 1) were cut into 50–100 mm small fibers, and then they
potential for biopolymers based nanocomposite development. were ground using a knife mill (Retch SM100) with a screen size of
Due to their main advantages including nontoxicity, biodegrad- 2 mm (Fig. 1). Chitosan (740063 Aldrich), alginate (W201502 Aldrich)
ability, wide availability and biocompatibility, the biopolymers can be and k-carrageenan (22048 Sigma) were purchased from Sigma Aldrich.
considered as good candidates for the replacement of petroleum-based All analytical grade chemicals used for the treatment of raw fibers and
polymers; however, their mechanical, thermal, and barrier properties for the extraction of cellulosic materials were also purchased from
must be enhanced to make them attractive materials for various ap- Sigma–Aldrich.
plications (El Miri et al., 2015a,b). Nevertheless, the nanoreinforcement
strategy, by adding suitable reinforcing nanofillers, is an effective way 2.2. Extraction of cellulose microfibers (CMF)
to enhance the physico-chemical properties of biopolymers, by produ-
cing high performance biopolymer-based nanocomposite materials (El Pure cellulose microfibers (CMF) were firstly extracted from raw AF
Miri et al., 2015a). Indeed, the potential for achievement of useful according to our previous works (El Achaby et al., 2017; El Miri et al.,
biopolymer based nanocomposites must be driven by the excellent in- 2016, 2015a). Briefly, ground raw AF (2 mm) were treated in distilled
trinsic properties of nanofillers and their nanoscale dispersion within water for 1 h at 60 °C. Then, the resulted AF were treated three times
biopolymer matrix (El Miri et al., 2015a). In this context, various bio- with 4 wt% NaOH solution at 80 °C for 2 h under stirring, resulting in
polymers have been used as polymeric matrices to develop bionano- alkali treated Alfa fibers (ATAF) (Fig. 1), after which bleaching treat-
composite materials with high mechanical, optical, thermal, and barrier ment was carried out using a solution made up of equal parts (v:v) of
properties, using CNC as nanoreinforcing fillers (El Achaby et al., 2017; acetate buffer (27 g NaOH and 75 mL glacial acetic acid, diluted to 1 L
El Miri et al., 2015a,b). Among such biopolymers, chitosan (CH), of distilled water) and aqueous sodium chlorite (1.7 wt% NaClO2 in

500
M. El Achaby et al. Industrial Crops & Products 112 (2018) 499–510

Fig 1. Photographs of (a) raw and (b) crushed alfa fibers (AF), (c)
alkali treated alfa fibers (ATAF), (d) bleached cellulose microfibers
(CMF) and (e) suspended and (f) freeze dried cellulose nanocrystals
(CNC).

water). This treatment was done three times, resulting in pure white material for different physico-chemical characterizations (Fig. 1).
colored cellulose microfibers (Fig. 1). In these treatments the ratio of
the fibers to liquor was 1/20 (g/mL). 2.4. Preparation of nanocomposite films

2.3. Extraction of cellulose nanocrystals (CNC) Nanocomposite films based on chitosan (CH), alginate (ALg) and k-
carrageenan (k-CA) polysaccharides as biopolymeric matrices filled
CNC were successfully extracted from the as-extracted cellulose with 1; 3; 5 and 8 wt% CNC, as nanodispersed phase, were prepared
microfibers (CMF) by acid hydrolysis process, as described in our pre- using solvent casting method (Fig. 2). For CH-CNC nanocomposite
vious works (El Achaby et al., 2017; El Miri et al., 2016, 2015b). For films, the desired amount of CNC (1; 3; 5 and 8 wt%) was dispersed into
that, CMF were subjected to preheated 64 wt% sulfuric acid hydrolysis 250 mL of 1% acetic acid aqueous solutions and stirred for 5 min at
at 50 °C for 30 min under mechanical stirring. Then, the mixture was 65 °C. Then, CH powder was slowly added to CNC suspensions and the
diluted with ice cubes to stop the reaction and was washed by succes- resulting mixture (3 g total dry mass) was stirred for 2 h at 65 °C, then
sive centrifugations at 12000 rpm at 15 °C for 20 min at each step, re- cooled to room temperature under stirring. For ALg-CNC or k-CA-CNC
sulting in white CNC suspension, which was subjected to a dialysis nanocomposite films, the desired amount of CNC (1; 3; 5 and 8 wt%)
against distillated water until it reached neutral pH. Afterward, the was dispersed into 250 mL in aqueous solutions containing 1 g of gly-
obtained CNC aqueous suspension was homogenized by the use of a cerol and stirred for 5 min at 65 °C. Then, AL (or k-CA) powder was
probe type ultrasonic homogenizer (BRANSON, Sonifier 250) for 5 min slowly added to CNC suspensions and the resulting mixture was stirred
in an ice bath, resulting white stable CNC suspension, in the form of a for 2 h min at 65 °C, then cooled to room temperature under stirring.
gel (Fig. 1). The concentration of CNC aqueous suspension was calcu- After cooling to room temperature, all mixtures were casted onto
lated gravimetrically. Finally, by using freeze drying process a small Petri dishes and the water evaporated at ambient temperature for
quantity of homogenized CNC suspension was transformed into solid 3 days. Neat CH, ALg and k-CA films were also prepared according to

501
M. El Achaby et al. Industrial Crops & Products 112 (2018) 499–510

Fig. 2. Schematic representation of the processing of nanocomposite films based on CNC as nanofillers and ALg, k-CA and CH as biopolymer matrices.

the same procedures mentioned above, without the addition of CNC. accumulation of 16 scans. The FTIR spectra were taken in the trans-
The produced nanocomposite films were coded as Y-CNC-X, with Y mittance mode. The XPS analysis of freeze dried CNC was conducted on
referred to biopolymeric matrix CH, ALg or k-CA and the X referred to an Escalab 250 apparatus (Thermo Electron) equipped with a mono-
CNC weight fractions (1, 3, 5 and 8 wt%) in each nanocomposite for- chromatic Al Kα radiation at 1486.6 eV. The area of analysis was
mulations. 400 μm diameter. The spectra were modified by setting the CeC con-
tribution in the C1 s emission to 285.0 eV. The thickness of nano-
composite films was determined using a Hanatek Variable Force
2.5. Measurement and characterization techniques Precision Thickness Gauge (FT3-V). Tensile tests for all bionano-
composite films were performed at room temperature using texture
The morphology of samples was evaluated by using an inverted Analyser (TA.XT plus). The specimens were cut in rectangular shapes
Nikon A1 laser scanning confocal microscope (LSCM). The samples with 80 mm in length and 10 mm in width. The gauge length was fixed
were stained with Congo Red (Merck CI22120) by incubating them in a at 30 mm and the crosshead speed was 5 mm/min. All tests were carried
0.01% solution in distilled water for 20 min. X-ray diffraction (XRD) out on a minimum of five samples and the reported results are average
characterizations of samples was performed on a Bruker diffractometer values.
D8 Advance using Cu-Kα radiation. The diffraction profiles were ob-
tained at diffraction angles between 5 and 50° at room temperature.
Atomic Force Microscopy (AFM) images of CNC was recorded using a 3. Results and discussions
BioScope Catalyst AFM (Bruker AXS, Santa Barbara, CA). Samples for
AFM observations were prepared by depositing a droplet of a diluted 3.1. Extraction of CMF and CNC
CNC suspension onto freshly cleaved mica sheets, after being sonicated
for 5 min, and allowing the solvent to dry in air. The measurements Micro and nanosized cellulose (CMF and CNC) were successfully
were conducted in tapping mode under ambient temperature at a scan extracted from raw AF, using the three well known steps that involves a
rate of 1.5 Hz. Transmission electron microscopy (TEM) for CNC char- combination of alkali, bleaching and acid hydrolysis treatments, ac-
acterization was performed using JEOL JEM-1230 (Jeol, Japan) oper- cording to our previous works (El Achaby et al., 2017; El Miri et al.,
ated at 80 kV and equipped with a LaB6 filament. Droplets of CNC 2016, 2015a). In order to remove the noncellulosic components, espe-
suspensions were deposited on glow-discharged carbon-coated grids. cially the hemicellulose and lignin molecules, the ground raw AF
The liquid in excess was blotted away with filter paper and a drop of 2% (2 mm) were firstly subjected to washing treatment using distilled
(w/v) uranyl acetate negative stain was added prior to air drying. The water followed by three subsequent alkali treatments to ensure at least
zeta potential measurement of CNC suspension was measured using a partial removal of the hemicellulose and lignin molecules. These
NICOMP ZLS Z3000 Zeta sizer apparatus. For this measurement, ca- treatments yielded about 51% yellow colored fibers (Fig. 1), which was
pillary cell was used and the CNC suspension was diluted and sonicated calculated by the weight of the obtained fibers divided by the initial
for 5 min before being analyzed. The thermal degradation behaviors of weight of AF. The alkali treated Alfa fibers (ATAF) were then subjected
all cellulosic materials were characterized using a thermogravimetric to a bleaching treatment to degrade the residual lignin molecules and
analyzer (METTLER TOLEDO, TGA2). All tests were conducted under a others impurities, resulting in solid pure cellulose microfibers (CMF),
nitrogen atmosphere, at a flow rate of 25 mL/min, from 25 to 800 °C shown by a clearly white color (Fig. 1), confirming that the all non-
with heating rate of 10 °C/min. cellulosic compounds were totally removed after bleaching treatment.
Fourier transform infrared spectroscopy (FTIR) was performed on a Herein, the yield of these CMF was about 39% in regard to the initial
Bruker Tensor 27 spectrometer equipped with a Golden Gate single amount of AF. Finally, the resulting bleached CMF were hydrolyzed
reflection ATR accessory. The experiments were carried out in the range with sulfuric acid, which was performed under appropriate conditions
from 4000 to 400 cm−1 with a resolution of 4 cm−1 and an that allowed the removal of amorphous domains from the bleached

502
M. El Achaby et al. Industrial Crops & Products 112 (2018) 499–510

Fig. 3. Laser scanning confocal fluorescence microscopy images of (a,b) AF, (c,d) ATAF and (e,f) CMF samples.

cellulose (Kargarzadeh et al., 2012). The removal of amorphous do- 3.3. TEM and AFM analysis of CNC
mains, via acid hydrolysis, was performed by cleaving CMF into bundles
of CNC with nanometric dimensions. As shown in Fig. 1, a viscous gel TEM and AFM microscopies were used to examine the morphology
for CNC suspension and a clearly white powder were obtained after and dimensions of CNC. The shape and the aspect ratio of CNC played
homogenization and freeze drying processes, respectively. The yield of an important role in determining reinforcing capability of the CNC for
CNC was 21% with respect to the initial amount of dried AF fibers, and polymer nanocomposite application (Sapkota et al., 2017a,b; Trache
57% with respect to the amount of bleached CMF. et al., 2017). Thus, these parameters are evaluated from TEM and AFM
observations (Fig. 4). From this result, it was observed that the CNC are
well separated and exhibited a needle-like shape, which is a well known
3.2. Morphological analysis of raw and treated cellulosic fibers morphology for CNC extracted from lignocellulosic materials using acid
hydrolysis process (El Miri et al., 2015b; Kargarzadeh et al., 2012;
Fig. 3 represents the fluorescence microscopy images of raw AF, Mariano et al., 2016). This shape can be considered as a special char-
ATAF and CMF. These images show that the morphology of fibers, in acteristic to be taken into account to use the needle-like CNC for de-
terms of the size and the defibrillation level, was changed from compact veloping CNC-based polymer nanocomposite materials.
fibrous structure for raw AF into totally defibrillated structure (micro- The average diameter and length of CNC were measured from AFM
fibers) for bleached cellulose. For the raw AF, it is worth noting that the and TEM observations at 5 ± 3 nm and 330 ± 30 nm, respectively,
presence of lignin, hemicelluloses and other noncellulosic components giving rise to an average aspect ratio of about 66, which is calculated as
act as cementing agent around the cellulose fibrils and help in the the length (L) divided by the diameter (D) for individual CNC, in-
formation of compact structure (Fig. 3a–b). With the subsequent dicating the potential application of CNC as nanoreinforcing agents in
treatment of alkali, some alkali labile linkages (ether and ester linkages) composite materials (Sapkota et al., 2017a,b,c; Trache et al., 2017).
between lignin monomers or between lignin and other carbohydrates Indeed, it is widely accepted that the starting raw material and the
may have been broken (El Miri et al., 2015b; Trache et al., 2014), which pretreatment and acid hydrolysis conditions are the most important
helped in the partial defibrillation of fibers, as shown in Fig. 3c–d. The factors affecting the dimensions of the produced CNC (Trache et al.,
laser scanning fluorescence microscopy micrographs obtained for CMF 2017). Herein, the aspect ratio obtained for the as-extracted CNC is
(Fig. 3e–f) show that the bleaching treatment resulted in the total de- comparable to the values reported in the literature for CNC extracted
fibrillation of fibers into individual microfibers with smaller diameter from various sources such as sugarcane bagasse (L/D = 32–64)
(average diameter of 10 μm), confirming the all noncellulosic com- (Teixeira et al., 2011), corncob (52–63) (Silvério et al., 2013), sisal
pounds were totally removed after bleaching treatment, which was (43–60) (Garcia de Rodriguez et al., 2006) and pineapple leaf (50–60)
suitable for acid hydrolysis process to produce CNC. These results are in (Marcos dos Santos et al., 2013); and higher than that measured for
good agreement with literature for obtaining pure CMF from lig- CNC from rice straw (9–10.5) (Lu and Hsieh, 2011), cotton (9.5–12.5)
nocellulosic fibers using the same bleaching treatment (Cudjoe et al., (Camarero Espinosa et al., 2013) or cotton linters (20–24) (Paulo et al.,
2017; El Miri et al., 2015b; Thambiraj and Ravi Shankaran, 2017; 2013), coconut husks (35–44) (Rosa et al., 2010), ramie (12) (Junior De
Trache et al., 2014). Menezes et al., 2009) and Luffa cylindrica fibers (46) (Siqueira et al.,
2013). Contrastingly, it was smaller than that of CNC from tunicate
(148) (Sacui et al., 2014) and bacteria (60–94) (Olsson et al., 2010;

503
M. El Achaby et al. Industrial Crops & Products 112 (2018) 499–510

Fig. 4. (a,b) TEM and (c,d) AFM images of CNC.

Sacui et al., 2014). in which a white gel appearance is observed for the as-isolated CNC
suspension, showing a good stability. It should be noted that the sta-
3.4. Zeta potential and aqueous colloidal stability of CNC bility of aqueous CNC suspensions is crucial in the preparation of
polymer based nanocomposites, especially when water soluble poly-
The zeta potential is an important parameter to evaluate the dis- mers are used as polymeric matrices and the nanocomposite systems are
persion state of CNC and its aqueous suspension stability. In this study, made by the solvent casting method.
the average value of zeta potential obtained for aqueous CNC suspen-
sion was found to be − 47.39 mV, which can be explained by effective 3.5. XPS analysis of CNC
electrostatic repulsions created by the presence of negatively charged
surface of CNC (Camarero Espinosa et al., 2013), restricting them to XPS analysis was carried out to better identify the presence of sul-
coagulate or flocculate, as schematically illustrated in Fig. 5. This result fate groups on the surface of CNC and to quantify the surface chemical
confirms that the as-extracted CNC suspension has a good stability, composition in terms of oxygen, carbon and sulfur. Fig. 6a shows the
because the absolute value is higher than 25 mV, as previously high- XPS survey spectra peaks of the as-extracted CNC. The peaks at 533.00,
lighted in our previous works (El Miri et al., 2016, 2015a,b). Fig. 5 286.44, and 169.07 eV generated from O 1s, C 1s, and S 2p, respectively
shows a photograph illustrating the CNC aqueous suspension stability, (Li et al., 2015; Thambiraj and Ravi Shankaran, 2017). The peak at
1071.8 which can be generated from Na 1 s was not detected con-
firming that the Na atoms were not presented in CNC sample. From the
survey spectra, the S 2p peak region was further determined using high
resolution analysis, and the obtained result is shown in Fig. 6b. The
atomic concentration was determined at 57.10 and 42.50 for C and O
respectively. The O/C ratio was calculated at 0.74, wherein 0.83 is the
theoretical value of pure cellulose (Zoppe et al., 2014). The atomic
concentration of S was found to be 0.40%, corresponding to 2.04 sulfate
groups per 100 anhydroglucose units, basing on the calculation pro-
cedure described elsewhere (Hamad and Hu, 2010; Li et al., 2015). This
result is comparable with that reported by Li et al. (2015) for sulfuric
acid hydrolyzed CNC.

3.6. XRD diffraction and FTIR analysis of raw and treated cellulosic fibers

Fig. 7a shows the XRD profiles of freeze dried CNC in comparison to


the raw and treated fibers (AF, ATAF and CMF). From this result, all
Fig. 5. Aqueous suspension of CNC at 5.30 mg/mL and schematic representation of cellulosic materials present the same cellulose I characteristic peaks at
possible electrostatic repulsions between indvidual CNC.
2θ of around 14.9°, 16.3° and 22.5°, corresponding to the cellulose I

504
M. El Achaby et al. Industrial Crops & Products 112 (2018) 499–510

Fig. 6. (a) XPS survey spectrum and (b) high-resolution of S2p photoelectron spectrum of CNC sample.

crystallographic planes 11̅0, 110 and 200, respectively (El Miri et al., 2013; Kargarzadeh et al., 2012; Rosa et al., 2010). The peaks at 1511
2015b; Flauzino Neto et al., 2013). As expected, the magnitude of the and 1246 cm−1 can be assigned to the C]C stretching vibrations from
observed crystalline peaks increased after different chemical treatments the aromatic ring and CeO out-of-plane-stretching-vibration of the aryl
because of the removal of amorphous lignin and hemicellulose group in lignin and hemicelluloses molecules (Kargarzadeh et al., 2012;
(Kargarzadeh et al., 2012). It is worth noting that the crystal structure Rosa et al., 2010). However, the mentioned peaks were not detected in
of cellulose component was not affected during the alkali, bleaching bleached CMF and hydrolyzed CNC spectra, suggesting that the lignin
and acid hydrolysis treatments. From the XRD data, the crystallinity and hemicelluloses were successfully removed after alkali and
index (CrI) was calculated using the Segal equation as previously de- bleaching treatments (Kargarzadeh et al., 2012; Mariano et al., 2016).
scribed in our works (El Miri et al., 2015b). The CrI was found to be Commonly, in all spectra the broad peaks observed at
about 42%, 56%, 71%, and 90% for raw AF, ATAF, CMF, and CNC 3370–3250 cm−1 was attributed to the stretching vibrations of CH and
samples, respectively. The increasing of CrI from raw AF to CMF was OH groups from the principal functional groups present in lig-
ascribed to the progressive removal of amorphous noncellulosic mate- nocellulosic materials (Barakat et al., 2014). Additionally, the peak
rials (lignin, hemicellulose and others) (El Miri et al., 2015b; detected at 1633 cm−1 is due to the bending mode of the absorbed
Kargarzadeh et al., 2012), and from CMF to CNC the increasing of CrI water (Barakat et al., 2014; Flauzino Neto et al., 2013). Besides, in all
was attributed to the hydrolysis of amorphous regions of CNC, resulting studied samples, the peak at 1159 cm−1 was associated to CeOeC
in CNC with relatively high crystallinity (90%). Thus, the as-obtained asymmetric stretching of the cellulose (Chen et al., 2016). Remarkably,
CNC exhibit a high crystallinity, which is a required property for using the intensity of this peak was gradually increased indicating that the
CNC as nanofillers for nanocomposites development (Paulo et al., cellulose content was increased after different chemical treatments
2013). starting from raw AF to the production of CNC. Moreover, the peaks
The FTIR is a very powerful technique to study the changes in the observed at 1033 and 899 cm−1 in all spectra could be attributed to the
chemical composition during different stages of treatment (Flauzino CeOeC pyranose ring skeletal vibration and the glycosidic C1eH de-
Neto et al., 2013; Marcos dos Santos et al., 2013). Fig. 7b shows the formation with ring vibration contribution and OH bending, respec-
FTIR spectra of raw AF, ATAF, CMF and CNC. For raw AF, the peak at tively, which is characteristic of β-glycosidic linkages between glucose
1729 cm−1 is attributed to the acetyl and uronic ester groups of in the cellulose (Chuan-Fu Liu et al., 2006). In addition, the peaks at
hemicellulose or the ester linkage of carboxylic group of ferulic and p- 1426 cm−1, 1370 cm−1 and 1315 cm−1 were contributed to the CH2
coumaric acids of lignin and/or hemicelluloses (Flauzino Neto et al., symmetric bending of cellulose, CeH asymmetric deformation and CeO

Fig. 7. (a) XRD patterns and (b) FTIR spectra of raw AF, ATAF, CMF and CNC samples.

505
M. El Achaby et al. Industrial Crops & Products 112 (2018) 499–510

undergo two step degradation process, which confirmed by the ob-


servation of two peaks in DTG curve, with Tmax1 and Tmax2 of 279 °C
and 340 °C. The first degradation stage of AF sample, which began at
Tonset of 217 °C, was attributed to the degradation of hemicelluloses and
lignin molecules. The second degradation stage (Tmax2 = 340 °C) can be
attributed to the degradation of cellulose (Mariano et al., 2016). It has
been reported that the degradation of cellulose starts with the genera-
tion of glucose units followed by the depolymerization and dehydration
process, and then completed by the oxidation and breakdown of the
charred residue to lower molecular weight gaseous products (Flauzino
Neto et al., 2013; Mariano et al., 2016; Trache et al., 2014).
It is noted that the degradation of ATAF and CMF samples follow
almost identical degradation mechanism, as demonstrated by TGA/DTG
curves (Tonset = 235 °C; Tmax = 354 °C), with Tonset is 18 °C higher than
that observed for raw AF sample (217 °C); indicating that the major
parts of hemicelluloses and lignin were removed after alkali treatment,
because their degradation at relatively low temperature was not de-
tected in ATAF sample, regarding to raw AF sample. The same trend
was reported for sugarcane bagasse and kenaf bast fibers after alkali
and bleaching treatments (El Miri et al., 2015b; Kargarzadeh et al.,
2012). Remarkably, the high percentage of residue at high tempera-
tures observed in raw AF as compared to that in the ATAF was due to
the presence of ash and lignin (Kargarzadeh et al., 2012).
For CNC two degradation stages were detected, which start at a low
Tonset of 170 °C, with Tmax1 and Tmax2 of 262 °C and 354 °C, respectively.
These results agreed with many conducted works in the literature for
CNC extracted by sulfuric acid hydrolysis (Kargarzadeh et al., 2012;
Mariano et al., 2016; Paulo et al., 2013). The first degradation me-
chanism (Tmax1 = 262 °C) of CNC can be explained by the presence of
sulfate groups on the surface of CNC, which are capable to catalyze the
primary dehydration phase in the cellulose component at the relatively
lower temperature. Afterwards, the second degradation stage occurred
at the higher temperature (Tmax2 = 354 °C) was attributed to degrada-
tion of cellulose chains that slowly degraded and formed char products.
In regard to the thermal degradation behavior of bleached cellulose
Fig. 8. (a) TGA and (b) DTG curves of AF, ATAF, CMF and CNC samples. (CMF) (Tonset = 235 °C), the thermal stability of CNC decreased as a
result of the presence sulfate groups, which considered as a major
symmetric stretching within the polysaccharide aromatic rings of cel- drawback for CNC extracted by sulfuric acid hydrolysis.
lulose and CeH bending of cellulose, respectively (Chen et al., 2016).
These peaks are observed in all spectra with increased intensity in the
case of bleached CMF and hydrolyzed CNC samples. Finally, the small 3.8. Processing of nanocomposite films
peak at 1202 cm−1 in the CNC spectrum is attributed to S]O vibration,
due to the insertion of sulfate groups on the surface of CNC during the The CNC were dispersed into Chitosan (CH), alginates (ALg) and k-
hydrolysis process (Lu and Hsieh, 2010). Consequently, the FTIR results carrageenan (k-CA) biopolymer matrices at various CNC loadings for
are in accord with other literatures, in which the extracted pure CMF objective to evaluate their ability to improve the tensile properties of
and acid hydrolyzed CNC showed a similar spectra (Chen et al., 2016; biopolymers and to produce new biopolymers based nanocomposite
Kargarzadeh et al., 2012; Lu and Hsieh, 2011, 2010; Mariano et al., films. CH, ALg and k-CA polysaccharides are the water soluble polymers
2016; Rosa et al., 2010). This suggesting that the produced pure CMF and their treatment in water can easily be achieved due to their hy-
and CNC have been successfully isolated from cellulose rich AF. drophilic nature. CH, AL and k-CA have also many functional groups,
such as hydroxyl groups (CH, ALg, CA), carboxylic groups (ALg), sulfate
groups (CA) and amino groups (CH) that could be useful to create
3.7. Thermal stability of raw and treated cellulosic fibers strong interfacial interactions between CNC and biopolymer matrices.
On the other hand, the sulfuric acid hydrolyzed CNC exhibit free hy-
TGA/DTG analysis was conducted under nitrogen atmosphere to droxyl and inserted anionic sulfate groups on their surfaces, which
evaluate the thermal degradation of all cellulosic materials at different make them highly water dispersible nanomaterials (Fig. 5). This com-
stages of treatment. The obtained results are shown in Fig. 8. It is well patibility between the CNC and these biopolymers can cause a good
known that the lignocellulosic materials degrade at low to moderate dispersion/distribution of CNC, and therefore driving the formation of a
temperature (Hajaligol et al., 2001). In general, the thermal decom- complex network in the final polymeric systems, thus improving the
position of such materials begins at lower temperature for hemi- properties of the resulted composite film materials. Accordingly, the
celluloses followed by an early stage of degradation of lignin molecules mixture of such polymers and CNC in water can be easily done in
and then decomposition of cellulose (Kargarzadeh et al., 2012; Mariano controlled conditions, enabling the formation of a homogeneous and
et al., 2016). stable aqueous CNC-biopolymers mixtures. By casting these latter on
From the results shown in Fig. 8 all of the samples showed a small petri dishes and evaporating of water, films with high quality, smooth
weight loss around 100 °C as a result of the evaporation of adsorbed surface, good flexibility, and about 70-μm-thick were produced, ex-
moisture, which is related to the hydrophilic nature of these cellulosic ample of these films are presented in Fig. 2.
materials (Kargarzadeh et al., 2012; Trache et al., 2014). The raw AF

506
M. El Achaby et al. Industrial Crops & Products 112 (2018) 499–510

elastic modulus. Herein the aim is to evaluate the ability of CNC from
Alfa fibers as nanoreinforcing fillers to improve the tensile properties of
some biopolymers (ALg, k-CA and CH). For that, the tensile properties
of neat ALg, k-CA and CH and its nanocomposite films, with different
CNC loadings (1, 3, 5 and 8 wt%), were investigated by uniaxial tensile
testing. The obtained typical stress–strain curves of all studied ALg-, k-
CA- and CH-based nanocomposite films are presented in Fig. 9a–c, re-
spectively, and the measured tensile parameters, such as Young’s
modulus, tensile strength, toughness and elongation at break are ploted
in Fig. 10a–c. The toughness is defined as the work of fracture, and was
calculated from the area under the stress–strain curve. From these data,
it can be seen that all nanocomposite films exhibited superior tensile
properties than those observed for neat CH, ALg and k-CA biopolymers,
thus confirming that the incorporation of CNC (up to 8 wt%) had a
positive effect on the improving of tensile properties of the selected
biopolymers.
For ALg-CNC nanocomposites, the modulus, strength and toughness
increased gradually with increasing of CNC content from 1 to 8 wt%.
Indeed, when 8 wt% CNC was added (i.e. ALg-CNC-8), the modulus,
strength and toughness were increased from 1310 to 2846 MPa (117%),
from 50 to 97 MPa (94%) and from 5.3 to 8.3 MJ/m3 (56%), respec-
tively (Fig. 10). In contrast, it have been reported that when high
content of CNC (6–8 wt%) was incorporated in ALg nanocomposite the
modulus and strength start to decrease in regard to nanocomposite
containing 4–5 wt% CNC. It was claimed that the increasing of these
tensile parameters is related to the agglomeration phenomenon of CNC
within the ALg matrix at high CNC contents (Huq et al., 2012; Wang
et al., 2017). Herein, in our research the incorporation of CNC at re-
latively high content (8 wt%) resulted in more increasing of the selected
tensile parameters in regard to 5 wt% CNC, which can be attributed to
the original source and high crystallinity of CNC and their inherent
properties, such as high aspect ratio. It has been reported that cellulose
derivatives are compatible with ALg biopolymer due to their similar
structures (Abdollahi et al., 2013; Huq et al., 2012), suggesting that the
incorporation of cellulose-based nanomaterial into ALg matrix can re-
sulted in good dispersion/distribution of the nanoreinforcing phase
within ALg biopolymer matrix, and thus improving the tensile proper-
ties of the resulted nanocomposite systems. Herein, the interesting
combination of improved modulus, strength and toughness in the ALg-
CNC nanocomposites is due to the nanoscale dispersion of CNC, and the
unique characteristics provided by the dispersed CNC. The same trend
was reported by Svagan et al. (2007) for microfibrillated cellulose
(MFC) filled glycerol-plasticized amylopectin nanocomposites, in which
the modulus, strength and toughness are together increased after the
addition of MFC fillers. The fine dispersion of CNC within ALg biopo-
lymer is ensured by strong interfacial hydrogen and ionic interactions
between the free hydroxyl groups, which are naturally available on the
surface of CNC, and the functional groups of ALg biopolymer, including
hydroxyl and carboxyl groups. Similarly, it have been reported that the
improving of tensile properties of CNC-reinforced ALg nanocomposites
is attributed to the interfacial hydrogen and ionic bonding between
both phases (Abdollahi et al., 2013; Wang et al., 2017).
The same trend was found for k-CA-CNC nanocomposite films, in
Fig. 9. Typical stress-strain curves of (a) ALg-CNC, (b) k-CA-CNC and (c) CH-CNC na-
which the modulus, strength and toughness were also increased with
nocomposite films at different CNC loadings (1, 3, 5 and 8 wt%). increasing of CNC content from 1 to 8 wt%. The addition of 8 wt% CNC
(i.e. k-CA-CNC-8) resulted in increasing of the modulus and strength
from 1510 to 3089 (104%) and from 48 to 118 (145%), respectively
3.9. Tensile properties of nanocomposite films
(Fig. 10). Furthermore, it should be noted that in regard to the neat k-
CA biopolymer, the k-CA-CNC nanocomposites not only have a higher
Biopolymer-based films usually suffers from low tensile properties,
strength and modulus but also a higher toughness that reached a
which limit their applications in various fields. However, these draw-
maximum value of 11.1 MJ/m3 for the k-CA nanocomposite containing
backs can be overcome by adding suitable nanoreinforcing fillers into
3 wt% CNC, which is higher than that determined for MFC filled gly-
biopolymers for enhancing their tensile properties. Indeed, it has been
cerol-plasticized amylopectin nanocomposites (9.4 MJ/m3) (Svagan
reported that CNC have a significant reinforcing impact on the tensile
et al., 2007). It is worth noting that no comparable results were pre-
properties of biopolymers (El Achaby et al., 2017; El Miri et al., 2015a;
viously reported in the literatures regarding CNC-filled k-CA nano-
Wang et al., 2017), due to their relatively large aspect ratio and high
composite films. Savadekar et al. (2012) prepared k-CA nanocomposites

507
M. El Achaby et al. Industrial Crops & Products 112 (2018) 499–510

Fig. 10. (a) Young’s modulus, (b) tensile strength, (c) toughness and (d) elongation at break of neat ALg, k-CA and CH and their nanocomposites with different contents of CNC (1, 3, 5
and 8 wt%).

filled with 0.1-1.0 wt% mechanically produced nanofibrated cellulose of the modulus, strength and toughness parameters (El Achaby et al.,
(NFC) (average diameter of 242 nm). They found that the strength was 2017). In regard to neat CH, the addition of 5 wt% CNC resulted in
increased by 44% with increasing of NFC content up to 0.4 wt% and increasing of the modulus and strength from 1392 to 2506 MPa (80%)
starts to decrease when 0.5-1.0 wt% NFC content was incorporated, due and from 46 to 92 MPa (100%), respectively. The toughness parameter
to the agglomeration phenomenon of NFC within k-CA biopolymer was reached its maximum value for the CH nanocomposite containing
matrix. Shankar et al. (2015) reported k-CA nanocomposite reinforced 3 wt% CNC. In these nanocomposite systems, the fact that 5 wt% CNC is
with 0.5–10 wt% hydrochloric acid hydrolyzed chitin nanofibrils (CNF) the optimum loading for increasing of modulus and strength was also
(average diameter of 5–10 nm and average length of 150–200 nm). reported in the literatures (Borysiak and Grząbka-Zasadzińska, 2016;
From their results, the strength was increased by 42% at 5 wt% CNF Khan et al., 2012). Khan et al. (2012) reported that the increase of the
content and then decreased with more addition of CNF (10 wt%); strength in CNC-filled CH nanocomposites was attributed to (i) the fa-
whereas the modulus was linearly increased with increasing of CNF vorable interactions between the macromolecular chains of CH and
content from 0 to 10 wt% (with maximum increase of 65%), attributing CNC, which is generated from the interactions between the anionic
this improvement to the formation of a percolating network based on sulfate groups of CNC and the cationic amine groups of CH, favoring a
hydrogen bonding interactions. good interface interaction between CH and CNC, in line with the work
Herein, the modulus, strength and toughness are together improved reported by de Mesquita et al. (2010), and (ii) the nanoreinforcing ef-
in the k-CA nanocomposites containing CNC, suggesting that CNC are fect of CNC occurred through effective stress transfer at CNC-CH in-
homogeneously dispersed within k-CA biopolymer. Might be the for- terface (Khan et al., 2012). This trend is in agreement with our previous
mation of hydrogen bonding in k-CA-CNC nanocomposite films are work when CNC from sugarcane bagasse were incorporated in CH-
responsible for creating interfacial interaction between CNC and k-CA Polyvinyl alcohol blend polymeric matrix (El Miri et al., 2015b).
biopolymer. Additionally, the presence of glycerol in these nano- On the contrary, the addition of CNC had a great impact on the
composite systems can help increase dispersion and interaction with elongation at break values for all prepared biopolymer-based nano-
CNC, as previously highlighted in the literature for MFC filled glycerol- composites (Fig. 10). With the incorporation of CNC, the elongation at
plasticized amylopectin nanocomposites (Svagan et al., 2007), hence, break dropped from 17%, 27% and 22% for neat ALg, k-CA and CH
the improvement of the tensile parameters of k-CA-CNC nanocompo- biopolymers to 12%, 12% and 11% for their nanocomposites containing
sites prepared with glycerol could be attributed to these explanations. 8 wt% CNC, respectively. The decrease of the elongation at break
For CH-based nanocomposite systems, the addition of 5 wt% CNC parameter with incorporating CNC nanofiller is a well known phe-
shows a more pronounced reinforcing effect than that of 8 wt% nomenon, which may be ascribed to the rigid behavior of CNC and their
(Fig. 10), which is related to the agglomeration phenomenon of CNC interfacial interactions with biopolymer matrices, which prevent the
within the CH biopolymer matrix, leading to a less pronounced increase motion of the macromolecular chains of biopolymers (Abdollahi et al.,

508
M. El Achaby et al. Industrial Crops & Products 112 (2018) 499–510

2013; Borysiak and Grząbka-Zasadzińska, 2016; Shankar et al., 2015; References


Wang et al., 2017).
Apparently, the special morphology (needle-like shape), large as- Abdollahi, M., Alboofetileh, M., Rezaei, M., Behrooz, R., 2013. Comparing physico-me-
pect ratio (66) and the functionalized surface of the as-produced sul- chanical and thermal properties of alginate nanocomposite films reinforced with
organic and/or inorganic nanofillers. Food Hydrocoll. 32, 416–424.
furic acid hydrolyzed CNC are responsible for the significant re- Bandera, D., Sapkota, J., Josset, S., Weder, C., Tingaut, P., Gao, X., Foster, E.J.,
inforcement impact on tensile properties of the studied biopolymers Zimmermann, T., 2014. Influence of mechanical treatments on the properties of
(ALg, k-CA and CH). Additionally, homogeneous and fine dispersion of cellulose nanofibers isolated from microcrystalline cellulose. React. Funct. Polym. 85,
134–141.
CNC along with favorable interfacial interactions between the func- Barakat, A., Gaillard, C., Steyer, J.-P., Carrere, H., 2014. Anaerobic biodegradation of
tional groups of CNC and those of biopolymer matrices is essential to cellulose–xylan–lignin nanocomposites as model assemblies of lignocellulosic bio-
achieve improvement in the final tensile properties of the as-developed mass. Waste Biomass Valorization 5, 293–304.
Belkhir, S., Koubaa, A., Khadhri, A., Ksontini, M., Nadji, H., Smiti, S., Stevanovic, T.,
biopolymer-based nanocomposite films. These ideal conditions result in 2013. Seasonal effect on the chemical composition of the leaves of Stipa tenacissima L
mechanically strong and flexible ecofriendly nanocomposite films, with and implications for pulp properties. Ind. Crops Prod. 44, 56–61.
a very interesting combination of high modulus, high strength and high Besbes, I., Vilar, M.R., Boufi, S., 2011. Nanofibrillated cellulose from Alfa, Eucalyptus and
Pine fibres Preparation, characteristics and reinforcing potential. Carbohydr. Polym.
toughness. From these very important results, CNC produced from raw
86, 1198–1206.
lignocellulosic AF have a good ability to improve the properties of Bian, J., Peng, F., Peng, X.-P., Xu, F., Sun, R.-C., Kennedy, J.F., 2012. Isolation of
biopolymers. hemicelluloses from sugarcane bagasse at different temperatures: structure and
properties. Carbohydr. Polym. 88, 638–645.
Borysiak, S., Grząbka-Zasadzińska, A., 2016. Influence of the polymorphism of cellulose
4. Conclusion on the formation of nanocrystals and their application in chitosan/nanocellulose
composites. J. Appl. Polym. Sci. 133, 42864.
Pure cellulose microfibers (CMF) with a diameter of 10 μm and yield Brahim, S.B., Cheikh, R.B., 2007. Influence of fibre orientation and volume fraction on the
tensile properties of unidirectional Alfa-polyester composite. Compos. Sci. Technol.
of 39% have been extracted from Alfa fibers (Stipa tenacissima plant) by 67, 140–147.
using alkali and bleaching treatments. After subjecting the as-extracted Buffiere, J., Balogh-Michels, Z., Borrega, M., Geiger, T., Zimmermann, T., Sixta, H., 2017.
CMF to sulfuric acid hydrolysis, cellulose nanocrystals (CNC) have been The chemical-free production of nanocelluloses from microcrystalline cellulose and
their use as Pickering emulsion stabilizer. Carbohydr. Polym. 178, 48–56.
successfully obtained with a yield of 57% (in regard to initial weight of Camarero Espinosa, S., Kuhnt, T., Foster, E.J., Weder, C., 2013. Isolation of thermally
CMF), a needle-like shape, a crystallinity of 90%, an aspect ratio of 66, stable cellulose nanocrystals by phosphoric acid hydrolysis. Biomacromolecules 14,
negatively charged surface (zeta potential of −41.39 mV and 2.04 1223–1230.
Carneiro-da-Cunha, M.G., Cerqueira, M.A., Souza, B.W.S., Carvalho, S., Quintas, M.A.C.,
sulfate groups per 100 anhydroglucose units) and a good aqueous col- Teixeira, J.A., Vicente, A.A., 2010. Physical and thermal properties of a chitosan/
loidal stability. Due to their excellent inherent properties and functio- alginate nanolayered PET film. Carbohydr. Polym. 82, 153–159.
nalized surface, the as-extracted CNC showed a high ability to improve Chen, Y.W., Lee, H.V., Juan, J.C., Phang, S.M., 2016. Production of new cellulose na-
nomaterial from red algae marine biomass Gelidium elegans. Carbohydr. Polym. 151,
the tensile properties of various biopolymers with different function-
1210–1219.
alities. The incorporation of 1, 3, 5 and 8 wt% CNC into alginate, k- Cudjoe, E., Hunsen, M., Xue, Z., Way, A.E., Barrios, E., Olson, R.A., Hore, M.J.A., Rowan,
carragennan and chitosan matrices, via solvent casting method, resulted S.J., 2017. Miscanthus Giganteus: a commercially viable sustainable source of cel-
in nanocomposite films with largely improved Young’s modulus, tensile lulose nanocrystals. Carbohydr. Polym. 155, 230–241.
Dafe, A., Etemadi, H., Zarredar, H., Mahdavinia, G.R., 2017. Development of novel car-
strength and toughness parameters. An increase of 117% and 94%, boxymethyl cellulose/k-carrageenan blends as an enteric delivery vehicle for pro-
104% and 145% and 60% and 85% were reached for Young’s modulus biotic bacteria. Int. J. Biol. Macromol. 97, 299–307.
and tensile strength in the case of alginate-, k-carragennan- and chit- de Mesquita, J.P., Donnici, C.L., Pereira, F.V., 2010. Biobased nanocomposites from layer-
by-layer assembly of cellulose nanowhiskers with chitosan. Biomacromolecules 11,
osan-based nanocomposite films containing 8 wt% CNC, respectively. 473–480.
This large improvement of tensile parameters was attributed to the El Achaby, M., Essamlali, Y., El Miri, N., Snik, A., Abdelouahdi, K., Fihri, A., Zahouily, M.,
homogeneous and fine dispersion of CNC along with favorable inter- Solhy, A., 2014. Graphene oxide reinforced chitosan/polyvinylpyrrolidone polymer
bio-nanocomposites. J. Appl. Polym. Sci. 131, 41042.
facial interactions between the functional groups of CNC and those of El Achaby, M., El Miri, N., Aboulkas, A., Zahouily, M., Bilal, E., Barakat, A., Solhy, A.,
biopolymer matrices, resulting in mechanically strong and flexible 2017. Processing and properties of eco-friendly bio-nanocomposite films filled with
ecofriendly nanocomposite films. cellulose nanocrystals from sugarcane bagasse. Int. J. Biol. Macromol. 96, 340–352.
El Miri, N., Abdelouahdi, K., Barakat, A., Zahouily, M., Fihri, A., Solhy, A., El Achaby, M.,
Through this study, we have demonstrated a possible strategy to
2015a. Bio-nanocomposite films reinforced with cellulose nanocrystals: rheology of
give an added value to raw Alfa plant which is a fast growing perennial film-forming solutions, transparency, water vapor barrier and tensile properties of
plant, rich in cellulose, inexpensive and renewable source. The ex- films. Carbohydr. Polym. 129, 156–167.
El Miri, N., Abdelouahdi, K., Zahouily, M., Fihri, A., Barakat, A., Solhy, A., El Achaby, M.,
traction of CNC, with excellent properties, from Alfa fibers can be
2015b. Bio-nanocomposite films based on cellulose nanocrystals filled polyvinyl al-
considered as moderately-feasible source. Indeed, the extracted CNC cohol/chitosan polymer blend. J. Appl. Polym. Sci. 132, 42004.
and the used method (sulfuric acid hydrolysis) are very attractive for El Miri, N., El Achaby, M., Fihri, A., Larzek, M., Zahouily, M., Abdelouahdi, K., Barakat,
the realization of a sustainable and economically viable bio-sourced A., Solhy, A., 2016. Synergistic effect of cellulose nanocrystals/graphene oxide na-
nosheets as functional hybrid nanofiller for enhancing properties of PVA nano-
nanomaterials for future growth of CNC applications in the industry. As composites. Carbohydr. Polym. 137, 239–248.
stated above, the raw Alfa fibers are abundant in nature, covering large Flauzino Neto, W.P., Silvério, H.A., Dantas, N.O., Pasquini, D., 2013. Extraction and
areas in North African countries, which can be ensure the availability of characterization of cellulose nanocrystals from agro-industrial residue –Soy hulls.
Ind. Crops Prod. 42, 480–488.
this source for its utilization for the production of CNC for various in- Garcia de Rodriguez, N.L., Thielemans, W., Dufresne, A., 2006. Sisal cellulose whiskers
dustrial applications. reinforced polyvinyl acetate nanocomposites. Cellulose 13, 261–270.
Grishkewich, N., Mohammed, N., Tang, J., Tam, K.C., 2017. Recent advances in the ap-
plication of cellulose nanocrystals. Curr. Opin. Colloid Interface Sci. 29, 32–45.
Acknowledgments Habibi, Y., Lucia, L.A., Rojas, O.J., 2010. Cellulose nanocrystals Chemistry, self-assembly,
and applications. Chem. Rev. 110, 3479–3500.
The financial assistance of the Office Chérifien des Phosphates (OCP Hajaligol, M., Waymack, B., Kellogg, D., 2001. Low temperature formation of aromatic
hydrocarbon from pyrolysis of cellulosic materials. Fuel 80, 1799–1807.
S.A.) in the Moroccan Kingdom toward this research is hereby ac-
Hamad, W.Y., Hu, T.Q., 2010. Structure-process-yield interrelations in nanocrystalline
knowledged. This work was performed as part of collaboration between cellulose extraction. Can. J. Chem. Eng. 88, 392–402.
the INRA of Montpellier and the Materials Science and Hanana, S., Elloumi, A., Placet, V., Tounsi, H., Belghith, H., Bradai, C., 2015. An efficient
enzymatic-based process for the extraction of high-mechanical properties alfa fibres.
Nanoengineering department (MSN) of the Mohamed 6 Polytechnic
Ind. Crops Prod. 70, 190–200.
University (UM6P) in Morocco. The authors would like to thank Dr. Huq, T., Salmieri, S., Khan, A., Khan, R.A., Le Tien, C., Riedl, B., Fraschini, C., Bouchard,
Cédric Gaillard from INRA-Nantes, France for their help to achieve this J., Uribe-Calderon, J., Kamal, M.R., Lacroix, M., 2012. Nanocrystalline cellulose
work. The authors would also like to thank Pr. Jones Alami, head of (NCC) reinforced alginate based biodegradable nanocomposite film. Carbohydr.
Polym. 90, 1757–1763.
MSN Department of UM6P, for his help to improve this work.

509
M. El Achaby et al. Industrial Crops & Products 112 (2018) 499–510

Junior De Menezes, A., Siqueira, G., Curvelo, A.A.S., Dufresne, A., 2009. Extrusion and Rosa, M.F., Medeiros, E.S., Malmonge, J.A., Gregorski, K.S., Wood, D.F., Mattoso, L.H.C.,
characterization of functionalized cellulose whiskers reinforced polyethylene nano- Glenn, G., Orts, W.J., Imam, S.H., 2010. Cellulose nanowhiskers from coconut husk
composites. Polymer 50, 4552–4563. fibers: effect of preparation conditions on their thermal and morphological behavior.
Kargarzadeh, H., Ahmad, I., Abdullah, I., Dufresne, A., Zainudin, S.Y., Sheltami, R.M., Carbohydr. Polym. 81, 83–92.
2012. Effects of hydrolysis conditions on the morphology, crystallinity, and thermal Sánchez-García, M.D., Hilliou, L., Lagarón, J.M., 2010. Morphology and water barrier
stability of cellulose nanocrystals extracted from kenaf bast fibers. Cellulose 19, properties of nanobiocomposites of κ/ι-hybrid carrageenan and cellulose nano-
855–866. whiskers. J. Agric. Food Chem. 58, 12847–12857.
Khan, A., Khan, R.A., Salmieri, S., Le Tien, C., Riedl, B., Bouchard, J., Chauve, G., Tan, V., Sacui, I.A., Nieuwendaal, R.C., Burnett, D.J., Stranick, S.J., Jorfi, M., Weder, C., Foster,
Kamal, M.R., Lacroix, M., 2012. Mechanical and barrier properties of nanocrystalline E.J., Olsson, R.T., Gilman, J.W., 2014. Comparison of the properties of cellulose
cellulose reinforced chitosan based nanocomposite films. Carbohydr. Polym. 90, nanocrystals and cellulose nanofibrils isolated from bacteria, tunicate, and wood
1601–1608. processed using acid, enzymatic, mechanical, and oxidative methods. ACS Appl.
Li, M.C., Wu, Q., Song, K., Lee, S., Yan, Q., Wu, Y., 2015. Cellulose nanoparticles: Mater. Interfaces 6, 6127–6138.
structure- morphology-rheology relationship. ACS Sustainable Chem. Eng. 3, Sapkota, J., Gooneie, A., Shirole, A., Martinez Garcia, J.C., 2017a. A refined model for the
821–832. mechanical properties of polymer composites with nanorods having different length
Liu, C.-F., Ren, J.-L., Xu, F., Liu, J.-J., Sun, J.-X., Sun, R.-C., 2006. Isolation and char- distributions. J. Appl. Polym. Sci. 134, 45279.
acterization of cellulose obtained from ultrasonic irradiated sugarcane bagasse. J. Sapkota, J., Martinez Garcia, J.C., Lattuada, M., 2017b. Reinterpretation of the me-
Agric. Food Chem. 54, 5742–5748. chanical reinforcement of polymer nanocomposites reinforced with cellulose na-
Liu, J., Zhan, X., Wan, J., Wang, Y., Wang, C., 2015. Review for carrageenan-based norods. J. Appl. Polym. Sci. 134, 45254.
pharmaceutical biomaterials: favourable physical features versus adverse biological Sapkota, J., Shirole, A., Foster, E.J., Martinez Garcia, J.C., Lattuada, M., Weder, C., 2017c.
effects. Carbohydr. Polym. 121, 27–36. Polymer nanocomposites with nanorods having different length distributions.
Lu, P., Hsieh, Y.-L., 2010. Preparation and properties of cellulose nanocrystals Rods, Polymer 110, 284–291.
spheres, and network. Carbohydr. Polym. 82, 329–336. Savadekar, N.R., Karande, V.S., Vigneshwaran, N., Bharimalla, A.K., Mhaske, S.T., 2012.
Lu, P., Hsieh, Y.-L., 2011. Preparation and characterization of cellulose nanocrystals from Preparation of nano cellulose fibers and its application in kappa-carrageenan based
rice straw. Carbohydr. Polym. 87, 564–573. film. Int. J. Biol. Macromol. 51, 1008–1013.
Maafi, E.M., Malek, F., Tighzert, L., Dony, P., 2010. Synthesis of polyurethane and Shankar, S., Reddy, J.P., Rhim, J.-W., Kim, H.-Y., 2015. Preparation, characterization,
characterization of its composites based on alfa cellulose fibers. J. Polym. Environ. and antimicrobial activity of chitin nanofibrils reinforced carrageenan nanocompo-
18, 638–646. site films. Carbohydr. Polym. 117, 468–475.
Malherbe, S., Cloete, T.E., 2002. Lignocellulose biodegradation: fundamentals and ap- Shankar, S., Wang, L.-F., Rhim, J.-W., 2016. Preparations and characterization of algi-
plications. Rev. Environ. Sci. Bio/Technol. 1, 105–114. nate/silver composite films: effect of types of silver particles. Carbohydr. Polym. 146,
Marcos dos Santos, R., Pires Flauzino Neto, W., Alves Silvério, H., Ferreira Martins, D., 208–216.
Oliveira Dantas, N., Pasquini, D., 2013. Cellulose nanocrystals from pineapple leaf, a Silvério, H.A., Pires, W., Neto, F., Oliveira Dantas, N., Pasquini, D., 2013. Extraction and
new approach for the reuse of this agro-waste. Ind. Crops Prod. 50, 707–714. characterization of cellulose nanocrystals from corncob for application as reinforcing
Mariano, M., Cercená, R., Soldi, V., 2016. Thermal characterization of cellulose nano- agent in nanocomposites. Ind. Crops Prod. 44, 427–436.
crystals isolated from sisal fibers using acid hydrolysis. Ind. Crops Prod. 94, 454–462. Siqueira, G., Bras, J., Follain, N., Belbekhouche, S., Marais, S., Dufresne, A., 2013.
Marrakchi, Z., Khiari, R., Oueslati, H., Mauret, E., Mhenni, F., 2011. Pulping and pa- Thermal and mechanical properties of bio-nanocomposites reinforced by Luffa cy-
permaking properties of Tunisian Alfa stems (Stipa tenacissima)—effects of refining lindrica cellulose nanocrystals. Carbohydr. Polym. 91, 711–717.
process. Ind. Crops Prod. 34, 1572–1582. Sofla, M.R.K., Brown, R.J., Tsuzuki, T., Rainey, T.J., 2016. A comparison of cellulose
Marrakchi, Z., Oueslati, H., Belgacem, M.N., Mhenni, F., Mauret, E., 2012. Biocomposites nanocrystals and cellulose nanofibres extracted from bagasse using acid and ball
based on polycaprolactone reinforced with alfa fibre mats. Compos. Part A Appl. Sci. milling methods. Adv. Nat. Sci.: Nanosci. Nanotechnol. 7, 35004.
Manuf. 43, 742–747. Svagan, A.J., Azizi Samir, M.A.S., Berglund, L.A., 2007. Biomimetic polysaccharide na-
Martins, D.F., de Souza, A.B., Henrique, M.A., Silvério, H.A., Neto, W.P.F., Pasquini, D., nocomposites of high cellulose content and high toughness. Biomacromolecules 8,
2015. The influence of the cellulose hydrolysis process on the structure of cellulose 2556–2563.
nanocrystals extracted from capim mombaça (Panicum maximum). Ind. Crops Prod. Tan, X.Y., Abd Hamid, S.B., Lai, C.W., 2015. Preparation of high crystallinity cellulose
65, 496–505. nanocrystals (CNCs) by ionic liquid solvolysis. Biomass Bioenergy 81, 584–591.
Nadji, H., Diouf, P.N., Benaboura, A., Bedard, Y., Riedl, B., Stevanovic, T., 2009. Teixeira, E. de M., Bondancia, T.J., Teodoro, K.B.R., Corrêa, A.C., Marconcini, J.M.,
Comparative study of lignins isolated from Alfa grass (Stipa tenacissima L.). Mattoso, L.H.C., 2011. Sugarcane bagasse whiskers: extraction and characterizations.
Bioresour. Technol. 100, 3585–3592. Ind. Crops Prod. 33, 63–66.
Olsson, R.T., Kraemer, R., López-Rubio, A., Torres-Giner, S., Ocio, M.J., Lagarón, J.M., Thambiraj, S., Ravi Shankaran, D., 2017. Preparation and physicochemical character-
2010. Extraction of microfibrils from bacterial cellulose networks for electrospinning ization of cellulose nanocrystals from industrial waste cotton. Appl. Surf. Sci. 412,
of anisotropic biohybrid fiber yarns. Macromolecules 43, 4201–4209. 405–416.
Omri, M.A., Triki, A., Guicha, M., Hassen, M., Ben Arous, M., Bulou, A., 2016. Effect of Trache, D., Donnot, A., Khimeche, K., Benelmir, R., Brosse, N., 2014. Physico-chemical
wool fibers on thermal and dielectric properties of Alfa fibers reinforced polyester properties and thermal stability of microcrystalline cellulose isolated from Alfa fibres.
composite. Mater. Chem. Phys. 170, 312–318. Carbohydr. Polym. 104, 223–230.
Paiva, M.C., Ammar, I., Campos, A.R., Cheikh, R.B., Cunha, A.M., 2007. Alfa fibres Trache, D., Hussin, M.H., Hui Chuin, C.T., Sabar, S., Fazita, M.R.N., Taiwo, O.F.A.,
Mechanical, morphological and interfacial characterization. Compos. Sci. Technol. Hassan, T.M., Haafiz, M.K.M., 2016. Microcrystalline cellulose: isolation, character-
67, 1132–1138. ization and bio-composites application—a review. Int. J. Biol. Macromol. 93,
Paulo, J. Morais S., Rosa, M. De F., De Sá Moreira De Souza Filho, M., Dias Nascimento, 789–804.
L., Magalhães Do Nascimento, D., Cassales, A.R., 2013. Extraction and character- Trache, D., Hussin, M.H., Haafiz, M.K.M., Thakur, V.K., 2017. Recent progress in cellulose
ization of nanocellulose structures from raw cotton linter. Carbohydr. Polym. 91, nanocrystals: sources and production. Nanoscale 9, 1763–1786.
229–235. Wang, L.-F., Shankar, S., Rhim, J.-W., 2017. Properties of alginate-based films reinforced
Phanthong, P., Karnjanakom, S., Reubroycharoen, P., Hao, X., Abudula, A., Guan, G., with cellulose fibers and cellulose nanowhiskers isolated from mulberry pulp. Food
2017. A facile one-step way for extraction of nanocellulose with high yield by ball Hydrocoll. 63, 201–208.
milling with ionic liquid. Cellulose 24, 2083–2093. Yao, K., Meng, Q., Bulone, V., Zhou, Q., 2017. Flexible and responsive chiral nematic
Pickering, K.L., Efendy, M.G.A., Le, T.M., 2016. A review of recent developments in cellulose nanocrystal/poly(ethylene glycol) composite films with uniform and tun-
natural fibre composites and their mechanical performance. Compos. Part A Appl. able structural color. Adv. Mater. 29, 1701323.
Sci. Manuf. 83, 98–112. Yu, C., 2015. Chapter 2 –natural textile fibres: vegetable fibres. Textiles and Fashion. pp.
Rangan, A., Manchiganti, M.V., Thilaividankan, R.M., Kestur, S.G., Menon, R., 2017. 29–56. http://dx.doi.org/10.1016/B978-1-84569-931-4.00002-7.
Novel method for the preparation of lignin-rich nanoparticles from lignocellulosic Zoppe, J.O., Ruottinen, V., Ruotsalainen, J., Rönkkö, S., Johansson, L.-S., Hinkkanen, A.,
fibers. Ind. Crops Prod. 103, 152–160. Järvinen, K., Seppälä, J., 2014. Synthesis of cellulose nanocrystals carrying tyrosine
Remminghorst, U., Rehm, B.H.A., 2006. Bacterial alginates: from biosynthesis to appli- sulfate mimetic ligands and inhibition of alphavirus infection. Biomacromolecules
cations. Biotechnol. Lett. 28, 1701–1712. 15, 1534–1542.

510

Вам также может понравиться