Carbohydrate Polymers 211 (2019) 181-194

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Carbohydrate Polymers 211 (2019) 181–194

Contents lists available at ScienceDirect

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

Cellulose nanocrystals reinforced κ-carrageenan based UV resistant T


transparent bionanocomposite films for sustainable packaging applications
Mithilesh Yadava, Fang-Chyou Chiua,b,

a
Department of Chemical and Materials Engineering, Chang Gung University, Taoyuan, 333, Taiwan
b
Department of General Dentistry, Chang Gung Memorial Hospital, Taoyuan, 333, Taiwan

ARTICLE INFO ABSTRACT

Keywords: In this work, κ-carrageenan bionanocomposite films were prepared by solution casting of a mixture of κ-car-
Fourier-transform infrared spectral analysis rageenan, glycerol, and various amounts of cellulose nanocrystals (CNCs, 0–9 wt.%). The structure and mor-
Mechanical property phology of the bionanocomposite films were characterized by Fourier-transform infrared spectroscopy, X-ray
Nanocomposite diffraction, transmission electron microscopy, and scanning electron microscopy. Compared with κ-carrageenan
Cellulose nanocrystal
films, the κ-carrageenan bionanocomposite films showed better mechanical and barrier properties (water and
UV) and thermal stability. The water contact angle increased from 23.30° to 71.80° and the water vapor per-
meation decreased from 8.93 gm−1 s−1 Pa−1 to 4.69 × 10−11 gm−1 s−1 Pa−1 in the κ-carrageenan films loaded
with 9–7 wt.% CNCs, respectively. The tensile strength and elongation at break of the films increased from
38.33 ± 3.79 MPa to 52.73 ± 0.70 MPa and from 21.50 ± 3.72% to 28.27 ± 2.39%, respectively, after CNC
loading increased from 0 wt.% to 7.0 wt.%. These results indicated that the κ-carrageenan nanocomposite films
have potential applications in food packaging.

1. Introduction for preventing dehydration (Karbowiak, Debeaufort, Champion, &


Voilley, 2006). Park (1996) found that κ-carrageenan transparent films
Petrochemical-based plastic films composed of polyethylene, poly- show a higher tensile strength (TS) than ι- and γ-carrageenan films. Fan,
vinylchloride (PVC), polypropylene (PP), and polystyrene (PS) are Peng, Li, Wang, and Wang (2013) concluded that the optimal TS and
widely used as packaging materials because of their low cost, light elongation at break (EB) of blend films of κ-carrageenan and alginate
weight, good mechanical behavior, water resistance, and heat seal can be achieved when the κ-carrageenan content is 30 wt.%. Lafargue,
ability. However, these materials are conditioned first before use be- Lourdin, and Doublier (2007) reported the mechanical properties of κ-
cause of their non-biodegradable nature (Sorrentino, Gorrasi, & carrageenan/pea starch blend films. Larotonda, Torres, Goncalves,
Vittoria, 2007). Therefore, researchers have developed packaging ma- Sereno, and Hilliou (2016) reported that the thin, flexible, and trans-
terials (Youssef & El- Sayed, 2018; Youssef et al., 2018; Youssef, EL- parent films of κ-carrageenan/rice starch show higher rheological, ul-
Sayed, Salama, EL-Sayed, & Dufresne, 2015; Youssef, EL-Sayed, EL- traviolet (UV) barrier, oxygen barrier, and hydrophobic properties than
Sayed, Salama, & Dufresne, 2016) from sustainable biomass sources commercial κ-carrageenan. Park, Lee, Jung, and Park (2001) reported
that are environmentally friendly, inexpensive, light weight, possess that adding ascorbic acid can affect the TS, EB, and water vapor per-
good thermomechanical properties, and provide a reasonable barrier to meability (WVP) of κ-carrageenan/chitosan blend. However, the ap-
the transfer of liquids and gases (Khan et al., 2010). Natural films based plications of κ-carrageenan films are limited by their poor mechanical,
on polysaccharides are biodegradable films that have been widely thermal, and barrier properties. Thus, κ-carrageenan has been re-
studied (Dou, Zhang, He, Yin, & Cui, 2016; Xie et al., 2019). Carra- inforced by preparing nanocomposites with various nanofillers. Wahab
geenans are sulfated polysaccharides that are derived from the cell and A B D Razak (2016) revealed that κ-carrageenan/clay bionano-
walls of red sea weeds (Roy, Shankar, & Rhim, 2019) and are composed composite films exhibit optimum tensile properties at 3 wt.% of HNT.
of a linear chain of sulfated galactans. Carrageenans are categorized Another study found that κ-carrageenan/SiO2 films as a packaging
into three types according to their sulfate content: κ (20%), ι (33%), and material demonstrate enhanced mechanical and antimicrobial proper-
γ (41%). Carrageenans show a high potential as a film-forming mate- ties (Venkatesan, Rajeswari, & Thiyagu, 2017). Compared with κ-car-
rial. Edible carrageenan films can be applied in meat, poultry, and fish rageenan films, κ-carrageenan/ZnO films show higher mechanical and


Corresponding author at: Department of Chemical and Materials Engineering, Chang Gung University, Taoyuan, 333, Taiwan.
E-mail address: maxson@mail.cgu.edu.tw (F.-C. Chiu).

https://doi.org/10.1016/j.carbpol.2019.01.114
Received 12 December 2018; Received in revised form 29 January 2019; Accepted 31 January 2019
Available online 01 February 2019
0144-8617/ © 2019 Elsevier Ltd. All rights reserved.
M. Yadav, F.-C. Chiu Carbohydrate Polymers 211 (2019) 181–194

solubility properties in food packaging applications (Saputri, experiments.


Praseptiangga, Rochima, Panatarani, & Joni, 2018). Rhim (2012)
evaluated the mechanical, water vapor barrier, and water resistance 2.2. Synthesis of CNCs
properties of κ-carrageenan/agar blend films prepared by solution
casting with and without nanoclay. CNCs were synthesized using sulfuric acid hydrolysis (Korolovych
Although numerous studies reported on κ-carrageenan nano- et al., 2018). Cellulose microcrystalline (CMC, 3 g) was hydrolyzed with
composites with different types of fillers, few used cellulose nanocrys- sulfuric acid solution (20 mL, 64 wt.%) at 48 ℃. The resultant suspen-
tals (CNC) as a nanofiller to reinforce κ-carrageenan. Cellulose is the sion was added to 200 mL of cold water (4 ℃). The reaction was per-
structural component of the primary cell wall of green plants. It con- formed for 60 min under continuous magnetic stirring and then stopped
stitutes the most abundant renewable polymer resource on earth. Two by adding 200 mL of cold water (4 ℃). After the reaction product
types of nanocellulose can be obtained from cellulose: whiskers and precipitated, the supernatant was decanted. This suspension was wa-
nanofibrils (Azizi Samir, Alloin, & Dufresne, 2005). The term “whis- shed by centrifugation at 10,000 rpm for 5 min at room temperature.
kers” (or “nanowhiskers/nanocrystals”) refers to nanoparticles that The washing was repeated three times before the product was dialyzed
exhibit a needle-like appearance, whereas the term “nanofibrils” (or in millipore water. The millipore water was changed daily until the pH
“microfibrillated cellulose”) refers to long and flexible nanoparticles reached 7. Subsequently, the suspension was subjected to ultra-soni-
composed of alternating amorphous and crystalline domains (Khalil, cation (DELTA Ultrasonic Cleaner DC300 H) for 1 h at an operating
Bhat, & Yusra, 2012). CNCs as sustainable nanomaterials show abun- frequency of 40 kHz and a power dissipation of 300 W to break down
dance, renewability, and biodegradability, high mechanical properties, the unfractionated cellulose into small nanoparticles. Finally, the
high reinforcing capability, and low density (Rescignano et al., 2014). sample was freeze dried. The final product was CNCs with a length of
As the main components of cellulose fibers (Veluswamya et al., 2016, 200–500 nm and a diameter of 20–50 nm. Details about the preparation
2017), the crystalline and amorphous parts can be generally separated of CNCs are discussed in S1. The obtained yield of the CNCs was ap-
by mechanical and chemical processes. The amorphous regions of cel- proximately 65%.
lulose can be broken under controlled conditions to liberate colloidal
particles known as CNCs. This process is achieved by acid hydrolysis, 2.3. Synthesis of κ-carrageenan/CNC nanocomposite films
which entails the destruction of the surrounding amorphous regions
between the microfibrils of cellulose while the crystalline segments In accordance with S2, κ-carrageenan/CNC films were fabricated as
remain intact. This breakdown of CNCs is attributed to the fact that the follows. The κ-carrageenan solution was prepared by dissolving 1 g of κ-
kinetics of hydrolysis is faster in the amorphous region than in the carrageenan powder in 100 mL of distilled water under magnetic stir-
crystalline domains because of the higher permeability of the former ring for 30 min. The glycerol plasticizer (25% of κ-carrageenan) was
than the latter (Azizi Samir et al., 2005). The geometric dimensions then mixed with the κ-carrageenan solution while stirring for another
depend on the initial source of cellulose, which accounts for widths 30 min. Nanocomposite samples were fabricated by dispersing different
(5–20 nm) and lengths (100 nm–2 μm). CNCs do not flocculate with amounts of CNCs (0, 1, 3, 5, 7, and 9 wt.%) in 100 mL of distilled water
water because of the electrostatic repulsion arising from the surface, for 30 min at room temperature. The obtained suspension was added to
leading to suspensions that are stable for several months (Tingaut, the κ-carrageenan solution, stirred for 1 h at room temperature, and
Zimmermann, & Sèbe, 2012). Azizi and Mohamad (2018) studied the then sonicated for 20 min at 25 ℃ in a bath-type ultrasound sonicator.
effect of CNC loading on the mechanical and barrier properties of κ- The κ-carrageenan/CNC suspensions were then poured into a glass petri
carrageenan bionanocomposite films. The addition of 4% CNC in κ- dish and dried at 40 ℃ for 24 h until the solvent was completely eva-
carrageenan improves the mechanical properties of biodegradable porated and a self-standing film was achieved. The resultant κ-carra-
composite films based on κ-carrageenan reinforced by CNCs from kenaf geenan/CNC nanocomposite film with 0, 1, 3, 5, 7, and 9 wt.% loading
fibers (Zarina and Ahmad, 2015). Garcia, Hilliou, and Lagaron (2010) of CNCs were coded as CNC0, CNC1, CNC3, CNC5, CNC7, and CNC9,
concluded that κ-carrageenan/cellulose bionanocomposites containing respectively. The thickness of the films (0.020–0.040 mm) was mea-
3 wt.% cellulose nanowire exhibit the lowest reduction in WVP, that is, sured with the help of a Teclock dial thickness gauge (SM-112, Japan),
ca. 71%, and largely attributed this reduction to a filler-induced water and triplicate measurements were taken.
solubility reduction. However, the effect of CNCs on the UV properties
of carrageenan was not analyzed. Many studies focused on κ-carra- 2.4. Fourier-transform infrared spectroscopy (FTIR)
geenan/CNC nanocomposites (Azizi & Mohamad, 2018; Garcia et al.,
2010; Zarina & Ahmad, 2015), but information about the UV barrier FTIR spectra were recorded using a BRUKER, TENSOR 27 IR spec-
properties of κ-carrageenan/CNC films remains lacking to date. Thus, trometer. The samples were analyzed at a range of 400–4500 cm−1.
the present study characterized the UV barrier properties of κ-carra- Prior to the experiment, the samples were dried in an air circulation
geenan/CNC films and fabricated κ-carrageenan/CNC bionanocompo- oven at 40 ℃.
site films by using a simple solution mixing technique. The effects of
CNCs on the water solubility, water absorbency, moisture content, and 2.5. Field emission scanning electron microscopy (FESEM)
mechanical and thermal properties of κ-carrageenan films were also
systematically explored. Results showed that the synthesized κ-carra- The surface morphology and cross section of the prepared samples
geenan/CNC nanocomposite films can serve as sustainable packaging were examined via SEM (Jeol JSM-7500F).
materials because of their enhanced physico-chemical properties.
2.6. Transmission electron microscopy (TEM)
2. Materials and methods
Carbon film-covered copper grids (5–6 nm film thickness, 200 mesh,
2.1. Materials EMS FF200) were used for imaging CNC samples. One drop of the CNC
suspension (0.1 mg/mL) was placed on a grid for 30 min and wicked
Glycerol and κ-carrageenan were purchased from Sigma Aldrich and away with a filter paper. Further, the samples were stained by depos-
used as received. Ultra-fine microcellulose (ARBOCEL UFC) was pur- iting a drop of (2 wt.%) phosphotungstic acid solution on the grid for
chased from J. RETTENMAIER & SÖHNE France (JRS). Sulfuric acid 20 min and wicking the excess solution away with a wet filter paper.
(Merck) was used to maintain acid concentration. The other chemical The grid was dried in the oven at 70 ℃ before insertion into the mi-
reagents were of analytical grade. Millipore water was used for all croscope. Finally, the CNCs were analyzed by a JEOL (JEM-1230

182
M. Yadav, F.-C. Chiu Carbohydrate Polymers 211 (2019) 181–194

electron microscope, Tokyo, Japan) operated at an accelerating voltage Seongcheol, Hyun, & Kim, 2015):
of 200 kV.
Ws Wd
Swelling Ratio (%) = Sr(%) = × 100,
Wd (3)
2.7. Optical microscopy (OM)
Where, Ws is the weight of the swollen samples and Wd is the weight of
The selected samples were dried in an air-circulating oven for 6 h at the dry samples. All measurements were performed in three replicates.
40 ℃ and used for microscopic analysis. An optical microscope
(Olympus BX-50) was used to examine the phase morphology and 2.13. Film water solubility (FWS)
dispersion status of the nanofillers.
The solubility of the films was calculated as the percentage of dis-
2.8. X-ray diffraction (XRD) solved dry matter after immersion in water in accordance with a pre-
viously described method (Orsuwan, Shankar, Wang, Sothornvit, &
XRD was measured by an X-ray diffractometer (D2, Bruker, Rhim, 2016). For this calculation, film samples were sectioned into
Germany). The X-ray source was CuKα radiation operating at 40 kV and 2 × 2 cm2 pieces and dried at room temperature for 24 h to determine
40 mA. The degree of crystallinity was calculated using the following the initial dry weight (Wi). Then, the film samples were immersed in
equation (Segal, Creely, Martin, & Conrad, 1959): 20 mL of distilled water with mild shaking, removed, and then dried at
40 ℃ for 24 h to determine the undissolved final dry weight (Wf). FWS
I002 Iam
CI(%) = × 100,
(1) was calculated using the following equation (Noshirvani,
I002
Ghanbarzadeh, Fasihi, & Almasi, 2016):
where I002 is the intensity of the crystalline region of cellulose
Wi Wf
(2θ = 22.5°) and Iam is the intensity of the amorphous region FWS= × 100.
Wi (4)
(2θ = 16.3°).

2.9. Thermogravimetric analysis (TGA) 2.14. Moisture absorption

The thermal stability of the samples was characterized using a Moisture absorption was measured as described by Almasi,
thermogravimetric analyzer on a TA Q50 system. The samples were Ghanbarzadeh, and Entezami (2010). Dried CNC0, CNC1, CNC3, CNC5,
scanned from room temperature to 700 ℃ at a heating rate of 10 ℃ CNC7, and CNC9 films of size 2 × 2 cm2 were preconditioned at 0%
/min under a nitrogen environment. relative humidity (RH) for 24 h. After weighing, they were conditioned
in a humid chamber (Giant Force, Taiwan) at 25 ℃ to ensure 98% RH.
2.10. Density The samples were weighed at desired intervals of time until the equi-
librium state was reached. The moisture absorption of the samples was
Film density was determined from the specimen weight and volume. calculated as follows (Noshirvani et al., 2016):
The specimen was weighed using an analytical balance (A & D Co., Ltd.,
Wt Wo
Japan) with a precision of 0.1 mg. The specimen volume was calculated Moisture absorption(%) = × 100,
(5)
Wo
from specimen area and thickness. The thickness was measured using a
Teclock dial thickness gauge (SM-112, Japan) with a precision of Where, Wo is the initial weight of the samples and Wt is the weight
0.01 mm at five different positions in each specimen, and the average of the samples after time at 98% RH. All measurements were performed
values were taken. The densities of the CNC0, CNC1, CNC3, CNC5, in three replicates.
CNC7, and CNC9 films are presented in Table 2. The data were calcu-
lated as the mean value of three specimens of each film type. All films 2.15. Water vapor permeation (WVP)
showed almost the same density, suggesting that the incorporation of
CNCs did not significantly affect the density of the films. This result can The WVP of the κ-carrageenan nanocomposite films was determined
be attributed to the low amount of loaded CNCs. by the gravimetric method in accordance with the ASTM-E96/E96-05
Standard (ASTM-E96/E96-05, 2005). The water method described in
2.11. Opacity and UV visibility this standard was implemented in the current work. Petri dishes of 6 cm
diameter were used as test dishes. In each test dish, 20 mL of distilled
Opacity was determined by measuring the film absorbance at water were added, leaving a distance of approximately 1 cm between
600 nm using a UV spectrophotometer (JASCO, V-650) and calculating the water surface and the film. The film samples were sealed to the dish
by the following equation: mouth by a water-resistant sealant. The dishes were placed in a con-
trolled-environment room with 30 ℃ temperature and 78% RH, and
Opacity = Abs600/d (2) their weight was measured at intervals of 12 h over a 48 h period on a
Where, Abs600 is the value of absorbance at 600 nm and d is the film balance. The weight of each dish linearly decreased with time. The
thickness (mm). WVP was calculated using Eq. (6). The parameter weight loss/time was
obtained from the slope of weight loss vs. time plot. The WVP of the
2.12. Water absorbency film was calculated using the following formula:

WVP (g m−1 s−1 Pa−1) = (w/t). γ. (A)−1 (Δp)−1, (6)


Water absorbency of the bionanocomposite films was tested in
terms of swelling. Film samples were sectioned into 2 × 2 cm2 pieces where w is the weight loss of the Petri dishes (g), γ is the film thickness
and pre-weighed (5 mg). Each sample was immersed in a petri dish (m), A is the cross- section area of the film (m2), t is the time (s), and Δp
filled with 20 mL of deionized water and remained undisturbed for 24 h is the vapor pressure difference (3167.2 Pa at 25 ℃). Each sample was
at room temperature until an equilibrium swelling was reached. The measured in triplicate.
swollen samples were removed from deionized water, quickly wiped
with filter paper to remove droplets on the surface, and then weighed. 2.16. Water contact angle
Tests were repeated four times to minimize error. The swelling ratio
(Sr) can be calculated using the following expression (Yadav, The hydrophilicity of the polymer blend and polymer

183
M. Yadav, F.-C. Chiu Carbohydrate Polymers 211 (2019) 181–194

Fig. 1. FTIR spectra of CMC, CNC, CNC0, κ-carrageenan and CNC9.

nanocomposites films was evaluated by contact angle measurements 3. Results and discussion
using a Drop Shape Analysis System (model DIGIDROP GBX instrument,
France). Deionized water was used to evaluate the hydrophilic prop- 3.1. Preparation of cellulose nanocrystals, homogenous suspension, and κ-
erties of these films. Angle measurements were performed in triplicate, carrageenan/CNC nanocomposite films
and their average values were taken.
Strong acid hydrolysis with H2SO4 is usually performed prior to or
after mechanical, high-pressure homogenizer, ultrasonic, or hydro-
2.17. Mechanical properties dynamic cavitation treatment (Majoinen, Kontturi, Ikkala, & Gray,
2012) to yield CNCs. The hydrolysis process basically involves re-
Tensile tests of the CNC0, CNC1, CNC3, CNC5, CNC7 and CNC9 moving the amorphous regions present in the CNCs, leaving the crys-
films (according to ASTM D638) were conducted at a crosshead speed talline regions intact. Hence, the dimensions of the cellulose whiskers
of 10 mm/min by using a Gotech AI-3000 system. obtained after hydrolysis are mainly dependent on the percentage of
amorphous regions that varies for each organism. The fabrication of
renewable biopolymer films by a process other than casting is not re-
ported due to its thermosensitivity. Other film-forming procedures,

184
M. Yadav, F.-C. Chiu Carbohydrate Polymers 211 (2019) 181–194

such as slit-die extrusion, blown-film extrusion, and calendaring, in- separation, indicating good adhesion between the κ-carrageenan and
volve high temperatures, which consequently lead to undesirable re- the CNCs. This adhesion was found due to the presence of hydrogen
actions such as biopolymer degradation. bonding and was beneficial to the mechanical properties of the nano-
composite films (Kang et al., 2018). The TEM images also confirmed the
3.2. Fourier transform infrared spectroscopy good mixing of the CNCs in the κ-carrageenan matrix. As shown in
Fig. 3(e), the rod-like CNCs were well distributed in the κ -carrageenan/
Many studies (Turquois, Acquistapace, Vera, & Welti, 1996) have CNC films. However, some agglomerates were observed in the micro-
performed infrared spectroscopy of polysaccharides. The FTIR tech- structure as the CNC concentration reached 9 wt.%. This agglomeration
nique is widely used to investigate the miscibility of renewable poly- was found due to the presence of hydrogen-bonded free hydroxyl
mers and nanomaterials due to its rapid, non-destructive, and powerful groups within the CNCs. Similar results were reported in previous re-
nature. When chemical functional groups interact, changes such as search papers (Garcia et al., 2010). The uniform and homogeneous
shifting of absorption bands can be detected in FTIR spectra. These distribution of the CNCs in the nanocomposite film is responsible for the
changes can indicate good miscibility of polymers. The FTIR spectra of enhanced physico-mechanical properties of the film.
the fabricated CNCs showed that the sharp peak at 3350 cm−1 might be
due to the OeH vibrations in hydrogen bonds (Li & Renneckar, 2011). 3.4. XRD
The absorption bands between 2800–3000 cm−1 originated from C–H
stretching and bending vibrations (Wang & Roman, 2011). Some other To confirm the presence of crystalline, CMC, CNC, κ-carrageenan,
absorption bands were assigned at 567 (OeH out of plane bending vi- CNC0, CNC1, CNC3, CNC5, CNC7, and CNC9 were characterized by
brations), 1058 (C–O stretching at C-3 position), and 1163 cm−1 XRD, and the diffractogram is shown in Fig. 4. XRD patterns were
(C–OeC stretching motion) (Nikonenko, Buslov, Sushko, & Zhbankov, collected from 10° to 60° (2θ). The CMCs exhibited characteristic
2005). Aside from these peaks, the CNCs showed two more peaks at crystalline peaks at around 2θ = 22.5°, 15.2°, 16.4°, and 34.8°. How-
1645 and 1240 cm−1, confirming that the acidification of the CMCs was ever, the CNCs showed four characteristic crystalline peaks at around
successfully performed (Lu & Hsieh, 2010). Furthermore, the abundant 2θ = 22.9°, 15.5°, 16.3°, and 34.7°, which corresponded to (2 0 0), (−1
oxygen functional groups rendered the CNCs hydrophilic, which im- 1 0), (1 1 0), and (4 0 0) lattice planes, respectively, which is a char-
proved their solubility in water. In κ-carrageenan (Sen & Erboz, 2010), acteristic of cellulose I structure (Khilil, Borges, Almeida, Boukherroub,
the characteristic peaks at 3362 cm−1 (OeH stretching vibration), & Omrani, 2018). The peak at 22.9° was broader in the CNCs than in the
2978 cm−1 (CeH stretching vibration) (Cerqueira et al., 2011), CMCs, indicating that the crystallite size of the CNCs was smaller than
1279 cm−1 (presence of ester sulfate group), 912 cm−1 (3, 6-anhy- that of the CMCs. The crystallinity index (CI) of the CMCs and CNCs
drogalactose group), and 835 cm−1 (galactose-4-sulfate group) were were 0.49 and 0.52, respectively. The increase in the CI of the CNCs was
assigned. After adding glycerol in κ-carrageenan (κ-carrageenan con- probably due to the acid treatment of the CMCs.
taining glycerol = CNC0), the peaks at 3362, 2978, 1279, 912, and 835 The κ-carrageenan and CNC0 did not show any clear peaks in XRD,
cm−1 shifted to 3342, 2982, 1288, 943, and 867 cm−1, suggesting that indicating its amorphous structure. After the loading of CNC with
glycerol interacted with κ-carrageenan through intermolecular hy- CNC0, the strong peak of CNC at 22.9° almost disappeared. Similar
drogen bonds and indicating the good miscibility between them. When results were observed by Arrieta, Fortunati, Dominici, Lopez, and
CNC was added in CNC0 (κ-carrageenan containing glycerol), the peaks Kenny (2015). They found that the expected diffraction peaks of cel-
at 3342–3416 cm−1 (OeH stretching vibration), 2982–2932 cm−1 lulose I are not found in poly(lactic acid) films containing low contents
(CeH stretching vibration), 1288–1261 cm−1 (presence of ester sulfate of CNCs (up to 5 wt.%). In another study, Cao, Chen, Chang, Muir, and
group), 943–924 cm−1 (3,6-anhydrogalactose group), and Falk (2008) reported that the XRD patterns of polyurethane films
867–851 cm−1 (galactose-4-sulfate group) sharpened, indicating that containing < 10% CNCs are amorphous. These results indicate that
CNC interacted with CNC0 through intermolecular hydrogen bonds and incorporating small contents of CNCs (up to 3 wt.%) can only slightly
suggested the good miscibility (Xu, Li, Kennedy, Xie, & Huang, 2007) increase the intensity of the characteristic peaks of κ-carrageenan,
between κ-carrageenan and the CNCs. The FTIR spectra of CMC, CNC, whereas adding large contents of CNCs (9 wt.%) can obviously increase
CNC0, κ-carrageenan and CNC9 are shown in Fig. 1. the intensity of the characteristic peaks of κ-carrageenan. The XRD
pattern of CNC9 is nearly the same as that of the CNCs, implying that
3.3. Morphology CNCs were well dispersed in the κ-carrageenan matrix and that the
amorphous structure of κ-carrageenan was not affected by the in-
Fig. 2(a) shows the FESEM image of the CNCs in powder form. After corporation of CNCs.
acid hydrolysis, the microfibrils of the cellulose converted into nano-
fibrils in needle shape. (Fig. 2b). The TEM image of the CNCs confirmed 3.5. TGA
that the lengths and widths of the needles are 200–500 nm and
20–50 nm, respectively (Fig. 2b). The needle shape-like CNCs appeared Vegetable-based food packaging films are widely used in the in-
longer under the SEM, with lengths up to several microns. Some ag- dustry because of their less processing time, high yield, and less pro-
gregations of CNC needles formed larger bundles during freeze drying. duction cost, but temperatures lower their stability in terms of de-
However, these bundles of CNC needles were loosely packed and easily gradation. The thermogravimetric behavior of these films has been
separated by water in preparing the TEM sample for observation reported by many papers (Martelli, Barros, de Moura, Mattoso, & Assis,
(Fig. 2b). The FESEM image of the CN0 and CN9 films showed a 2013; Otoni et al., 2017). Thermal analysis techniques measure thermal
homogeneous and smooth surface structure (Fig. 3(a) and (c)), sug- transitions, chemical reactions and decompositions, viscoelastic prop-
gesting the homogeneous un-agglomerated dispersion of CNCs in the κ- erties, and thermal conductivity as a function of temperature, heating
carrageenan matrix, resulting in good adhesion between the fillers and rate, deformation, and atmosphere. These techniques give insight into
the matrix. The optical microscope (OM) images of the CNCs in dry the specific thermal properties of polymer materials and products. More
form (c) and in wet form (d) at 20× magnification are presented in importantly, they can be used to determine the composition of plastic
Fig. 2. Cross-sectional FESEM micrographs of CN0 and CN9 are shown and rubber compounds and to gain information regarding the condition
in Fig. 3(b) and (d), respectively. The cross-section images confirmed or processing history of specific samples relative to reference samples.
that the smoothness of the κ-carrageenan increased upon the addition of The TGA/DTGA curve was used to determine the thermal stability of
CNC nanoparticles, suggesting the influence of the un-aggregation of CMC, CNC, κ-carrageenan, CNC0, CNC1, CNC3, CNC5, CNC7, and
CNC particles. The images also showed no indication of phase CNC9 by monitoring the weight change that occurred when the samples

185
M. Yadav, F.-C. Chiu Carbohydrate Polymers 211 (2019) 181–194

Fig. 2. (a) FESEM image of CNC (b) TEM image of CNC (c) OM images of CNC in dry form and in wet form (d) at Magnification 20×. All optical microscopy samples
of the fractions were made from dispersions of 5 g dm−3.

were heated in an inert atmosphere. The mass loss (TGA) and derivative steps rather than one. In specific, CNC1 degraded differently at 36
mass loss (DTGA) curves of the samples are given in Figs. 5 and 6. The ℃–100 ℃, 100 ℃–208 ℃, 208 ℃–240 ℃, and 240 ℃–300 ℃, with Tmax
CMCs and CNCs showed similar weight loss in the given TGA/DTGA values of 48 ℃, 180 ℃, 221 ℃, and 245 ℃ identified in the four-step
curve. Two-step degradations were observed during the thermal ana- degradation (DTGA Fig. 6). The three other samples CNC3, CNC5, and
lysis of the samples. The first degradation, which occurred at 30–200 ℃, CNC7 showed multi-step thermal degradation. The first degradation
is attributable to the evaporation of residual water present in the ma- step was found in the range of 32 ℃–124 ℃, which was assigned to
terial. The second degradation at 346 ℃, which occurred at 230–400 ℃, water desorption; the second degradation step, in the range of 124
is characterized by a series of degradation reactions of cellulose, in- ℃–242 ℃, resulted from dehydration of the fiber-forming material
cluding dehydration, decomposition, and depolymerization of the gly- (Islam & Karim, 2010). The last process over 242 ℃ is attributable to
coside units (Vasconcelos et al., 2017). the evaporation of carbonaceous char material (Zhang, Ji, Wang, Tan, &
For κ-carrageenan (Fig. 6), the smaller DTGA peak between 33 ℃ Xia, 2012). Furthermore, the given TGA/DTGA curve shows that the
and 150 ℃ with a weight loss represented the evolution of light vola- thermal stability of CNC9 (190–236 at 227 ℃, 236–282 at 243 ℃) is
tiles together with an additional removal of bound moisture. Between almost same as those of the CNC1, CNC3, CNC5, and CNC7. The ob-
150 ℃ and 222 ℃, a second peak with a loss of 8.4% at 215 ℃ in mass tained TGA/DTGA results in the present work suggest that the thermal
showed the devolatilization of more thermally stable heavy volatiles in stability of the nanocomposite films was retained.
κ-carrageenan. A further increase in temperature to 700 ℃ showed a
third peak at around 237 ℃ (222 ℃–260 ℃) with the highest rate of 3.6. Barrier properties
weight loss (Martins et al., 2012). This last step with a mass loss of
9.38% corresponded to the decomposition of carbonaceous materials. Water sorption is one factor that governs the final product quality,
The degradation of CNC0 occurred in three steps. It degraded differ- considering that the films should be exposed with the outer environ-
ently at 35 ℃–130 ℃, 150 ℃–238 ℃, and 238 ℃–300 ℃, with Tmax ment. Decrease in water sorption facilitates better product quality. The
values of 45 ℃, 220 ℃, and 242 ℃ identified in the three-step de- equilibrium water absorption rate of neat κ-carrageenan films (CNC0)
gradation (DTGA Fig. 6). The degradation of CNC1 occurred in four was 885.94%, which decreased to 561.79% after the incorporation of

186
M. Yadav, F.-C. Chiu Carbohydrate Polymers 211 (2019) 181–194

Fig. 3. (a) FESEM surface image of κ-carrageenan (b) FESEM cross section image of κ-carrageenan (c) FESEM surface image of CNC9 (d) FESEM cross section image of
CNC9 (e) TEM image of CNC9 (f) Digital image of CNC9.

CNCs in the κ-carrageenan film (Table 1). The decrease in water ab- shows the water solubility of the CNC0, CNC1, CNC3, CNC5, CNC7 and
sorption of the κ-carrageenan nanocomposite film is because cellulose is CNC9 nanocomposite films. As CNC content was increased from 0% to
more crystalline than κ-carrageenan and, upon reinforcement, it acts as 9%, the water solubility decreased from 60.94% to 47.97%. This result
a barrier to κ-carrageenan molecules to swell and thereby decrease the was due to the hydrogen bonding generated between hydroxyl groups
water sorption. of κ-carrageenan and CNCs, which led to the formation of 3D networks
A decreased water solubility of hydrophilic materials is also im- of cellulose. The generated 3D networks limit the solubility of the
portant because a certain degree of water resistance is desirable for polymer and lead to further reinforcement, which restricts the move-
most food applications to avoid film disintegration when in contact ment of low-molecular-weight polymers and other compounds to the
with humid food surfaces, such as meat and fresh-cut fruits. Table 1 water (Oleyaei, Zahedi, Ghanbarzadeh, & Moayedi, 2016). The

187
M. Yadav, F.-C. Chiu Carbohydrate Polymers 211 (2019) 181–194

Fig. 4. XRD patterns of κ-carrageenan, CNC, CMC, CNC0, CNC1, CNC3, CNC5, CNC7 and CNC9.

solubility of a κ-carrageenan nanocomposite film with (0.5 wt.%) ZnO is 8.93 × 10−11gm-1 s-1 Pa-1, which reduced by 7.37, 6.25, 5.36, and
higher than that of neat κ-carrageenan film (Saputri et al., 2018). 4.69 × 10−11gm-1 s-1 Pa-1 after the addition of 1, 3, 5, and 7 wt.%
The WVP of food packaging films depends on the film purity, the CNCs, respectively. The reduction of WVP in the κ-carrageenan nano-
hydrophilic–hydrophobic ratio, the ratio between crystalline and composites can be explained by the physical barrier to the passage of
amorphous zones, and the polymeric chain mobility (Souza, Cerqueira, water provided by CNC. That is, the water must flow along the surface
Teixeira, & Vicente, 2010). One of the major purposes of edible films is of CNC; failure to pass through slows down diffusion and thus reduces
to block moisture transfer between the food and the surrounding at- permeability (Azeredo et al., 2009). The barrier properties are en-
mosphere. Therefore, the WVP should be as low as possible. Perme- hanced if the filler is less permeable and has a good dispersion into the
ability is the contribution of diffusivity and solubility of the permeant matrix (Lagaron, Catala, & Gavara, 2004). However, at 9 wt.% loading
through the solid matrix, and this parameter is changed by the structure of CNC, the permeability increases (4.69–9.15 × 10−11 gm−1
of the matrix when using the same permeant. The WVP values of the κ- s−1 Pa−1) drastically because of the increased chances of agglomera-
carrageenan nanocomposites with different CNC loadings are given in tion at high filler loading (Chang, Ruijuan, Zheng, Yu, & Ma, 2010).
the Table 1. The WVP of κ-carrageenan containing glycerol films Garcia et al. (2010) also reported the same agglomeration behavior

188
M. Yadav, F.-C. Chiu Carbohydrate Polymers 211 (2019) 181–194

Fig. 5. TGA curve of CMC, CNC, κ-carrageenan, CNC0, CNC1, CNC3, CNC5, CNC7 and CNC9.

upon loading of more than (1 wt.%) CNC fiber in κ carrageenan-con- CNC9 exhibited the highest and lowest hydrophilicity among all tested
taining glycerol films. Other research papers also showed improvement films, respectively (Table 1). The decreased hydrophilicity of the
in the moisture barrier properties of polymer films with the in- composite film may be attributed to the 3D cellulosic network of CNCs,
corporation of CNCs (Paralikar, Simonsen, & Lombardi, 2008; Svagan, which reduced the number of hydroxyl groups participating in hy-
Hedenqvist, & Berglund, 2009). These improved moisture barrier drogen bonding with the CNCs. The contact angle of CNC9 was lower
properties of CNC-incorporated κ-carrageenan would certainly increase than that of CNC0. The hydrophilicity of the composite films can be
the shelf life of food in packaging applications. controlled by adjusting the amount of CNCs dispersed on the CNC0
Liquid creates a spherical droplet with a small surface area if the surface. Furthermore, the work of adhesion (W12) depends on the
solid liquid interfacial energy is high. The contact angle between a li- contact angle and surface tension of the liquid. In the present study,
quid and solid becomes zero or close to zero that the liquid spreads over W12 is the work that must be done to separate two adjacent phases 1
the solid easily during wetting. In non-wetting, the contact angle is and 2 of a liquid–liquid or liquid–solid phase boundary from one an-
greater than 90° so that the liquid tends form a sphere and run off the other. Conversely, it is the energy that is released during wetting.
surface easily. In the present study, the contact of κ-carrageenan was
Work of adhesion = W12 = (1 + Cos ϴ) Y. (7)
not found due to the substantial spreading of liquid over the surface of
κ-carrageenan/CNC nanocomposites. The contact angles for the CNC0, W12 (Jose, George, Maya, & Thomas, 2015) can be correlated to the
CNC1, CNC3, CNC5, CNC7, and CNC9 films were 23.30°, 23.70°, interaction of the filler with the matrix and to the interaction of the
46.45°, 55.65°, 57.75°, and 71.80°, respectively (Table 1). CNC0 and filler with a liquid that is comparable with the matrix polymer. The

189
M. Yadav, F.-C. Chiu Carbohydrate Polymers 211 (2019) 181–194

Fig. 6. DTGA curve of CMC, CNC, κ-carrageenan, CNC0, CNC1, CNC3, CNC5, CNC7 and CNC9.

Table 1
Physical Properties of κ-carrageenan nanocomposite films.
Samples Code Water absorbency Film solubility WVP (×10−11gm−1 s−1 Pa−1) Water Contact Angle(°) Work of Adhesion

CNC0 885.94 60.94 8.93 23.30 139.69


CNC1 825.00 54.63 7.37 23.70 139.38
CNC3 821.43 53.97 6.25 46.45 122.97
CNC5 778.05 51.22 5.36 55.65 113.88
CNC7 761.54 49.23 4.69 57.75 111.60
CNC9 561.79 47.97 9.15 71.80 95.54

work of adhesion of all selected samples is shown in Table 1. CNC0 and proportional to work of adhesion, implying that hydrophilic films show
CNC9 showed the highest (139.69°) and lowest (95.54°) work of ad- high TS.
hesion, respectively. The work of adhesion decreased with the increase Moisture absorption is a key factor that influences the performance
in CNC concentration. Thus, the effective dispersion of nanofillers into of nanocomposite films for food packaging applications. A natural
the matrix might have decreased the work of adhesion, which is the abundant biopolymer polymer, κ-carrageenan, limits its application due
work required to separate the liquid from the solid surface (Thomas, to its hygroscopic nature. In the present work, κ-carrageenan was re-
Thomas, Abraham, & Bandyopadhyay, 2008). In addition, a previous inforced with CNCs. Due to the better interaction of κ-carrageenan and
study (Shang, Williams, & Soderholm, 1994) found that TS is inversely CNCs by hydrogen bonding, the chances to absorb moisture is

190
M. Yadav, F.-C. Chiu Carbohydrate Polymers 211 (2019) 181–194

Table 2 spectroscopy. A JASCO UV/Vis Spectrometer was used to measure the


Values of thickness, Density, Opacity, tensile strength (TS) and elongation-at- light transmittance of the films in a light wavelength area from 190 nm
break (EB) obtained for κ-carrageenan/CNC films. to 900 nm. The scan speed used in the analysis was 1000 nm/min, and
Samples Thickness Density Opacity Tensile Elongation at three replicates of each material were measured. The absorbance of the
Code (mm) (g/cm3) (Abs600/ Strength Break (%) κ-carrageenan/CNC nanocomposite films is presented in Fig. 8P. To
mm) (MPa) avoid the oxidative rancidity catalyzed by UV light, the packaging film
could able to absorb UV light (Chen et al., 2013). When the films are
CNC0 0.020 1.700 3.781 38.33 ± 3.79 21.50 ± 3.72
CNC1 0.030 1.705 6.422 38.43 ± 5.94 22.93 ± 1.50 not UV responsive, the packaging films must be irradiated with UV light
CNC3 0.040 1.710 10.410 39.83 ± 0.38 23.83 ± 2.71 to destroy microbial charge (Narayanan, Loganathan, Valapa, Thomas,
CNC5 0.025 1.720 11.239 40.07 ± 2.80 24.33 ± 3.00 & Varghese, 2017). In the present study, the absorbance of nano-
CNC7 0.025 1.730 22.224 52.73 ± 0.70 28.27 ± 2.39
composites near the UV region (270–300 nm) increased with increasing
CNC9 0.035 1.750 22.254 39.10 ± 1.04 25.83 ± 2.61
fiber loading, and CNC loading did not significantly affect the UV re-
sponse of the prepared nanocomposites films. Thus, this nanocomposite
film did not cause yellowing or de-coloration when exposed to light
because the nanocomposites were not affected by photo-oxidation re-
action. The UV spectra of CNC0, CNC1, CNC3, CNC5, CNC7, and CNC9
are shown in Fig. 8. The spectra of all films exhibited lower transmit-
tance values at UV light (190–380 nm) than that at visible light
(380–750 nm). Fig. 8Q confirms that the UV light transmittance of the
κ-carrageenan nanocomposite films continuously decreases compared
with that of κ-carrageenan films due to the UV-shielding ability of
CNCs. This decrease in transmittance with the CNC content is indicative
of the uniform dispersion of CNCs in the κ-carrageenan matrix. The
supplied figure shows that the UV light transmittance of CNC0 at
820 nm is 76.28% and decreases to 21.40% as the CNC content of
composites is increased to 9 wt.%. The reduction in light transmittance
in all films still maintained optical transparency, and no visual ag-
gregates were observed (Fig. 8S). The order of optical transparency of
films is as follows:

CNC0 > CNC1 > CNC3 > CNC5 > CNC7 > CNC9

76.28 > 63.52 > 48.14 > 46.06 > 21.42 > 21.40.
Fig. 7. Moisture absorption of CNC0, CNC1, CNC3, CNC5, CNC7 and CNC9
films. The transmittance spectra clearly show that the transmittance re-
duced with increasing CNC content because of light scattering. Similar
decreased. Composites with poor adhesion between the matrix and the results were found in recent papers (Hietala, Mathew, & Oksman,
filler allow moisture to enter the composite material and thereby in- 2013). The blocking effect (Beff) of CNCs for κ-carrageenan nano-
crease the moisture sensitivity (Balakrishnan, Gopi, Sreekala, & composites were calculated using the following formula (Castillo et al.,
Thomas, 2018). The effect of CNC content on the moisture absorption of 2013):
films is shown in Fig. 7. The moisture absorption of the bionano- T carrageenan T carrageenan/CNC
composite films was lower than that of κ-carrageenan film. This result Beff =
Percentage of CNC with respect to the carrageenan
,
may be due to the less availability of free hydroxyl groups of the matrix
(8)
that participated in hydrogen bonding with the CNCs. Thus, the effect
of time on moisture absorption was studied by varying its time from where T κ-carrageenan and T κ-carrageenan/CNC are the transmittance of κ-
12 h to 96 h, and results are presented in Fig. 7. The moisture absorp- carrageenan film and the κ-carrageenan/CNC nanocomposites, respec-
tion increased as the time was prolonged from 12 h to 60 h, and this tively.
increase can be explained due to the greater availability of −OH groups Fig. 8R shows the Beff of the CNCs at 300, 700, and 750 nm (UV-B,
of κ-carrageenan. Beyond 60 h, the decrease in moisture absorption UV-A, and visible regions, respectively).
which might be due to the 3D cellulosic network of CNCs, which re- Furthermore, the transmittance of light decreased in the κ-carra-
stricts the chain mobility and reduces the number of hydroxyl groups geenan/CNC nanocomposites after a certain value of visible light
(Das et al., 2011). The equilibrium moisture uptake value depends on (500 nm) compared with the κ-carrageenan films, confirming the UV
hydrophilicity and morphology (macro-voids, free volume, crystal size, absorption potential of the former. The above findings clearly show that
and crystallinity degree. Recently, Venkatesan et al. (2017) have re- κ-carrageenan/CNCs can be used to protect food materials from UV
ported that the moisture uptake of κ-carrageenan nanocomposite films degradation.
decreases when SiO2 is introduced into the neat polymer matrix.
3.8. Mechanical properties
3.7. Opacity and UV visibility
Good mechanical properties are among the basic requirements for
Transparent food packaging secures the visual evidence of the food food packaging films, considering that poor flexibility or strength may
content, which is an essential factor that affects the consumers’ sensi- lead to premature failure or cracking during production, handling,
bility of food quality. Transparent food packaging clubbed with light storage, or use (Sothornvit & Rodsamran, 2008). The most investigated
labelling appeals to consumers because it provides a clear view of the mechanical properties of packaging films are TS and EB. These prop-
food content and its condition. Opacity is an important factor that de- erties support the correlation of the mechanical properties of films to
termines the quality of nanocomposite films. The opacity of all required their compositions and chemical structures. The fabricated bionano-
samples is presented in Table 2. The transparency of the CNC0, CNC1, composite films were sectioned into rectangular shapes with a size of
CNC3, CNC5, CNC7, and CNC9 films was studied using UV/Visible light 10 mm × 60 mm × 0.02 mm for tensile tests. The tensile tests of the

191
M. Yadav, F.-C. Chiu Carbohydrate Polymers 211 (2019) 181–194

Fig. 8. UV spectra of CNC0, CNC1, CNC3, CNC5, CNC7 and CNC9 (P) Absorbance curve (Q) Percentage transmittance curve (R) Blocking effect at different
wavelengths and (S) Visual appearance.

CNC0, CNC1, CNC3, CNC5, CNC7, and CNC9 films (in accordance with MPa (Venkatesan et al., 2017). The TS values of all selected films are
ASTM D638) were conducted at a crosshead speed of 10 mm/min by comparable to those of plastic-based films, such as films based on LDPE
using a Gotech AI-3000 system. The TS and EB were taken to be the (8–10 MPa), HDPE (19–31 MPa), EVOH (6–19 MPa), PCL (4 MPa), PS
average value of the five tests for each composition. Tensile properties (31–49 MPa), PLA (45 MPa), PVC (42–55 MPa), PP (27–98 MPa), and
(TS and EB) of the CNC0, CNC1, CNC3, CNC5, CNC7, and CNC9 bio- PET (157–177 MPa) (Otoni et al., 2017). The κ-carrageenan/SiO2 na-
nanocomposites films are shown in Table 2. Loading with a small nocomposite films showed higher TS (53.9 MPa) than the κ-carrageenan
amount of CNC significantly improved the tensile properties. The TS of neat polymer (46.8 MPa) (Venkatesan et al., 2017). The fabricated κ-
CNC0 (κ-carrageenan containing glycerol) (38.33 ± 3.79 MPa) is evi- carrageenan nanocomposite film loaded with 0.5 wt.% ZnO showed
dently lower than those of CNC1 (38.43 ± 5.94 MPa), CNC3 highest TS for food packaging applications (Saputri et al., 2018).
(39.83 ± 0.38 MPa), CNC5 (40.07 ± 2.80 MPa), CNC7 Similar to the TS results, the EB values of CNC0, CNC1, CNC3,
(52.73 ± 0.70 MPa), and CNC9 (39.10 ± 1.04 MPa). These results CNC5, CNC7, and CNC9 also increased, with 21.50 ± 3.72,
can be attributed to the strong hydrogen bonding between the CNCs 22.93 ± 1.50, 23.83 ± 2.71, 24.33 ± 3.00, 28.27 ± 2.39, and
and the κ-carrageenan matrix phase due to their chemical similarity 25.83 ± 2.61%, respectively (Table 2). In specific, the EB values of the
(Huq et al., 2012). Furthermore, the tensile strengths of the films de- CNC1, CNC3, CNC5, CNC7, and CNC9 films were 6.67%, 10.85%,
creased from 52.73 ± 0.70 MPa to 39.10 ± 1.04 MPa when the 13.18%, 31.47%, and 20.16% higher than that of the κ-carrageenan
loading amount of the CNC was increased from 7 wt.% to 9 wt.%. This film, respectively. Furthermore, the EB of the films decreased from
result is caused by the agglomeration of CNCs in the κ-carrageenan 28.27 ± 2.39% to 25.83 ± 2.61% when the loading amount of the
matrix. The aggregation of CNCs may occur while the fiber is in na- CNCs was increased from 7 wt.% to 9 wt.%. The decrease in EB of the
nophase, and the tendency to form large macroscopic fibers increases at nanocomposite films can be explained by the restrictions to the se-
high filler loading. These phenomena result in inter-fiber interaction quential motion at the filler–polymer interface (Wang et al., 2005). The
due to hydrogen bonding-induced self-association (Dorigato, Dzenis, & EB of the prepared nanocomposites is comparable to those of films
Pegoretti, 2013). Similar results were previously reported (Wu, Zhang, based on LDPE (300%–900%), HDPE (20%–50%), PCL (800%–1000%)
Rong, & Friedrich, 2005). The TS of the κ-carrageenan control film (Bastarrachea, Dhawan, & Sablani, 2011), PS (2%–3%), PVC
(38.33.MPa) obtained in this study was similar to those of films in (20%–180%), PP (200%–1000%), PET (70%), and PVDC (10%–40%).
previous reports, such as 23.4 MPa (Zarina & Ahmad, 2015), 28.1 MPa However, the TS and EB of the bionanocomposites with high CNCs
(Wahab & A B D Razak, 2016), 30.68 MPa (Saputri et al., 2018), content were still higher than those of the neat matrix. Thus, the CNCs
39.44 MPa (Rhim, 2012), 42.5 MPa (Larotonda et al., 2016), and 46.8. acted as good reinforcing agents in the κ-carrageenan films.

192
M. Yadav, F.-C. Chiu Carbohydrate Polymers 211 (2019) 181–194

4. Conclusion polyvinyl alcohol biocomposite films. Composite Part B, 42, 376–381.


Dorigato, A., Dzenis, Y., & Pegoretti, A. (2013). Filler aggregation as a reinforcement
mechanism in polymer nanocomposites. Mechanics of Materials, 61, 79–90.
The needle-shaped Cellulose nanocrystals (CNCs) were synthesized Dou, Y., Zhang, B., He, M., Yin, G., & Cui, Y. (2016). The structure, tensile properties and
from ultra-fine microcellulose by using acid hydrolysis. The CNCs re- water resistance of hydrolyzed feather keratin-based bioplastics. Chinese Journal of
inforced κ-carrageenan transparent bionanocomposite films were pre- Chemical Engineering, 24, 415–420.
Fan, L., Peng, K., Li, M., Wang, L., & Wang, T. (2013). Preparation and properties of
pared by using solution casting method. FTIR analysis revealed the carboxymethyl κ-carrageenan/alginate blend fibers. Journal of Biomaterials Science,
formation of hydrogen bond at the interface of κ-carrageenan and Polymer Edition, 24, 1099–1111.
CNCs. The XRD and FESEM studies showed the homogeneous disper- Garcia, M. D. S., Hilliou, L., & Lagaron, J. M. (2010). Morphology and water barrier
properties of nanobiocomposites of κ/i-hybrid carrageenan and cellulose nano-
sion of the CNCs in the κ-carrageenan matrix, without any aggregation. whiskers. Journal of Agricultural and Food Chemistry, 58, 12847–12857.
The effects of CNCs concentration on the physical and mechanical Hietala, M., Mathew, A. P., & Oksman, K. (2013). Bionanocomposites of thermoplastic
properties (density, moisture content, opacity, color, water solubility, starch and cellulose nanofibers manufactured using twin-screw extrusion. European
Polymer Journal, 49, 950–956.
WVP, TS, and EB) of κ-carrageenan film were systematically studied.
Huq, T., Salmieri, S., Khan, A., Khan, R. A., Le Tien, C., Riedl, B., Fraschini, C., Bouchard,
TGA data showed that the thermal stability of all selected samples was J., Uribe-Calderon, J., Kamal, M. R., & Lacroix, M. (2012). Nanocrystalline cellulose
retained. The WVP decreased from 8.93 to 4.69 × 10−11gm-1 s-1 Pa-1 (NCC) reinforced alginate based biodegradable nanocomposite film. Carbohydrate
after 7 wt.% CNC loading. With the incorporation of 7 wt.% CNC, the TS Polymers, 90, 1757–1763.
Islam, M. S., & Karim, M. R. (2010). Fabrication and characterization of poly(vinyl al-
and EB of the films increased from 38.33 ± 3.79 MPa to cohol)/alginate blend nanofibers by electrospinning method. Colloids and Surfaces A:
52.73 ± 0.70 MPa and from 21.50 ± 3.72% to 28.27 ± 2.39%, re- Physicochemical and Engineering Aspects, 366, 135–140.
spectively. Compared with the κ-carrageenan film, the κ-carrageenan/ Jose, T., George, S. C., Maya, M. G., & Thomas, S. (2015). Functionalized MWCNT and
PVA nanocomposite membranes for dielectric and pervaporation applications.
CNC films had higher water barrier properties and mechanical prop- Journal of Chemical Engineering and Process Technology, 6, 1000233.
erties. Therefore, the obtained κ-carrageenan/CNC bionanocomposite Kang, X., Kuga, S., Wang, C., Zhao, Y., Wu, M., & Huang, Y. (2018). Green preparation of
films have potential applications in food packaging. cellulose nanocrystal and its application. ACS Sustainable chemistry and engineering, 6,
2954–2960.
Karbowiak, T., Debeaufort, F., Champion, D., & Voilley, A. (2006). Wetting properties at
Acknowledgments the surface of iota carrageenan-based edible films. Journal of Colloid and Interface
Science, 294, 400–410.
Khalil, H. P. S. A., Bhat, A. H., & Yusra, A. F. I. (2012). Green composites from sustainable
Mithilesh Yadav thanks to Chang Gung Memorial Hospital cellulose nanofibrils: A review. Carbohydrate Polymers, 87, 963–979.
(BMRP392) for providing the Post-doctoral fellowship. The financial Khan, R. A., Salmieri, S., Dussault, D., Uribe-Calderon, J., Kamal, M. R., Safrany, A., et al.
support from Ministry of Science and Technology (Taiwan) under (2010). Production and properties of nanocellulose-reinforced methylcellulose-based
biodegradable films. Journal of Agricultural and Food Chemistry, 58, 7878.
contract MOST 107-2811-E-182-503 is appreciated as well.
Khilil, F., Borges, J., Almeida, P. L., Boukherroub, R., & Omrani, A. D. (2018). Extraction
of cellulose nanocrystals with structure I and II and their applications for reduction of
Appendix A. Supplementary data graphene oxide and nanocomposite elaboration. Waste and Biomass Valorization.
https://doi.org/10.1007/s12649-018-0202-4.
Korolovych, V. F., Cherpak, V., Nepal, D., Ng, A., Shaikh, N. R., Grant, A., et al. (2018).
Supplementary material related to this article can be found, in the Cellulose nanocrystals with different morphologies and chiral properties. Polymer,
online version, at doi:https://doi.org/10.1016/j.carbpol.2019.01.114. 145, 334–347.
Lafargue, D., Lourdin, D., & Doublier, J. L. (2007). Film-forming properties of a modified
starch/kappa-carrageenan mixture in relation to its rheological behavior.
References Carbohydrate Polymers, 70, 101–111.
Lagaron, J. M., Catala, R., & Gavara, R. J. (2004). Structural characteristics defining high
Almasi, H., Ghanbarzadeh, B., & Entezami, A. A. (2010). Physicochemical properties of barrier properties in polymeric materials. Journal of Materials Science & Technology,
starch/CMC/nanoclay biodegradable films. International Journal of Biological 20, 1.
Macromolecule, 46, 1–5. Larotonda, F. D. S., Torres, M. D., Goncalves, M. P., Sereno, A. M., & Hilliou, L. (2016).
Arrieta, M. P., Fortunati, E., Dominici, F., Lopez, J., & Kenny, J. M. (2015). Hybrid carrageenan-based formulations for edible film preparation: Benchmarking
Bionanocomposite films based on plasticized PLA–PHB/cellulose nanocrystal blends. with kappa carrageenan. Journal of Applied Polymer Science, 42263, 1–10.
Carbohydrate Polymers, 121, 265–275. Li, Q., & Renneckar, S. (2011). Supramolecular structure characterization of molecularly
ASTM-E96/E96-05 (2005). Standard test methods for Water vapor transmission of materials. thin cellulose I nanoparticles. Biomacromolecules, 12, 650–659.
West Conshohocken, USA: ASTM International. Lu, P., & Hsieh, Y. L. (2010). Preparation and properties of cellulose nanocrystals: Rods,
Azeredo, H. M. C., Mattoso, L. H. C., Wood, D., Williams, T. G., Avena-Bustillos, R. J., & spheres, and network. Carbohydrate Polymers, 82, 329–336.
McHugh, T. H. (2009). Journal of Food Science and Technology, 74, 31. Majoinen, J., Kontturi, E., Ikkala, O., & Gray, D. G. (2012). SEM imaging of chiral nematic
Azizi, S., & Mohamad, R. (2018). Mechanical and barrier properties of kappa-carra- films cast from cellulose nanocrystals suspensions. Cellulose, 19, 1599–1605.
geenan/cellulose nanocrystals bio- nanocomposite films. IOP Conferences Series: Martelli, M. R., Barros, T. T., de Moura, M. R., Mattoso, L. H. C., & Assis, O. B. G. (2013).
Materials Science and Engineering, 368, 012013. Effect of chitosan nanoparticles and pectin content on mechanical properties and
Azizi Samir, A., Alloin, F., & Dufresne, A. (2005). Review of recent research into cellulosic water vapor permeability of banana puree films. Journal of Food Science, 78,
whiskers, their properties and their application in nanocomposite Field. N98–N104.
Biomacromolecules, 6, 612. Martins, J. T., Cerqueira, M. A., Bourbon, A. I., Pinheiro, A. C., Souza, B. W. S., & Vicente,
Balakrishnan, P., Gopi, S., Sreekala, M. S., & Thomas, S. (2018). UV resistant transparent A. A. (2012). Synergistic effects between κ-carrageenan and locust bean gum on
bionanocomposite films based on potato starch/cellulose for sustainable packaging. physicochemical properties of edible films made thereof. Food Hydrocolloids, 29,
Starch/Stärke, 70, 1700139. 280–289.
Bastarrachea, L., Dhawan, S., & Sablani, S. S. (2011). Engineering properties of poly- Narayanan, M., Loganathan, S., Valapa, R. B., Thomas, S., & Varghese, T. O. (2017). UV
meric-based antimicrobial films for food packaging: A review. Food Engineering protective poly (lactic acid)/rosin films for sustainable packaging. International
Reviews, 3, 79–93. Journal of Biological Macromolecule, 99, 37–45.
Cao, X., Chen, Y., Chang, P. R., Muir, A. D., & Falk, G. (2008). Starch-based nano- Nikonenko, N. A., Buslov, D. K., Sushko, N. I., & Zhbankov, R. G. (2005). Spectroscopic
composites reinforced with flax cellulose nanocrystals. Express Polymer Letter, 2, manifestation of stretching vibrations of glycosidic linkage in polysaccharides.
502–510. Journal of Molecular Structure, 752, 20–24.
Castillo, L., Lopez, O., Lopez, C., Zaritzky, N., García, M. A., Barbosa, S., & Villar, M. Noshirvani, N., Ghanbarzadeh, B., Fasihi, H., & Almasi, H. (2016). Starch–PVA nano-
(2013). Thermoplastic starch films reinforced with talc nanoparticles. Carbohydrate composite film incorporated with cellulose nanocrystals and MMT: A comparative
Polymers, 95, 664–674. study. International Journal of Food Engineering, 12, 37–48.
Cerqueira, M. A., Souza, B. W. S., Simões, J., Teixeira, J. A., Domingues, M. R. R. M., Oleyaei, S. A., Zahedi, Y., Ghanbarzadeh, B., & Moayedi, A. A. (2016). Modification of
Coimbra, M. A., et al. (2011). Structural and thermal characterization of galacto- physicochemical and thermal properties of starch films by incorporation of TiO2
mannans from non-conventional sources. Carbohydrate Polymers, 83, 179–185. nanoparticles. International Journal of Biological Macromolecule, 89, 256–264.
Chang, P. R., Ruijuan, J., Zheng, P., Yu, J., & Ma, X. (2010). Preparation and properties of Orsuwan, A., Shankar, S., Wang, L. F., Sothornvit, R., & Rhim, J. W. (2016). Preparation
glycerol plasticized starch (GPS) cellulose nanoparticle (CN) composites. of antimicrobial agar/banana powder blend films reinforced with silver nano-
Carbohydrate Polymer, 79, 301–305. particles. Food Hydrocolloids, 60, 476–485.
Chen, H. M., Zhang, W. B., Du, X. C., Yang, J. H., Zhang, N., Huang, T., et al. (2013). Otoni, C. G., Bustillos, R. J. A., Azeredo, H. M. C., Lorevice, M. V., Moura, M. R., Mattoso,
Crystallization kinetics and melting behaviors of poly (l-lactide)/graphene oxides L. H. C., et al. (2017). Recent advances on edible films based on fruits and vegetables:
composites. Thermochimica Acta, 566, 57–70. A review. Food Science and Food Safety, 16, 1151.
Das, K., Ray, D., Bandyopadhyay, N. R., Sahoo, D., Mohanty, A. K., & Misra, M. (2011). Paralikar, S. A., Simonsen, J., & Lombardi, J. (2008). Polyvinyl alcohol/cellulose nano-
Physico-mechanical properties of the jute micro/nanofibril reinforced starch/ crystal barrier membranes. Journal of Membrane Science, 320, 248–258.

193
M. Yadav, F.-C. Chiu Carbohydrate Polymers 211 (2019) 181–194

Park, H. (1996). Gas and mechanical barrier properties of carrageenan-based biopolymer (2016). Incorporation of ZnO and their composite nanostructured material into a
films. Food Science and Industry, 29, 47–53. cotton fabric platform for wearable device applications. Carbohydrate Polymers, 157,
Park, S. Y., Lee, B. I., Jung, S. T., & Park, H. J. (2001). Biopolymer composite films based 1801–1808.
on carrageenan and chitosan. Materials Research Bulletin, 36, 511–519. Veluswamya, P., Suhasini, S., Archana, J., Navaneethan, M., Majumdar, A., Hayakawa,
Rescignano, N., Fortunati, E., Montesano, S., Emiliani, C., Kenny, J. M., Martino, S., et al. Y., et al. (2017). Fabrication of hierarchical ZnO nanostructures on cotton fabric for
(2014). PVA bio-nanocomposites: A new take-off using cellulose nanocrystals and wearable device applications. Applied Surface Science, 418, 352–361.
PLGA nanoparticles. Carbohydrate Polymers, 99, 47–58. Venkatesan, R., Rajeswari, N., & Thiyagu, T. T. (2017). Preparation, characterization and
Rhim, J. W. (2012). Physical-mechanical properties of Agar/κ-carrageenan blend film and mechanical properties of κ-carrageenan/SiO2 nanocomposite films for antimicrobial
derived clay nanocomposite film. Journal of Food Science, 77, 66–73. food packaging. Bulletin of Materials Science, 40, 609–614.
Roy, S., Shankar, S., & Rhim, J. W. (2019). Melanin-mediated synthesis of silver nano- Wahab, I. F., & A B D Razak, S. I. (2016). Bionanocomposite film of κ-carrageenan/na-
particle and its use for the preparation of carrageenan-based antibacterial films. Food notube clay: Growth of hydroxyl apatite and model drug release. Digest Journal of
Hydrocolloids, 88, 237–246. Nanomaterials and Biostructures, 11, 963–972.
Saputri, A. E., Praseptiangga, D., Rochima, E., Panatarani, C., & Joni, I. M. (2018). Wang, H., & Roman, M. (2011). Formation and properties of chitosan cellulose nano-
Mechanical and solubility properties of bio-nanocomposite film of semi refined kappa crystal polyelectrolyte macro ion complexes for drug delivery applications.
carrageenan/ZnO nanoparticles. AIP Conference Proceedings 1927, 030040. https:// Biomacromolecules, 12, 1585–1593.
doi.org/10.1063/1.5021233. Wang, S., Song, C., Chen, G., Gu, T., Liu, J., Zhang, B., et al. (2005). Characteristics and
Segal, L., Creely, J. J., Martin, A. E., & Conrad, C. M. (1959). An empirical method for biodegradation properties of poly (3-hydroxybutyrate-co-3-hydroxyvalerate)/orga-
estimating the degree of crystallinity of native cellulose using X-ray diffractometer. nophilic montmorillonite (PHBV/OMMT) nanocomposite. Polymer Degradation
Textile Research Journal, 29, 786–794. Stability, 87, 69–76.
Sen, M., & Erboz, E. N. (2010). Determination of critical gelation conditions of κ-carra- Wu, C. L., Zhang, M. Q., Rong, M. Z., & Friedrich, K. (2005). Silica nanoparticles filled
geenan by viscosimetric and FT-IR analyses. Food Research Inter-national, 43, polypropylene: Effects of particle surface treatment, matrix ductility and particle
1361–1364. species on mechanical performance of the composites. Composite Science and
Shang, S. W., Williams, J. W., & Soderholm, K. J. M. (1994). How the work of adhesion Technology, 65, 635.
affects the mechanical properties of silica-filled polymer composites. Journal of Xie, H., Xiang, C., Lia, Y., Wang, L., Zhang, Y., Song, Z., et al. (2019). Fabrication of
Materials Science, 29, 2406–2416. ovalbumin/κ-carrageenan complex nanoparticles as a novel carrier for curcumin
Sorrentino, A., Gorrasi, G., & Vittoria, V. (2007). Potential perspectives of bio-nano- delivery. Food Hydrocolloids, 89, 111–121.
composites for food packaging applications. Trends in Food Science & Technology, 18, Xu, X., Li, B., Kennedy, J. F., Xie, B. J., & Huang, M. (2007). Characterization of konjac
84–95. glucomannane gellan gum blend films and their suitability for release of nisin in-
Sothornvit, R., & Rodsamran, P. (2008). Effect of a mango film on quality of whole and corporated therein. Carbohydrate Polymers, 70, 192–197.
minimally processed mangoes. Postharvest Biology and Technology, 47, 407–415. Yadav, M., Seongcheol, M., Hyun, J., & Kim, J. (2015). Synthesis and characterization of
Souza, B., Cerqueira, M., Teixeira, J., & Vicente, A. (2010). The use of electric fields for iron oxide/cellulose nanocomposite film. International Journal of Biological
edible coatings and films development and production: A review. Food Engineering Macromolecules, 74, 142–149.
Reviews, 2, 244–255. Youssef, A. M., & El- Sayed, S. M. (2018). Bionanocomposites materials for food packa-
Svagan, A. J., Hedenqvist, M. S., & Berglund, L. (2009). Reduced water vapor sorption in ging applications: Concepts and future outlook. Carbohydrate Polymers, 193, 19–27.
cellulose nanocomposites with starch matrix. Composite Science and Technology, 69, Youssef, A. M., EL-Sayed, S., Salama, H., EL-Sayed, H., & Dufresne, A. (2015). Evaluation
500–506. of bionanocomposites as packaging material on properties of soft white cheese during
Thomas, S. P., Thomas, S., Abraham, R., & Bandyopadhyay, S. (2008). Polystyrene/cal- storage period. Carbohydrate Polymers, 132, 274–285.
cium phosphate nanocomposites: Contact angle studies based on water and methy- Youssef, A., EL-Sayed, S., EL-Sayed, H., Salama, H., & Dufresne, A. (2016). Enhancement
lene iodide. Express Polymer Letters, 2, 528–538. of Egyptian soft white cheese shelf life using a novel chitosan/carboxymethyl cellu-
Tingaut, P., Zimmermann, T., & Sèbe, G. (2012). Cellulose nanocrystals and micro- lose/zinc oxide bionanocomposite film. Carbohydrate Polymers, 151, 9–19.
fibrillated cellulose as building blocks for the design of hierarchical functional ma- Youssef, A. M., El-Sayed, S. M., El-Sayed, H. S., Salama, H. H., Assem, F. M., & Abd El-
terials. Journal of Materials Chemistry, 22, 20105–20111. Salam, M. H. (2018). Novel bionanocomposite materials used for packaging skimmed
Turquois, T., Acquistapace, S., Vera, F. A., & Welti, D. H. (1996). Composition of carra- milk acid coagulated cheese (Karish). International Journal of Biological
geenan blends inferred from 13C-NMR and infrared spectroscopic analysis. Macromolecules, 115, 1002–1011.
Carbohydrate Polymers, 31, 269–278. Zarina, S., & Ahmad, I. (2015). Biodegradable composite films based on κ-carrageenan
Vasconcelos, N. F., Feitosa, J. P. A., Gama, F. M. P., Morais, J. P. S., Andrade, F. K., Filho, reinforced by cellulose nanocrystal from kenaf fibers. BioResources, 10, 256–271.
M. S. M. S., et al. (2017). Bacterial cellulose nanocrystals produced under different Zhang, J., Ji, Q., Wang, F., Tan, L., & Xia, Y. (2012). Effects of divalent metal ions on the
hydrolysis conditions: Properties and morphological features. Carbohydrate Polymers, flame retardancy and pyrolysis products of alginate fibers. Polymer Degradation and
155, 425–431. Stability, 97, 1034–1040.
Veluswamya, P., Suhasini, S., Khana, F., Ghoshd, A., Abhijit, M., Hayakawa, Y., et al.

194

You might also like