ArticlePDF Available

Antimicrobial Edible Film Prepared from Bacterial Cellulose Nanofibers/Starch/Chitosan for a Food Packaging Alternative

Wiley
International Journal of Polymer Science
Authors:

Abstract and Figures

As a contribution to the growing demand for environmentally friendly food packaging films, this work produced and characterized a biocomposite of disintegrated bacterial cellulose (BC) nanofibers and tapioca starch/chitosan-based films. Ultrasonication dispersed all fillers throughout the film homogeneously. The highest fraction of dried BC nanofibers (0.136 g) in the film resulted in the maximum tensile strength of 4.7 MPa. 0.136 g BC nanofiber addition to the tapioca starch/chitosan matrix increased the thermal resistance (the temperature of maximum decomposition rate from 307 to 317°C), moisture resistance (after 8 h) by 8.9%, and water vapor barrier (24 h) by 27%. All chitosan-based films displayed antibacterial activity. This characterization suggests that this environmentally friendly edible biocomposite film is a potential candidate for applications in food packaging.
This content is subject to copyright. Terms and conditions apply.
Research Article
Antimicrobial Edible Film Prepared from Bacterial Cellulose
Nanofibers/Starch/Chitosan for a Food Packaging Alternative
Hairul Abral ,
1
Angga Bahri Pratama ,
2
Dian Handayani ,
3
Melbi Mahardika ,
4
Ibtisamatul Aminah ,
3
Neny Sandrawati ,
3
Eni Sugiarti,
5
Ahmad Novi Muslimin ,
5
S. M. Sapuan ,
6
and R. A. Ilyas
7
1
Department of Mechanical Engineering, Andalas University, 25163 Padang Sumatera Barat, Indonesia
2
Program Studi Teknik Mesin, Universitas Dharma Andalas, 25000 Padang Sumatera Barat, Indonesia
3
Laboratory of Sumatran Biota/Faculty of Pharmacy, Andalas University, 25163 Padang, Sumatera Barat, Indonesia
4
Department of Biosystems Engineering, Institut Teknologi Sumatera, 35365 South Lampung, Indonesia
5
Laboratory of High-Temperature Coating, Research Center for Physics Indonesian Institute of Sciences (LIPI), Serpong, Indonesia
6
Department of Mechanical and Manufacturing Engineering, Faculty of Engineering, Universiti Putra Malaysia,
43400 UPM Serdang, Selangor, Malaysia
7
School of Chemical and Energy Engineering, Faculty of Engineering, Universiti Teknologi Malaysia, 81310 Skudai, Johor Bahru,
Johor, Malaysia
Correspondence should be addressed to Hairul Abral; habral@yahoo.com
Received 2 January 2021; Revised 9 March 2021; Accepted 15 March 2021; Published 1 April 2021
Academic Editor: Victor H. Perez
Copyright © 2021 Hairul Abral et al. This is an open access article distributed under the Creative Commons Attribution License,
which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
As a contribution to the growing demand for environmentally friendly food packaging lms, this work produced and
characterized a biocomposite of disintegrated bacterial cellulose (BC) nanobers and tapioca starch/chitosan-based lms.
Ultrasonication dispersed all llers throughout the lm homogeneously. The highest fraction of dried BC nanobers
(0.136 g) in the lm resulted in the maximum tensile strength of 4.7 MPa. 0.136 g BC nanober addition to the tapioca
starch/chitosan matrix increased the thermal resistance (the temperature of maximum decomposition rate from 307 to
317
°
C), moisture resistance (after 8 h) by 8.9%, and water vapor barrier (24 h) by 27%. All chitosan-based lms displayed
antibacterial activity. This characterization suggests that this environmentally friendly edible biocomposite lm is a
potential candidate for applications in food packaging.
1. Introduction
The demand for aordable environmentally friendly plastic
substitutes made from renewable sources continues to
increase resulting in a growing interest in the research
community in plant-based replacements for nondegradable
plastics [13]. For the food packaging industry, promising
cheap and abundant eco-friendly edible lm-forming mate-
rials include starches, pure bacterial cellulose nanobers,
and chitosan [4]. However, lms made of starch alone have
low mechanical and thermal properties, high moisture
absorption, and poor antimicrobial resistance [5, 6]. Many
attempts have been conducted previously to reduce these
weaknesses by adding environmentally friendly llers to the
starch lm [710]. Of these edible llers, cellulose ber and
chitosan have signicant potential, with one being one of
the most abundant in nature[10, 11]. Cellulose ber has good
mechanical properties and exibility but no antimicrobial
activity [12]. The tensile and thermal properties of the starch
lm were improved after reinforcement with randomly ori-
ented cellulose bers from kenaf [13], water hyacinth [14],
and softwood [15]. However, these short cellulose nanober
preparation methods tend to use potentially harmful chemi-
cals and result in a high residue of hemicellulose, lignin, or
other impurities. Nanober from bacterial cellulose pellicle
has none of these disadvantages because it consists of large
Hindawi
International Journal of Polymer Science
Volume 2021, Article ID 6641284, 11 pages
https://doi.org/10.1155/2021/6641284
amounts of pure cellulose bers with nanosized diameters
[16]. Short BC nanobers can be prepared from this pellicle
using a high-shear homogenizer with a rotating blade to
disintegrate the long bers into nanosized lengths [17].
Recently, previous work has reported the interesting results
that the microbial activity of the disintegrated BC/chitosan
lm was less than that of the BC sheet/chitosan one [1].
Chitosan has several advantages including antimicrobial
activity, controlled release of food ingredients and drugs, rel-
atively low cost, and widespread availability from a stable
renewable source [5, 18]. Numerous studies have been per-
formed on the development of edible biocomposite lms
made from chitosan, cellulose, and starch [19, 20]. Of course,
as a food packaging material, this polysaccharide-based edi-
ble lm should protect foods against deterioration due to
microorganisms, moisture, dust, odors, and mechanical
forces [21, 22]. There have been many previous studies
reporting on the improved properties of the biocomposite
lm. However, characterizations of chitosan and disinte-
grated BC nanober content on tapioca starch-based bio-
compositesproperties have yet to be reported. This study
understands the eect of the addition of both chitosan and
disintegrated nanober from BC pellicle on the properties
of this edible biocomposite lm more completely. All sam-
ples were characterized by eld emission scanning electron
microscopy (FESEM), X-ray diraction (XRD), Fourier
transform infrared (FTIR), and thermogravimetric analysis
(TGA). Opacity, moisture absorption (MA), water vapor per-
meability (WVP), and tensile properties were also measured.
2. Materials and Experiment
2.1. Materials. Local (Padang, Indonesia) commercially avail-
able tapioca starch (Cap Pak Tani brand) was washed once
with distilled water to obtain pure tapioca starch. The pure
wet starch was dried using a drying oven (Universal Oven
Memmert UN-55) for 20 h at 50
°
C. The dried starch was l-
tered through a 63 μm mesh cloth. The chemical composi-
tion of dry starch granules was analyzed according to a
previous amylopectin method [23]. The pure dried granular
starch for this present work contained 14.5% of amylose
and 85.5% of glycerol (Brataco brand) purchased from
Brataco (Padang, Indonesia) and chitosan (degree of deacety-
lation: 94%) from CV. Chi Multiguna (Indramayu, Indone-
sia). Acetic acid (CH
3
COOH, 5%) was used as the solvent
for chitosan.
2.2. Preparation of Single BC Nanober and Films. Single BC
nanober was isolated and characterized as in our previous
work [24]. Briey, a wet pellicle with a dimension
(248 × 151 × 22 mm) was purchased from a local small-
scale industry in Padang, Indonesia. This pellicle, which is a
common addition to drinks or desserts in the form of nata
de coco, is the result of a week-long fermentation of coconut
water, glucose, and acetic acid with Acetobacter xylinum in a
static closed container. Integration of the pellicle was carried
out with a homogenizer and ultrasonication probe. The crys-
tallinity index of the disintegrated BC nanober was about
71%. Figure 1 displays the steps of sample preparation from
raw BC until biocomposite. The lm sample was prepared
as follows:
Starch lm. About 10 g puried tapioca granules, 100 mL
distilled water, and 2 mL glycerol were mixed in a glass bea-
ker (250 mL) using a hot plate stirrer (Daihan Scientic
MSH-200) at 500 rpm and 65
°
C until completely gelatinized.
The gel suspension was sonicated using an ultrasonic cell
crusher (SJIA-1200W) at 600 W for 1 min, then poured onto
a petri dish (d= 145 mm) and dried in a drying oven (Mem-
mert Germany, Model 55 UN) at 50
°
C for 20 h.
Chitosan lm. Simultaneously, a mixture of 2.5 g chitosan
and 100 mL acetic acid was prepared in a glass beaker
(250 mL) and heated using a hot plate stirrer at 80
°
C for 2
hours until gelatinization. The chitosan gel was ltered using
74 μm cheesecloth. The gel was poured onto a petri dish
which was dried using the drying oven at 50
°
C for 20 h.
Starch/chitosan lm. The two gels, starch and chitosan,
were blended in the ratio of 80 : 20 (70 g total weight) using
an ultrasonic cell crusher at 600 W for 1 min keeping the
temperature below 65
°
C. The gel was poured onto a petri dish
for drying as described in the starch sample preparation.
Biocomposite lm. The blended starch/chitosan gels were
mixed with the appropriate BC suspension (10, 15, or 20 mL).
Each gel suspension was sonicated at 600 W for 1 min, then
cast onto a petri dish and dried in the drying oven at 50
°
C
for 20 h.
Abbreviations used for the studied samples with their
compositions are shown in Table 1.
2.3. Characterization
2.3.1. FESEM Morphology of the Fracture Surface. The tensile
samples fracture surface morphology after the tensile test
was observed using JIB-4610F FESEM from JEOL (Tokyo,
Japan). At about 25 mm from its fracture surface, the tensile
specimen was cut using a steel scissor (in perpendicular ten-
sile direction) and placed on a specimen holder. All samples
were coated with gold (Au). An accelerati ng voltage of 10 kV
with 8 mA was set up for testing.
2.3.2. X-Ray Diraction. The XRD pattern was recorded
using an XPert PRO PANalytical instrument (Philips Ana-
lytical, Netherlands) with CuKαradiation (λ=0:154)at
40 kV and 30 mA. The scanning range was 5
°
to 50
°
. The crys-
tallinity index (CI) of the biocomposites was calculated using
CI = I002 Iam
ðÞ
I002

× 100, ð1Þ
where I002 and Iam are the peak intensities of crystalline and
amorphous regions, respectively [25].
2.3.3. Opacity Measurement. The opacity of the lm was mea-
sured using a UV-Vis spectrophotometer (Shimadzu UV
1800, Japan) in the range 400-800 nm according to ASTM
D 1003-00 (Standard Test Method for Haze and Luminous
Transmittance of Transparent Plastics). Films of 0.38 mm
thickness were cut into 10 mm × 25 mm rectangles. The
opacity measurement was repeated 3 times.
2 International Journal of Polymer Science
2.3.4. Fourier Transform Infrared. FTIR spectra of lms were
recorded using a PerkinElmer FTIR spectrometer (Frontier
Instrument, USA), equipped with deuterated triglycine sul-
fate, DTGS, detector, and extended range KBr beam splitter.
This spectrometer was used in the frequency range in the
wavenumber range of 4000-600 cm
-1
, at resolution 4 cm
-1
with 32 scans per sample. Samples (10 mm × 10 mm) were
dried in the oven at 50
°
C until constant weight before
characterization. The samples were made in powder and
mixed with KBr as well as followed by the pressure within
the pellet ultrathin layer [26].
2.3.5. Moisture Absorption and Water Vapor Permeability.
Moisture absorption (MA). MA was determined using the
method described in a previous study [27]. All biocomposite
samples were dried in a drying oven (Memmert Germany,
Model 55 UN) at 50
°
C until a constant weight was achieved.
The dried sample was stored in a closed chamber with 75%
RH at 25
°
C. The samples were we ighed every 30 min for 7 h
with a precision balance (Kenko) with a 0.1 mg accuracy.
MA was calculated using
MA = whwo
ðÞ
wo
,ð2Þ
where whis the nal weight and wois the initial weight of the
sample. MA determination was repeated 5 times for each
lm.
Water vapor permeability (WVP). WVP was measured
according to the method described by previous work [28].
WVP determination was repeated 3 times for each lm.
2.3.6. Thermogravimetric Analysis (TGA) and Derivative
Thermogravimetry (DTG). TGA and DTG of all samples
were characterized using a dierential scanning calorimeter
(Linseis TA type PT 1600, Germany). About 25 mg of the
lm was positioned on the microbalance located inside the
Table 1: Composition of the starch lm, chitosan lm, and biocomposite lms used in the study.
Sample code Tapioca starch
(g)
Nanober suspension
(mL)
Dried nanobers
(g)
Aquades
(mL)
Glycerol
(mL)
Chitosan
(g)
Acetic acid
(mL)
GU 10 ——100 2 ——
CH —— —2.5 100
GU/CH 10 ——100 2 2.5 100
GU/CH/10BC 10 10 0.068 100 2 2.5 100
GU/CH/15BC 10 15 0.102 100 2 2.5 100
GU/CH/20BC 10 20 0.136 100 2 2.5 100
10% NaOH, 12 h Electrical blender
12000 rpm, 1 h
High-shear homogenizer
60 min,
10000 rpm
e suspension was filtered
200T mesh cheesecloth (74 𝜇m)
BC suspensionsGelatinized starchChitosan gel
Bionanocomposite
film
Dried in a drying
Poured onto a
petri dish
(d = 145 mm)
Blended using
ultrasonic cell
crusher at 600 W
Suspension (starch/BC) and chitosan gel
with composition 80:20
(70 g total weight)
Figure 1: Steps of sample preparations for biocomposite.
3International Journal of Polymer Science
furnace. The test was carried out from 35
°
C up to 550
°
C with
a heating rate of 10
°
C/min in a nitrogen atmosphere.
2.3.7. Tensile Properties. Tensile properties of the biocompo-
sites, including tensile strength and elongation at break, were
measured using COM-TEN 95T Series 5K (Pinellas Park,
USA) and were performed according to the ASTM D 638
type V standard. Before the test, all samples were stored in
a desiccator with 50 ± 5%relative humidity at 25
°
C for 48 h.
Samples were then tested at room temperature and RH 75%
using a tensile test speed of 5 mm/min. The testing of the lm
was repeated at least three times for each ber content.
2.3.8. Antimicrobial Activity. The antibacterial activity of the
starch/chitosan-based biocomposite lms was assayed using
the agar diusion method (Bauer, Kirby, Sherris, and Turck,
1966). Four microbe strains were used: Gram-positive Staph-
ylococcus aureus and Bacillus subtilis bacteria and Gram-
negative Escherichia coli and Pseudomonas aeruginosa. The
microbial suspensions in saline solution (NaCl 0.85% sterile)
were standardized using the McFarland scale to inoculate
petri dishes containing nutrient agar for bacter ia. 6 mm lm
diameter disks were placed on the inoculated agar then incu-
bated at 30
°
C for 24 h. The diameter of the growth inhibition
zones around the lm disks was gauged visually. All tests
10.0 kV 10 𝜇m8.3 3 SEM_SEI
(a)
10.0 kV 10 𝜇m8.9 3 SEM_SEI
(b)
10.0 kV 100 𝜇m8.9 3 SEM_SEI 10.0 kV 10 𝜇m8.8 3 SEM_SEI
(c)
10.0 kV 1 𝜇m9.5 3 SEM_SEI
(d)
10.0 kV 1 𝜇m9.0 3 SEM_SEI
(e)
Figure 2: FESEM images of the fracture surface for the pure starch lm (a), pure chitosan lm (b), starch/chitosan lm (c), GU/CH/15BC
lm (d), and GU/CH/20BC lm (e).
4 International Journal of Polymer Science
were carried out in triplicate, and the antibacterial activity
was expressed as the mean of the inhibition diameters (mm).
2.3.9. Statistical Analysis. Experimental data were analyzed
using IBM SPSS Statistics 25.0 (IBM Corporation, Chicago,
USA). One-way analysis of variance (ANOVA) and a ptest
were used to identify the signicance of any eects of varying
nanober content on properties of the biocomposites. Dun-
cans multiple range tests were used on the MA and WVP
results using a 95% (p0:05) condence level.
3. Results and Discussion
3.1. Morphological Biocomposites. Figure 2 displays FESEM
fracture surface micrographs for the pure starch lm (a), pure
chitosan lm (b), starch/chitosan lm (c), GU/CH/15BC lm
(d), and GU/CH/20BC lm (e). The surface of the starch lm
was rough (Figure 2(a)) probably as a result of a long tortu-
ous way of the polymer chains looking for the weak section
of the chain structure. Meanwhile, the CH lm had a smooth
fracture surface (Figure 2(b)) attributed to unimpeded crack
propagation and corresponding to its brittle properties.
10 20 30 40 50 60
0
200
400
600
800
1000
1200
1400
Intensity (count)
Shi
GU
CH
GU/CH
GU/CH/10BC
GU/CH/15BC
GU/CH/20BC
Figure 3: The XRD patterns of all samples.
Table 2: Crystallinity index, opacity value, and thermal properties of all samples.
Films Crystallinity index (%) Opacity (AUnm)Maximum decomposition temperature (
°
C)
GU 11 280:5±0:16
d
268
CH 17 450:8±0:04
f
311
GU/CH 14 202:3±0:98
a
307
GU/CH/10BC 35 208:1±1:09
b
313
GU/CH/15BC 36 237:9±0:63
c
314
GU/CH/20BC 37 328:3±0:76
e
317
Dierent letters a, b, c, d, e, and f in the same column indicate signicant dierences in means (p0:05).
100
80
60
40
20
0
400 500 600 700 800
Wavelength (nm)
Transmittance (%)
GU
CH
GU/CH
GU/CH/10BC
GU/CH/15BC
GU/CH/20BC
Figure 4: The transmittance of all samples.
5International Journal of Polymer Science
Figure 2(c) displays the GU/CH fracture surface, which was
rougher than the CH sample. The dierent chemical struc-
tures of both these substances produce a weaker structural
section in which the crack propagates with a longer tortuous
way resulting in microscopic features known as a beach mark
as shown by the yellow arrow in the inset of Figure 2(c) which
marks an interruption of the cracking progress. Adding BC
into the blends continuously increases the surface roughness
of the biocomposite (yellow arrow in Figures 2(d) and 2(e)).
In these gures, disintegrated BC nanobers were dispersed
homogeneously after ultrasonication. A similar result also is
supported by previous studies [14, 17, 24, 29, 30].
3.2. X-Ray Diraction. The X-ray diraction curves for all
studied lms are shown in Figure 3. All lms show a similar
semicrystalline pattern with prominent peaks at about 2θ=
20°and 23
°
. The crystallinity index (CI) of each sample is
shown in Table 2. The GU/CH blend lm has a CI value of
14% between the CI of chitosan (17%) and the CI of the
starch lm (11%). The addition of any fraction of nanobers
to the starch-based matrix improves the CI value of the
biocomposite lms (around 164% increase compared to
GU/CH). This increased value indicates better ller disper-
sion in the starch matrix thanks to ultrasonication [2]. Sam-
ples before mixing with nanobers display the main peak
position at 2θ=23
°. The addition of the nanobers shifted
the position toward the left side (2θ=20
°). According to pre-
vious work, shifting the peak position to the left side can be
associated with an increase in tensile residual stress resulting
from increases in the polymer chainsinterlayer spacing [31].
3.3. Transparency. In food packaging applications, high
transparency can be an essential property [32]. The transpar-
ency values for all samples are displayed in Figure 4. The GU
lm displays low transmittance at all wavelengths. The high-
est transparency at 800 nm belongs to the CH lm. There-
fore, after mixing starch with chitosan, the GU/CH lm
became more transparent. However, the BC nanober addi-
tion to this GU/CH lm signicantly decreased (p0:05)
the transparency of the biocomposite lm. This phenomenon
is because increasing amounts of the nanober increase the
amount of reected light in the lm. The lowest transparency
was consequently measured on the lm with the highest ber
loading, the GU/CH/20BC lm (28.9% less than the GU/CH
lm). This value still agrees with previous work [33]. Despite
the decreasing transparency, this lm was still clear enough
to see through easily. The consistency of transparency read-
ings in each of the repeats for each ber loading, as shown
by the small standard deviation values, conrms that the
cellulose bers are homogeneously dispersed in the starch-
based matrix. This result is consistent with Figure 2(e) which
shows beach marks spread evenly on all fracture surfaces of
the biocomposite lm.
3.4. FTIR Spectra. Structural changes in the starch-based lm
after mixing with chitosan or/and nanobers can be observed
using an FTIR curve. Figure 5 displays an FTIR curve of the
mean transmittance values for three samples of each biocom-
posite. Shifts of peak intensity, broadening of absorption
peaks, and appearance of new bands in the FTIR spectra cor-
respond to structural changes [34]. There are absorption
peaks in the GU lm at about 3247 cm
1
(OH stretching)
and 2917 cm
-1
due to CH stretching. The band at 1650 cm
-1
is present due to the deformation vibration of the absorbed
water molecules. The characteristic absorption peak of chito-
san is the band at 1559 cm
-1
, which is assigned to the stretch-
ing vibration of the amino group of chitosan. Another band
at 3367 cm
-1
is due to amine NH symmetric vibration. The
wavenumber and Tvalue of O-H stretching shift from
100
80
60
40
20
0
4000 3500 3000 2500 2000 1500 1000
Wavenumber (cm1)
Transmittance (%)
GU
CH
GU/CH
GU/CH/10BC
GU/CH/15BC
GU/CH/20BC
(a)
GU
CH
GU/CH
GU/CH/10BC
GU/CH/15BC
GU/CH/20BC
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
3500 3450 3400 3350 3300 3250 3200 3150 3100 3050
Absorbance
Wavenumber (cm1)
(b)
Figure 5: FTIR spectrum resulting from triplicate measurements of each lm. The full spectrum from 4000 to 250 cm
-1
(a). Sections of the
spectrum for O-H stretching vibration (b).
6 International Journal of Polymer Science
3290 cm
1
and 17.9% for the GU lm to 3288 cm
1
and
19.6% for the GU/CH blend lm. Similar shifting was also
observed on the GU/CH-based biocomposite lm due to
the presence of nanobers. For example, Tof the GU/CH
lm at about 3290 cm
1
is 19.6% but 24.5% for the
GU/CH/20BC lm. As expected, increasing concentrations
of BC increased the peak Tand shifted the wavenumber of
O-H functional groups. This case is probably a result of
increasing hydrogen bonds between starch and/or nanober
polymer chains and amino functional groups [35].
3.5. Moisture Absorption and Water Vapor Permeability.
Figure 6(a) shows the moisture absorption (MA) of the pure
starch lm, chitosan lm, and starch/chitosan-based biocom-
posite lm. The pure chitosan lm has the lowest MA (17.7%
after 8 h in the humid chamber). The presence of nanobers
25
20
15
10
5
0
021 3456789
7.0 7.5 8.0 8.5 9.0
16
18
20
22
A
B
C
C
D
E
1
Time (h)
Moisture absorption (%)
7.0
7.5
8.0
18
20
22
A
B
C
C
D
E
GU
CH
GU/CH
GU/CH/10BC
GU/CH/15BC
GU/CH/20BC
(a)
2.50E010
2.00E010
1.50E010
1.00E010
5.00E011
5.00E011
0.00E+000
0.00E+000
8 12162024
Time (h)
20 24
A
A,B
B,C
C,D
D
C,D
GU
CH
GU/CH
GU/CH/10BC
GU/CH/15BC
GU/CH/20BC
(b)
Figure 6: Average value of moisture absorption (a) and WVP (b) of each studied lm. Dierent letters a, b, c, d, and e in the inset indicate
signicant dierences (p0:05).
7International Journal of Polymer Science
in the starch/chitosan-based lm results in a decrease in MA
of the biocomposites. Higher nanober loading leads to
lower average MA values. This result is because nanober
and chitosan are more hydrophobic than the neat starch lm.
Dispersion of llers in the starch lm homogeneously results
in decreasing MA of the biocomposite lm. Better intermo-
lecular hydrogen bonding between the starch matrix and
the chitosan and ber improves the moisture resistance of
the biocomposite lm due to reducing the number of free
hydroxyl groups. This result is consistent with the FTIR pat-
tern (Figure 5) showing the weakest intensity of O-H stretch-
ing and O-H of absorbed water peaks in the lms with the
highest nanober content due to the reduction in free
hydroxyl groups. Similar ndings of reducing moisture
absorption with increased ber loadings have also been
reported previously [30]. Figure 5(b) shows water vapor per-
meability (WVP) of both starch and chitosan lms and bio-
composite lms. As expected, the pattern of WVP with the
addition of nanobers is similar to that of MA. There is a
decrease in the WVP value in lms that contain more nano-
bers. WVP of the GU/CH/20BC sample is 27% lower than
that of the GU/CH lm after 24 h. The decrease in WVP is
because moisture is absorbed less readily into the biocompo-
site for the reasons described above.
Also, well-dispersed nanober hinders the path for water
molecule diusion through the lm due to the more com-
pact, homogeneous polymer structures [29]. As shown in
Figure 6, the WVP value of the GU/CH/20BC lm is 3:
1011 gm
1s1Pa1(24 h), similar to that found in a previous
study on the improvement of the shelf life of yam starch/chi-
tosan-coated apples [36]. Therefore, this lm has a good
potential for the shelf life of various food types.
3.6. Thermal Properties. Figures 7(a) and 7(b) show TGA and
DTG curves of each tested lm as a function of temperature.
There are three stages of weight loss of the lm shown in a
TGA graph. The rst stage at 100-150
°
C is related to weight
loss in the lm due to the evaporation of the absorbed mois-
ture. This small amount of dehydration is evident in the DTG
curve (Figure 7(b)). The weight loss for the second stage at
250-350
°
C is attributed to the decomposition of starch, chito-
san, and nanobers. In the temperature range of 360570
°
C,
a third weight loss was observed due to a nal decomposition
to ash. The temperature of the maximum decomposition rate
(Tm) at the second stage was higher (311
°
C) for starch than
for chitosan (268
°
C). As expected, the addition of chitosan
to starch decreased the Tmvalue slightly (307
°
C). However,
the thermal resistance of the starch/chitosan-based lm
became higher with the addition of nanobers (Figures 7(a)
and 7(b)). For example, the Tmof the GU/CH lm increased
from 10
°
C to 317
°
C after adding dried nanobers of 0.136 g.
This increased value is probably because of the higher crys-
tallinity in the sample [30]. Also, the higher thermal resis-
tance resulted from better interfacial hydrogen bonding
between starch and nanober dispersed homogeneously
[17, 30]. This result is consistent with the high CI value of
lms with high nanober content, as shown in the XRD
curve (Table 2).
3.7. Tensile Properties. Figure 8 shows the tensile properties
of all tested samples. TS for the GU/CH lm was 2.6 MPa, a
value between pure GU (1.4 MPa) and CH (3.2 MPa). As
expected, the nanober addition to the GU/CH lm led to
an increase in its TS value. The maximum TS was 4.7 MPa,
measured on the GU/CH/20BC lm, reinforced with the
highest ber loading (0.136 g). This increased TS value prob-
ably results from the increased crystallinity index (Table 2),
better nanober dispersion (Figure 2), and better interfacial
hydrogen bonding between the nanobers with the GU/CH
matrix [20]. The pure chitosan lm is the least brittle of all
the lms with an EB of only 0.98%. After mixing the chitosan
with starch, the EB of the GU/CH lm became higher (11%).
GU
CH
GU/CH
GU/CH/10BC
GU/CH/15BC
GU/CH/20BC
100 200 300 400 500
0
20
40
60
80
100
Weight (%)
(a)
100 200 300 400 500
1.4
1.2
1.0
0.8
0.6
0.4
0.2
0.0
0.2
0.4
DTG (%/min)
GU
CH
GU/CH
GU/CH/10BC
GU/CH/15BC
GU/CH/20BC
(b)
Figure 7: TGA (a) and DTG (b) charts of all samples.
8 International Journal of Polymer Science
Even the addition of the BC nanobers to the GU/CH lm
improved its EB. This tendency is attributable to longer tor-
tuous pathways of the crack propagation through the matrix
due to the BC nanobers. Further nanober loading did not
result in statistically signicant changes in EB of the biocom-
posite lm.
3.8. Antibacterial Activity. Table 3 displays the diameters of
antibacterial activity inhibition zones against all microorgan-
isms tested in studied samples. BC nanobers did not inhibit
any microorganisms. This phenomenon is in good agree-
ment with previous work [37]. A similar appearance was also
displayed by the GU lm without antibacterial activity. How-
ever, all chitosan-contained lms were eective against
Gram-positive bacteria and Gram-negative bacteria. This
result could be due to the numerous active ingredients pres-
ent in chitosan. Chitosan could adsorb the electronegative
substance in the cell, and it disturbs the physiological activi-
ties of the bacteria and kills them [38].
4. Conclusion
This work characterized a tapioca starch/chitosan-based lm
reinforced by bacterial cellulose nanober. All chitosan-
based lms had antibacterial activity. 0.136 g nanober addi-
tion to this lm led to the highest tensile strength and the
highest thermal resistance. The presence of nanobers
increased moisture resistance and water barrier properties.
The addition of the nanobers led to a decrease in transpar-
ency. However, the resulting translucent biocomposite lm
could still be seen through clearly. Overall, this biocomposite
lm could become a food packaging alternative for replacing
hydrocarbon-based plastics.
Table 3: Antibacterial activity of the lms.
Films Diameter of inhibition zones (mm) against microorganisms
SA BC EC PA
CH 20:9±1:913:9±7:212:8±0:215:1±0:1
GU 0000
GU/CH 18 ± 1:612:3±6:412:3±6:912:0±3:5
GU/CH/10BC 14:7±6 11:9±6:111:4±5:511:3±6:5
GU/CH/15BC 12:1±7:210:5±4:212:5±2:510:3±3:9
GU/CH/20BC 14:5±5:810:6±4:910:3±1:513:4±3:3
BC nanobers 0000
SA = Staphylococcus aureus;BC=Bacillus subtilis;EC=Escherichia coli;PA=Pseudomonas aeruginosa.
1.4
3.2
2.6
3.6
4.1 4.7
E
D
C
B
C
A
5
4
3
2
1
0
GU
CH
GU/CH
GU/CH/10BC
GU/CH/15BC
GU/CH/20BC
Samples
Tensile strength (MPa)
(a)
C
24.7%
0.98%
A
B
11%
D
43.7%
CC
20.9%
21.5%
GU
CH
GU/CH
GU/CH/10BC
GU/CH/15BC
GU/CH/20BC
Samples
50
40
30
20
10
0
Elongation at break (%)
(b)
Figure 8: Tensile strength (a) and elongation at break (b) of each sample. Dierent letters a, b, c, d, and e in the vertical bar chart indicate
signicant dierences (p0:05).
9International Journal of Polymer Science
Data Availability
We have data supporting the results of this work. The Micro-
soft Word document data used to support the ndings of this
study are available from the corresponding author upon
request. Please email habral@yahoo.com.
Conflicts of Interest
We arm that there is no conict of interest.
Acknowledgments
Acknowledgment is addressed to Universitas Andalas for
supporting research funding with project name PDU
KRP1GB UNAND (number T/5/UN.16.17/PT.01.03/IS-
PDU-KRP1GB/2020).
References
[1] I. Dinika, D. K. Verma, R. Balia, G. L. Utama, and A. R. Patel,
Potential of cheese whey bioactive proteins and peptides in
the development of antimicrobial edible lm composite: a
review of recent trends,Trends in Food Science and Technol-
ogy, vol. 103, pp. 5767, 2020.
[2] M. Asro, H. Abral, A. Kasim, A. Pratoto, M. Mahardika, and
F. Hazulhaq, Characterization of the sonicated yam bean
starch bionanocomposites reinforced by nanocellulose water
hyacinth ber (Whf): the eect of various ber loading,Jour-
nal of Engineering Science and Technology, vol. 13, pp. 2700
2715, 2018.
[3] H. Abral, G. J. Putra, M. Asro, J. Park, and H. Kim, Eect of
vibration duration of high ultrasound applied to bio-
composite while gelatinized on its properties,Ultrasonics
Sonochemistry, vol. 40, no. Part A, pp. 697702, 2018.
[4] I. Gan and W. S. Chow, Antimicrobial poly(lactic acid)/cellu-
lose bionanocomposite for food packaging application: a
review,Food Packaging and Shelf Life, vol. 17, pp. 150161,
2018.
[5] L. V. Cabañas-Romero, C. Valls, S. V. Valenzuela et al.,
Bacterial cellulose-chitosan paper with antimicrobial and
antioxidant activities,Biomacromolecules, vol. 21, no. 4,
pp. 15681577, 2020.
[6] M. Babaee, M. Jonoobi, Y. Hamzeh, and A. Ashori, Biode-
gradability and mechanical properties of reinforced starch
nanocomposites using cellulose nanobers,Carbohydrate
Polymers, vol. 132, pp. 18, 2015.
[7] S. M. Noorbakhsh-Soltani, M. M. Zerafat, and S. Sabbaghi, A
comparative study of gelatin and starch-based nano-composite
lms modied by nano-cellulose and chitosan for food packag-
ing applications,Carbohydrate Polymers, vol. 189, pp. 4855,
2018.
[8] Z. Shariatinia and M. Fazli, Mechanical properties and anti-
bacterial activities of novel nanobiocomposite lms of chitosan
and starch,Food Hydrocolloids, vol. 46, pp. 112124, 2015.
[9] S. Chillo, S. Flores, M. Mastromatteo, A. Conte,
L. Gerschenson, and M. A. Del Nobile, Inuence of glycerol
and chitosan on tapioca starch-based edible lm properties,
Journal of Food Engineering, vol. 88, no. 2, pp. 159168, 2008.
[10] U. Qasim, A. I. Osman, A.. H. al-Muhtaseb et al., Renewable
cellulosic nanocomposites for food packaging to avoid fossil
fuel plastic pollution: a review,Environmental Chemistry Let-
ters, vol. 19, no. 1, pp. 613641, 2021.
[11] P. Cazón and M. Vázquez, Mechanical and barrier properties
of chitosan combined with other components as food packag-
ing lm,Environmental Chemistry Letters, vol. 18, no. 2,
pp. 257267, 2020.
[12] I. M. S. Araújo, R. R. Silva, G. Pacheco et al., Hydrothermal
synthesis of bacterial cellulose-copper oxide nanocomposites
and evaluation of their antimicrobial activity,Carbohydrate
Polymers, vol. 179, pp. 341349, 2018.
[13] S. Y. Z. Zainuddin, I. Ahmad, H. Kargarzadeh, I. Abdullah, and
A. Dufresne, Potential of using multiscale kenaf bers as rein-
forcing ller in cassava starch-kenaf biocomposites,Carbohy-
drate Polymers, vol. 92, no. 2, pp. 22992305, 2013.
[14] M. Asro, H. Abral, A. Kasim, A. Pratoto, M. Mahardika, and
F. Hazulhaq, Mechanical properties of a water hyacinth
nanober cellulose reinforced thermoplastic starch bionano-
composite: eect of ultrasonic vibration during processing,
Fibers, vol. 6, no. 2, p. 40, 2018.
[15] C. M. O. Müller, J. B. Laurindo, and F. Yamashita, Eect of
cellulose bers addition on the mechanical properties and
water vapor barrier of starch-based lms,Food Hydrocolloids,
vol. 23, no. 5, pp. 13281333, 2009.
[16] H. Abral, V. Lawrensius, D. Handayani, and E. Sugiarti, Prep-
aration of nano-sized particles from bacterial cellulose using
ultrasonication and their characterization,Carbohydrate
Polymers, vol. 191, pp. 161167, 2018.
[17] H. Abral, A. Hartono, F. Hazulhaq, D. Handayani,
E. Sugiarti, and O. Pradipta, Characterization of PVA/cassava
starch biocomposites fabricated with and without sonication
using bacterial cellulose ber loadings,Carbohydrate Poly-
mers, vol. 206, pp. 593601, 2018.
[18] T. Bourtoom and M. S. Chinnan, Preparation and properties
of rice starch-chitosan blend biodegradable lm,LWT- Food
Science and Technology, vol. 41, no. 9, pp. 16331641, 2008.
[19] M. M. Marvizadeh, N. Oladzadabbasabadi, A. Mohammadi
Nafchi, and M. Jokar, Preparation and characterization of
bionanocomposite lm based on tapioca starch/bovine gela-
tin/nanorod zinc oxide,International Journal of Biological
Macromolecules, vol. 99, pp. 17, 2017.
[20] M. Mahardika, H. Abral, A. Kasim, S. Arief, F. Hazulhaq, and
M. Asro,Properties of cellulose nanober/bengkoang starch
bionanocomposites: eect of ber loading,LWT- Food Sci-
ence and Technology, vol. 116, article 108554, 2019.
[21] N. A. Al-Tayyar, A. M. Youssef, and R. Al-hindi, Antimicro-
bial food packaging based on sustainable bio-based materials
for reducing foodborne pathogens: a review,Food Chemistry,
vol. 310, article 125915, 2020.
[22] S. Y. Sung, L. T. Sin, T. T. Tee et al., Antimicrobial agents for
food packaging applications,Trends in Food Science and
Technology, vol. 33, no. 2, pp. 110123, 2013.
[23] S. J. Mcgrance, H. J. Cornell, and C. J. Rix, A simple and rapid
colorimetric method for the determination of amylose in starch
products,Starch/Staerke,vol.50,no.4,pp.158163, 1998.
[24] H. Abral, M. M. Kadriadi, D. Handayani, E. Sugiarti, and A. N.
Muslimin, Characterization of disintegrated bacterial cellu-
lose nanobers/PVA bionanocomposites prepared via ultraso-
nication,International Journal of Biological Macromolecules,
vol. 135, pp. 591599, 2019.
[25] R. A. Ilyas, S. M. Sapuan, and M. R. Ishak, Isolation and char-
acterization of nanocrystalline cellulose from sugar palm bres
10 International Journal of Polymer Science
( _Arenga pinnata_ ),Carbohydrate Polymers, vol. 181,
pp. 10381051, 2018.
[26] M. Asro, H. Abral, A. Kasim, and A. Pratoto, XRD and FTIR
studies of nanocrystalline cellulose from water hyacinth
(Eichornia crassipes)ber,Journal of Metastable and Nano-
crystalline Materials, vol. 29, pp. 916, 2017.
[27] H. Abral, R. Soni Satria, M. Mahardika et al., Comparative
study of the physical and tensile properties of jicama (Pachyr-
hizus erosus) starch lm prepared using three dierent
methods,Starch/Staerke, vol. 71, pp. 19, 2019.
[28] H. Abral, A. Basri, F. Muhammad et al., A simple method for
improving the properties of the sago starch lms prepared by
using ultrasonication treatment,Food Hydrocolloids, vol. 93,
pp. 276283, 2019.
[29] A. Khan, R. A. Khan, S. Salmieri et al., Mechanical and barrier
properties of nanocrystalline cellulose reinforced chitosan
based nanocomposite lms,Carbohydrate Polymers, vol. 90,
no. 4, pp. 16011608, 2012.
[30] H. Abral, A. S. Anugrah, F. Hazulhaq, D. Handayani,
E. Sugiarti, and A. N. Muslimin, Eect of nanobers fraction
on properties of the starch based biocomposite prepared in
various ultrasonic powers,International Journal of Biological
Macromolecules, vol. 116, pp. 12141221, 2018.
[31] J. Epp, X-ray diraction (XRD) techniques for materials char-
acterization,Elsevier Ltd, 2016.
[32] W. Cheng, J. Chen, D. Liu, X. Ye, and F. Ke, Impact of ultra-
sonic treatment on properties of starch lm-forming disper-
sion and the resulting lms,Carbohydrate Polymers,vol. 81,
no. 3, pp. 707711, 2010.
[33] D. Merino, A. Y. Mansilla, T. J. Gutiérrez, C. A. Casalongué,
and V. A. Alvarez, Chitosan coated-phosphorylated starch
lms: water interaction, transparency and antibacterial prop-
erties,Reactive and Functional Polymers, vol. 131, pp. 445
453, 2018.
[34] M. S. Goyat, S. Ray, and P. K. Ghosh, Innovative application
of ultrasonic mixing to produce homogeneously mixed
nanoparticulate-epoxy composite of improved physical prop-
erties,Composites. Part A, Applied Science and Manufactur-
ing, vol. 42, no. 10, pp. 14211431, 2011.
[35] H. Abral, M. H. Dalimunthe, J. Hartono et al., Characteriza-
tion of tapioca starch biopolymer composites reinforced with
micro scale water hyacinth bers,Starch/Staerke, vol. 70,
no. 7-8, pp. 18, 2018.
[36] J. C. Martins da Costa, K. S. Lima Miki, A. da Silva Ramos, and
B. E. Teixeira-Costa, Development of biodegradable lms
based on purple yam starch/chitosan for food application,
Heliyon, vol. 6, no. 4, p. e03718, 2020.
[37] J. Kim, Z. Cai, H. S. Lee, G. S. Choi, D. H. Lee, and C. Jo, Prep-
aration and characterization of a bacterial cellulose/chitosan
composite for potential biomedical application,Journal of
Polymer Research, vol. 18, no. 4, pp. 739744, 2011.
[38] L. Y. Zheng and J. F. Zhu, Study on antimicrobial activity of
chitosan with dierent molecular weights,Carbohydrate
Polymers, vol. 54, no. 4, pp. 527530, 2003.
11International Journal of Polymer Science
... Chitosan (CH) is a derived polysaccharide from chitin found in shrimp and/or crustacean shells [25]. CH-based films have a variety of physical-functional attributes that makes them technically appropriate to apply in food contact materials [26]. Previous research has demonstrated that CH is unique in meat or meat products because of the good barrier to oxygen and moisture as well as being non-toxic, biodegradable, and cheap [27]. ...
Article
Full-text available
Edible films and essential oil (EO) systems have the potency to enhance the microbial quality and shelf life of food. This investigation aimed to evaluate the efficacy of chitosan films including essential oils against spoilage bacteria and foodborne pathogens associated with meat. Antimicrobial activity (in vitro and in vivo) of chitosan films (CH) incorporated with oregano oil (OO) and thyme oil (TO) at 0.5 and 1% was done against spoilage bacteria and foodborne pathogens, compared to the control sample and CH alone. Preliminary experiments (in vitro) showed that the 1% OO and TO were more active against Staphylococcus aureus compared to Escherichia coli O157:H7 and Salmonella Typhimurium. In in vivo studies, CH containing OO and TO effectively inhibited the three foodborne pathogens and spoilage bacteria linked with packed beef meat which was kept at 4 °C/30 days compared to the control. The total phenolic content of the EOs was 201.52 mg GAE L −1 in thyme and 187.64 mg GAE L −1 in oregano. The antioxidant activity of thyme oil was higher than oregano oil. The results demonstrated that the shelf life of meat including CH with EOs was prolonged ~10 days compared to CH alone. Additionally, CH-OO and CH-TO have improved the sensory acceptability until 25 days, compared to the control. Results revealed that edible films made of chitosan and containing EOs improved the quality parameters and safety attributes of refrigerated or fresh meat.
... Due to the high crystalline forms of CNCs, it arises in transparent and gas-impermeable forms with extremely high tensile strengths that are up to eight times that of steel [21]. Apart from that, it was found that CNCs derived from bacteria, tunicates, or Valonia are highly polydisperse in their molecular weights without any indication of a level-off degree of polymerization (LODP) [110,120,121]. The absence of a regular distribution of amorphous domains is likely to be the cause of this phenomenon. ...
Article
Full-text available
The increasing global concern over the contamination of natural resources, especially freshwater, has intensified the need for effective water treatment methods. This article focuses on the utilization of Cellulose nanocrystals (CNCs), sourced from lignocellulosic materials, for addressing environmental challenges. CNCs a product of cellulose-rich sources has emerged as a versatile and eco-friendly solution. CNCs boast unique chemical and physical properties that render them highly suitable for water remediation. Their nanoscale size, excellent biocompatibility, and recyclability make them stand out. Moreover, CNCs possess a substantial surface area and can be modified with functional groups to enhance their adsorption capabilities. Consequently, CNCs exhibit remarkable efficiency in removing a wide array of pollutants from wastewater, including heavy metals, pesticides, dyes, pharmaceuticals, organic micropollutants, oils, and organic solvents. This review delves into the adsorption mechanisms, surface modifications, and factors influencing CNCs’ adsorption capacities. It also highlights the impressive adsorption efficiencies of CNC-based adsorbents across diverse pollutant types. Employing CNCs in water remediation offers a promising, eco-friendly solution, as they can undergo treatment without producing toxic intermediates. As research and development in this field progress, CNC-based adsorbents are expected to become even more effective and find expanded applications in combating water pollution. Graphical Abstract
... Chitosan (CH) is a derived polysaccharide from chitin found in shrimp and/or crustacean shells [25]. CH-based films have a variety of physical-functional attributes that makes them technically appropriate to apply in food contact materials [26]. Previous research has demonstrated that CH is unique in meat or meat products because of the good barrier to oxygen and moisture as well as being non-toxic, biodegradable, and cheap [27]. ...
Thesis
Edible films and essential oil (EO) systems have the potency to enhance the microbial quality and shelf life of food. This investigation aimed to evaluate the efficacy of chitosan films including essential oils against spoilage bacteria and foodborne pathogens associated with meat. Antimicrobial activity (in vitro and in vivo) of chitosan films (CH) incorporated with oregano oil (OO) and thyme oil (TO) at 0.5 and 1% was done against spoilage bacteria and foodborne pathogens, compared to the control sample and CH alone. Preliminary experiments (in vitro) showed that the 1% OO and TO were more active against Staphylococcus aureus compared to Escherichia coli O157:H7 and Salmonella Typhimurium. In in vivo studies, CH containing OO and TO effectively inhibited the three foodborne pathogens and spoilage bacteria linked with packed beef meat which was kept at 4 °C/30 days compared to the control. The total phenolic content of the EOs was 201.52 mg GAE L−1 in thyme and 187.64 mg GAE L−1 in oregano. The antioxidant activity of thyme oil was higher than oregano oil. The results demonstrated that the shelf life of meat including CH with EOs was prolonged ~10 days compared to CH alone. Additionally, CH-OO and CH-TO have improved the sensory acceptability until 25 days, compared to the control. Results revealed that edible films made of chitosan and containing EOs improved the quality parameters and safety attributes of refrigerated or fresh meat.
Article
Full-text available
This review discusses the recent advancements in cost-effective fermentation methods for producing bacterial nanocellulose (BC) from food and agro-industrial waste. Achieving economical cell culture media is crucial for large-scale BC production, requiring nutrient-rich media at low cost to maximize cellulose yield. Various pretreatment methods, including chemical, physical, and biological approaches, are stated to break down waste into accessible molecules for cellulose-producing bacteria. Additionally, strategies such as dynamic bioreactors and genetic engineering methods are investigated to enhance BC production. This review also focuses on the environmental impact assessment and updated application challenges of BC such as medical applications, energy storage/electronics, filtration membranes, and food packaging. By providing insights from the recent literature findings, this review highlights the innovative potential and challenges in economically and efficiently producing BC from waste streams.
Article
Postharvest produces including fruits and vegetables are perishable and require preservation for prolonged shelf life. Edible coatings are emerging as a safe and value‐added alternative for traditional preservation methods such as refrigeration and chemical treatments. In this article, we have developed almond gum/chitosan (AC)‐based edible coating/formulation and investigated the effectiveness of coating on improving the shelf life of tomatoes and blueberries. The coated tomatoes and blueberries, after a shelf life of 25 and 20 days, respectively exhibited a significant improvement in weight loss (5%–20%), reduced acidity (1–3 folds), and kept the ascorbic acid content lower when compared to the uncoated tomatoes and blueberries. Both the coated tomatoes and blueberries did not exhibit spoilage indicating antimicrobial activity of the coating. Direct testing of the coating formulations exhibited significant anti‐microbial activity against Staphyococcus aureus and Pseudomonas aeroginosa . The polymer films, prepared using the coating formulation, exhibited low moisture content (34.8 ± 2.02% and 41.5 ± 1.34%), low solubility (42.56 ± 2.3% and 31.31 ± 3.1%), low tensile strength (0.34 ± 0.05 N m ⁻² and 0.75 ± 0.07 N m ⁻² ) and high elongation (63.2 ± 4.12% and 71.5 ± 3.8%). These measured film properties confirm that the polymer film has good flexibility and is suitable for edible packaging. Overall, the newly developed AC edible coating exhibits significant potential for a wide range of applications within the food processing industry, offering a viable substitute for conventional wax and lipid‐based coatings.
Article
Full-text available
Bacteria produce cellulose in the form of an extracellular matrix known for its ultrapure existence. Due to its distinctive features, such as high crystallinity, biocompatibility, biodegradability, excellent mechanical qualities, and high liquid-loading capacity, it has boosted significant advancements in biomedical, pharmaceutical, and healthcare. Bacterial cellulose (BC) has also been widely used as an antimicrobial packaging material/ ingredient in food due to its GRAS attribute. Komagataeibacter genus is a potential producer of BC and has been utilized by researchers and industries for cellulose production. The widely used synthetic culture medium yields cellulose with more prolonged incubation and high production costs. To overcome the drawbacks and to produce cellulose in an economical manner, waste of food, fruit/vegetable peel, starchy kitchen waste, agricultural waste, and other low-cost substrates have been widely exploited as an alternative nutritional source in the culture medium. Cell-free enzyme system for BC production is receiving a lot of attention due to its durability, reproducibility, and efficacy. Also, there has been a burst in the advancement in genetic intervention, genetic toolkit, metabolic engineering, and accessibility to whole genome sequences of BC-producing bacteria to regulate production. Furthermore, in the last few years, due to the increase in demand for BC with novel characteristics, researchers have also explored modern approaches, computational programming, genetic modifications, host recombinant expression, synthetic toolkits, and metabolic engineering for regulating and controlling BC production. This article aims to discuss inexpensive, affordable nutritional sources that can be readily employed for BC synthesis. Additionally, this article will list processes for gene regulation and recombination expression for accomplishing preferred alterations in the obtained cellulose as requisite for a distinct application and for adding new functionalities in the material. Graphical abstract
Article
Full-text available
The extensive use of petroleum-based synthetic and non-biodegradable materials for packaging applications has caused severe environmental damage. The rising demand for sustainable packaging materials has encouraged scientists to explore abundant unconventional materials. For instance, cellulose, extracted from lignocellulosic biomass, has gained attention owing to its ecological and biodegradable nature. This article reviews the extraction of cellulose nanoparticles from conventional and non-conventional lignocellulosic biomass, and the preparation of cellulosic nanocomposites for food packaging. Cellulosic nanocomposites exhibit exceptional mechanical, biodegradation, optical and barrier properties, which are attributed to the nanoscale structure and the high specific surface area, of 533 m 2 g −1 , of cellulose. The mechanical properties of composites improve with the content of cellulose nanoparticles, yet an excessive amount induces agglomeration and, in turn, poor mechanical properties. Addition of cellulose nanoparticles increases tensile properties by about 42%. Barrier properties of the composites are reinforced by cellulose nanoparticles; for instance, the water vapor permeability decreased by 28% in the presence of 5 wt% cellulose nanoparticles. Moreover, 1 wt% addition of filler decreased the oxygen transmission rate by 21%. We also discuss the eco-design process, designing principles and challenges.
Article
Full-text available
One of the most critical challenges for the food packaging industry to overcome is the development of biodegradable coatings from renewable sources. In this work, purple yam starch (PYS), chitosan (CS), and glycerol were blended to obtain biodegradable films for characterization as intended food coatings. The films had a homogeneous surface, and the amount of CS highly influenced the film thickness. Infrared spectroscopy indicated hydrogen bond interactions between PYS and CS in the films. Thermogram data suggested that glycerol contributed to the thermal stability of the films, due to its greater interaction with CS than to the PYS. Finally, the application of a YS/CS film on apples for 4 weeks was able to preserve the fruit quality, as weight loss from the coated apple was significantly lower than the uncoated apple (p = 0.44, Dunnet's posthoc test). YS/CS films have great prospects in the food packaging industry as a new biodegradable coating.
Article
Background Researchers are now focusing on the development of edible films made from food by-products to minimize waste. Cheese whey, which is usually drained, is one of those products that are harmful to the environment. However, several evidence has shown the potential of whey proteins, particularly lactoferrin, as a value-added product. Scope and approach The present review is therefore written with the aim of highlighting the potential of bioactive proteins and peptides as an antimicrobial agent in edible films. Key findings and conclusions Some bioactive proteins and peptides from cheese whey, such as lactoferrin, have a wide range of antimicrobial effects that can be incorporated into the edible film composite to increase the shelf-life of food products. Edible film composite can also be an alternative way of reducing plastic waste on the basis of its biodegradable characteristics. Various edible film composites used may prolong the shelf-life of food products with their endurance in the direction of oxygen, carbon dioxide, and steam transmission.
Article
The production of paper based bacterial cellulose-chitosan (BC-Ch) nanocomposites was accomplished following two different approaches. In the first, BC paper sheets were produced and then immersed in an aqueous solution of chitosan (BC-ChI); in the second, BC pulp was impregnated with chitosan prior to the production of the paper sheets (BC-ChM). BC-Ch nanocomposites were investigated in terms of physical characteristics, antimicrobial and antioxidant properties, and ability to inhibit the formation of biofilms on their surface. The two types of BC-Ch nanocomposites maintained the hydrophobic character, the air barrier properties, and the high crystallinity of the BC paper. However, BC-ChI showed a surface with a denser fiber network and with smaller pore than BC-ChM. Only 5% of the chitosan leached from the BC-Ch nanocomposites after 96 h of incubation in an aqueous medium, indicating that it was well retained by the BC paper matrix. BC-Ch nanocomposites displayed antimicrobial activity, inhibiting growth and having killing effect against the bacteria S.aureus and P.aeruginosa, and the yeast C.albicans. Moreover, BC-Ch papers showed activity against the formation of biofilm on their surface. The incorporation of chitosan increased the antioxidant activity of the BC paper. Paper based BC-Ch nanocomposites combined the physical properties of BC paper and the antimicrobial, antibiofilm and antioxidant activity of chitosan.
Article
Improvements in the effectiveness of packaging materials can help to prevent foodborne pathogens and reduce environmental waste. Traditionally, food is packaged in plastic that is rarely recyclable, negatively impacting the environment. Biodegradable packaging materials play an important role in maintaining the health of ecosystems. However, there are limitations in the utilization of bio-based materials, including poor barrier and mechanical properties which frequently cause a shorter shelf life compared to conventional food packaging materials. The incorporation of different nanomaterial in bio-based polymers such as (Chitosan, potato starch, carboxymethyl cellulose (CMC), corn starch and Arabic gum) can improve the various properties packaging materials by enhancing the antimicrobial activity, therefore preventing foodborne pathogens, correspondingly bringing notable enhancement in the bio-based materials properties as food packaging materials. This review deliberates the potential of using bionanocomposite films to solve the issues of both environmental waste and to reduce the spoilage of food products.
Article
Chitosan is an alternative to synthetic polymers for food packaging. The mechanical and barrier properties of pure chitosan films are promising. Chitosan properties can be modified by combining chitosan with other components such as plasticizers, other polysaccharides, proteins and lipids. Here we review mechanical and barrier properties of composite films based on chitosan. The major points are: (1) compared with synthetic plastic films, an important limitation of chitosan-based films is their mechanical properties, especially their capacity to elongation; (2) chitosan is a polymer that allows an easy combination with other polysaccharides, plasticizers, proteins and lipids; (3) this allows to develop mixed components and modify the film properties according to the nature of the food to be packaged.
Article
With the increasing demand for simple, efficient, environmentally friendly preparation methods to produce cellulose nanofibers for reinforcing a biodegradable film is increased, the role of nanofibers from the pure cellulose produced by bacteria becomes more important. This work characterized bacterial cellulose nanofibers disintegrated using a high shear homogenizer. These nanofibers, in 2.5, 5, and 7.5 mL suspensions, were mixed with PVA gel using ultrasonication. The resulting dried bionanocomposite film was also characterized. Adding nanofiber significantly increases (p ≤ 0.05) on tensile strength, thermal resistance, water vapor impermeability, and moisture resistance of PVA film but not strain at break. Tensile strength, tensile modulus, and elongation at the break of the 7.5 mL nanofiber reinforced film were 37.9 MPa (increased by 38%), 547.8 MPa (increased by 26%), and 10.7% (decreased from 17.2% for pure PVA), respectively compared to pure PVA. Transparency decreases slightly with increased nanofiber content. These properties indicate that this bionanocomposite film has potential in food packaging applications.
Article
Starch granules containing amylopectin-rich fractions like sago starch may remain insoluble and undamaged decreasing properties of the film. The aim of this study is to characterize native sago starch films prepared using ultrasonication. An ultrasonication probe was used during gelatinization for 2.5, 5, and 10 min respectively. Ultrasonication decreases the incomplete gelatinized granules resulting in a film with a more compact structure, and lower moisture vapor permeability than non-treated film. The longest duration resulted in a film with the highest transparency, and the highest thermal resistance. The duration for 5 min increased tensile strength of the film by 227%, and its moisture absorption decreased by 27.39% compared to non-sonicated film. After ultrasonication for 10 min, melting temperature increased by 7% in comparison to non-sonicated film. This work promotes a simple method to improve the tensile and physical properties of starch based film.
Article
The objective of this work is to study the physical and tensile properties of jicama (Pachyrhizus erosus) starch film prepared using three different methods. First, a film was prepared from starch granules after sifting using a sieve shaker. A second film was prepared from starch granules after ultrasonication. Another film was made by sonicating the starch gel. Ultrasonication was performed using an ultrasonic probe. These three different methods had a significant effect on the properties of the film (p≤0.05). The film from the starch granules after sifting using 63 μm mesh size and ultrasonication (labeled as S‐63U film) showed the optimum properties. Opacity for S‐63U film was almost half (48.6%) that of the equivalent non‐sonicated film. S‐63U film had the highest tensile strength (3.1 MPa), the lowest moisture absorption (18% after 8h in a humid chamber) and water vapor permeability. FESEM morphology of the fracture surface of the sonicated film displayed a more homogeneous structure compared to films without ultrasonication.