You are on page 1of 9

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/323760583

Characterization of chitosan edible films obtained with various polymer


concentrations and drying temperatures

Article  in  International Journal of Biological Macromolecules · March 2018


DOI: 10.1016/j.ijbiomac.2018.03.057

CITATIONS READS

0 184

6 authors, including:

Angie Katherine Homez Jara Luis Daniel Daza


University of Tolima 11 PUBLICATIONS   18 CITATIONS   
4 PUBLICATIONS   0 CITATIONS   
SEE PROFILE
SEE PROFILE

Diana MARCELA Aguirre Aldemar muñoz


5 PUBLICATIONS   6 CITATIONS    University of Tolima
8 PUBLICATIONS   257 CITATIONS   
SEE PROFILE
SEE PROFILE

Some of the authors of this publication are also working on these related projects:

MODELING ADSORPTION AT THE AIR-WATER INTERFACE OF FISH PROTEIN BY FUZZY LOGIC View project

Physicochemical characterization of starches View project

All content following this page was uploaded by Angie Katherine Homez Jara on 29 March 2018.

The user has requested enhancement of the downloaded file.


International Journal of Biological Macromolecules 113 (2018) 1233–1240

Contents lists available at ScienceDirect

International Journal of Biological Macromolecules

journal homepage: http://www.elsevier.com/locate/ijbiomac

Characterization of chitosan edible films obtained with various polymer


concentrations and drying temperatures
Angie Homez-Jara a,⁎, Luis Daniel Daza a,b, Diana Marcela Aguirre c,d, José Aldemar Muñoz a,
José Fernando Solanilla e, Henry Alexander Váquiro a
a
Centro de Desarrollo Agroindustrial del Tolima, CEDAGRITOL, Departamento de Producción y Sanidad Vegetal, Facultad Ingeniería Agronómica, Universidad del Tolima, 730006 Ibagué, Colombia
b
Grupo de Investigación en Productos Naturales de la Universidad del Tolima, GIPRONUT, Departamento de Química, Facultad de Ciencias, Universidad del Tolima, 730006 Ibagué, Colombia
c
Corporación Centro de Desarrollo Tecnológico Piscícola Surcolombiano, ACUAPEZ, 410008 Neiva, Colombia
d
Fundación Escuela Tecnológica de Neiva Jesús Oviedo, FET, Av. La Toma, 413001 Rivera, Colombia
e
Departamento de Agroindustria, Facultad de Ciencias Agrarias, Universidad del Cauca, 190003 Popayán, Colombia

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

Article history: Chitosan is a promising material that could be used for the development of edible coatings and films on an indus-
Received 1 February 2018 trial level because of its film-forming, biodegradable, non-toxic, and antimicrobial characteristics. The aim of this
Received in revised form 7 March 2018 work was to evaluate the effect of the polymer concentration (0.5%, 1.0%, and 1.5%) and drying temperature (2 °C,
Accepted 13 March 2018
25 °C, and 40 °C) on the physicochemical, mechanical, and thermal properties of chitosan edible films. Chitosan
Available online 14 March 2018
edible films were successfully produced using various processing conditions. The use of lower drying tempera-
Keywords:
tures had a positive effect on certain properties of the films, such as the moisture content (MC), solubility (S),
Biopolymers water vapor permeability (WVP), and optical properties. However, the use of greater drying temperatures
Thermogravimetric properties (40 °C), combined with a higher chitosan concentration, enhanced certain properties of the films, such as
Chitosan the tensile strength (TS), swelling power (SP), and greenness value, while diminishing their luminosity.
Fourier-transform infrared spectroscopy The chitosan films developed in this study showed many desirable characteristics, which may enable
their future use as packaging for food products.
© 2018 Elsevier B.V. All rights reserved.

1. Introduction The manufacture of edible films involves several factors, which influ-
ence the structure and properties of the film, such as the polymer
Currently, the majority of packaging materials used in the food in- source, polymer chemical structure, type of plasticizer, component con-
dustry are produced from petrochemical polymers. However, this mate- centrations, and drying temperature. The drying temperature and film-
rial is practically non-biodegradable, which causes a negative effect on forming material are the most challenging parameters to control during
the environment. In addition, compounds, such as bisphenol A (BPA), the production of edible films. The drying temperature determines the
are often present within their structures. These compounds can migrate degree of reorganization or crystallinity of the film, which can induce
to food. This is concerning because there is evidence to indicate that changes in the optical, mechanical, and barrier properties of the film
chronic exposure to such compounds is associated with genotoxic activ- [6]. Like the drying temperature, the film-forming material influences
ity and epigenetic modifications [1]. Consequently, there is increased the mechanical and barrier properties of the edible film, as well as its ad-
focus on the development of alternative packaging products based on sorption and surface properties [7]; however, the degree of this influ-
polymers, co-products, and agricultural residues [2]. Edible films can ence depends on the nature, source, structure, and concentration of
be used as substitutes for the synthetic polymers that are currently the material used.
used in a wide range of products. Such edible films are non-toxic, im- Various materials have been used to produce edible films, including
prove food safety, and extend shelf-life. They are also considered an gelatin, starch, gums, pectin, and chitosan [8]. Chitosan is a linear poly-
eco-friendly material because of their short degradation period [3]. saccharide consisting of (1,4)-linked-2-amino-deoxy-β-d-glucan; it is
Moreover, edible films can act as a protective barrier against external obtained through the deacetylation of chitin, which, after cellulose, is
factors, such as microorganisms, physical damage, moisture, oil, and the second most abundant polysaccharide in nature [9]. This compound
vapor [4,5]. could be used for the development of edible coatings and films on an in-
dustrial level because of its biodegradable, biofunctional, biocompatible,
⁎ Corresponding author. non-toxic, film-forming, and antimicrobial characteristics [10]. Chitosan
E-mail address: akhomezj@ut.edu.co (A. Homez-Jara). edible films have been successfully used as packaging materials to

https://doi.org/10.1016/j.ijbiomac.2018.03.057
0141-8130/© 2018 Elsevier B.V. All rights reserved.
1234 A. Homez-Jara et al. / International Journal of Biological Macromolecules 113 (2018) 1233–1240

preserve the quality of a wide variety of fruits, such as raspberry [11], films were calculated using Eqs. (1), (2), and (3), respectively.
apple [12], kiwi [13], strawberry [14,15], cherimoya [16], and loquat
[17]. Chitosan edible films have also been tested with regard to the MC ¼ ½ðW 0 −W 1 Þ=W 0   100 ð1Þ
packaging of meats, such as chicken [18], turkey [19], beef [20], pork
[21], and fish [22]. The use of this type of edible film demonstrated a SP ¼ ½ðW 2 −W 1 Þ=W 1   100 ð2Þ
positive impact on the packaged products. In the case of the packaging
of fruits, the application of the chitosan films resulted in decreased
S ¼ ½ðW 1 −W 3 Þ=W 1   100 ð3Þ
weight loss, ethylene gas production, and microorganism proliferation.
The use of chitosan edible films for the packaging of meats resulted in
delayed bacterial damage and improved sensorial acceptance. However,
to date, there is a lack of information on the formation of chitosan films 2.2.3. Mechanical properties
at cool temperatures. Considering that cooling is one of the most widely The tensile properties of the films were determined according to
used processes with regard to the conservation of food on industrial ASTM D882 [25]. Sample films, with dimensions of 5.0 cm × 1.0 cm,
levels, it is important to characterize the properties of chitosan edible were loaded into a texturometer (LS1, Lloyd Ltd., Largo, FL, USA). The
films produced under these conditions. crosshead speed and initial distance were set as 5 mm/min and 5 cm, re-
Considering the aforementioned factors, the aim of this study is to spectively. The tensile strength (TS (MPa)) and elongation at break (EB
produce edible films via the casting method, and to evaluate the influ- (%)) were determined using the Nexygen Plus software (Lloyd ltd., Ver-
ence of the drying temperature and chitosan concentration on the opti- sion 3.0., Largo, FL, USA). The measurements were performed in
cal, structural, barrier, mechanical, thermal, and surface properties of triplicate.
the resultant films.
2.2.4. Water vapor permeability (WVP)
2. Materials and methods The water vapor permeability (WVP) was determined by a water
method based on ASTM E96 [26], with a few modifications. The film
2.1. Film preparation samples were loaded into stainless-steel permeability cells, in which
6 mL of distilled water was deposited with an exposed area of 9.6 cm2.
The films were prepared using a method described by Costa et al. It was assumed that the RH was maintained at 100% within the cell.
[23], with some modifications. Chitosan solutions, with concentrations The cells were placed in a controlled chamber at a temperature of 25
of 0.5% (CH05), 1.0% (CH10), and 1.5% (CH15), were prepared by dis- ± 2 °C and RH of 75 ± 2%. Every hour, the mass of the cells was recorded
solving chitosan (Low molecular weight 50–190 kDa, Deacetylated using an analytical balance (Pioneer PA224C, OHAUS, Parsippany, NJ,
75–85%, Sigma Aldrich, St. Louis, MO, USA) in 1% v/v acetic solution USA), over a period of 8 h. A fan was placed inside the chamber to en-
(Food grade, Cimpa S.A.S., Colombia), under agitation at 25 °C, over a pe- sure that uniform conditions existed within the chamber. The WVP (g/
riod of 1 h. Subsequently, 0.1% of Tween® 80 (Density at 20 °C, m s Pa) was calculated using Eq. (4). Here, a is the mass of the perme-
1060 kg m−3; Merck, Darmstadt, Germany) and 0.3% of glycerol (food ability cell (g) at time t (s); FT is the film thickness (m); A is the exposed
grade, Sigma Aldrich, St. Louis, MO, USA) were added; the solutions area (m2); Psat is the saturation vapor pressure at the test temperature
were then homogenized at 60 °C, under agitation, for a period of (Pa); RH1 is the RH of the water within the cells, expressed as a fraction;
30 min. The chitosonium acetate solution was vacuum filtered and and RH2 is the RH of the chamber, expressed as a fraction.
allowed to stand overnight. Approximately 0.27 g/cm2 of each solution,
with the various concentrations, was cast in acrylic plates and dried at WVP ¼ aFT=tAP sat ðRH1 −RH2 Þ ð4Þ
2 °C for two weeks (T02), 25 °C for one week (T25), and 40 °C for 24 h
(T40). When dried, the films were stabilized for 24 h within a desiccator
cabinet (DH-1002, Acequilab, Colombia) under a relative humidity (RH) 2.2.5. Optical properties
of 54% (using magnesium nitrate salt) prior to analysis. The OP of the film was measured according to a method described by
Hu et al. [27]. The film samples (9 mm × 40 mm) were placed in a quartz
2.2. Physicochemical properties cuvette, which was then loaded into a spectrophotometer (Genesys 10S
UV-VIS, Thermo Fisher Scientific, Waltham, MA, USA); the absorbance
2.2.1. Thickness was measured at 600 nm. An empty cuvette was used as a blank control.
The thickness of each film was measured directly using a microme- The OP was calculated using Eq. (5), where Abs600 is the absorbance at
ter (193–101, Mitutoyo Manufacturing Co. Ltd., Tokyo, Japan). The re- 600 nm, and FT is the film thickness (mm).
sults were used to calculate the values associated with the mechanical,
water permeability, and opacity (OP) properties. OP ¼ Abs600 =FT ð5Þ

2.2.2. Moisture content (MC), swelling power (SP), and solubility (S) of the The color parameters of the films were determined using a Minolta
films colorimeter (Cr 410, Konica Minolta, Tokyo, Japan). A standard white
The moisture content (MC), swelling power (SP), and solubility (S) plate was used to calibrate the equipment (Y = 86.6, x = 0.3153, y =
were calculated using the methodology reported by Baron et al. [24], 0.3228). The results were expressed according to the CieLab color sys-
with a few modifications. Briefly, pieces of the chitosan edible films tem (L*: lightness, a*: redness, and b*: greenness). Measurements
(2 cm × 2 cm) were cut, and weighed (W0) within pre-weighed porce- were performed on the film samples, and the average results were re-
lain capsules. Subsequently, the pieces were dried at 100 °C for 24 h, and ported [28].
then weighed once again (W1). The dried films were placed in glass bot-
tles containing 50 mL of distilled water, which were then allowed to 2.2.6. Fourier-transform infrared spectroscopy (FTIR)
stand, with occasional agitation, for 24 h at room temperature (~25 The spectra of the chitosan films were obtained using Fourier-
°C). The surface of each swollen film was gently blotted with filter transform infrared spectroscopy (FTIR, IRAffinity-1S, Shimadzu,
paper; subsequently, the weight of the swollen film was measured Japan). The measurements were recorded over the range of
(W2). Finally, each piece was dried at 100 °C for 24 h, and its final weight 4000–700 cm−1, with a resolution of 4 cm−1 and 20 scans at room tem-
was recorded (W3). The MC (%), SP (%), and S (%) values of the edible perature. The attenuated total reflectance technique was applied [29].
A. Homez-Jara et al. / International Journal of Biological Macromolecules 113 (2018) 1233–1240 1235

CH: chitosan; MC: moisture content; S: solubility; SP: swelling power; TS: tensile strength; EB: elongation at break; WVP: water vapor permeability; L*: lightness; a*: redness; b*: greenness; OP: opacity. Values are expressed as means ± standard
2.2.7. Scanning electron microscopy (SEM)

1.98 ± 0.14d

3.55 ± 0.14d
1.72 ± 0.13b

3.13 ± 0.33e
5.13 ± 0.15a

4.09 ± 0.40a
2.43 ± 0.11c

2.17 ± 0.07f

2.13 ± 0.09f
OP (A/mm)
The surface morphologies of the chitosan edible films were exam-
ined using scanning electron microscopy ((SEM) Quanta FEG 250, FEI,
HILLSBORO, OR, USA), with an accelerating voltage of 10 kV. Prior to


the analysis, the films were coated using a gold-sputtering technique.

7.14 ± 0.23a,b,c,e,f

7.49 ± 0.46ª,b,c,e
6.84 ± 0.16b,c,e,f

6.91 ± 0.09b,c,e,f

6.98 ± 0.10b,c,e,f
a,c,e

6.61 ± 0.45b,c,f

12.83 ± 0.03
5.66 ± 0.12d

5.82 ± 0.18d
2.3. Thermal analysis

7.81 ± 0.14
2.3.1. Differential scanning calorimetry (DSC)

b*
Differential scanning calorimetry (DSC) was performed using a ther-
mal analyzer (Q200, TA Instruments, Japan), as described by Rizzi et al.

−0.82 ± 0.01d,f,g
−0.75 ± 0.01d,e,f

−0.76 ± 0.04d,e,f
−0.88 ± 0.00b,f,g
−0.93 ± 0.02b,g
−0.92 ± 0.05b,g

−0.73 ± 0.02d,e
a

−1.04 ± 0.04c
[30]. Samples, weighing approximately 3–6 mg, were placed into alumi-

−1.52 ± 0.01

0.08 ± 0.01
num pans and hermetically sealed. An empty pan was used as a refer-
ence. The samples were cooled to −60 °C at 20 °C/min, and then
heated from −60 °C to 250 °C at a constant rate of 20 °C/min, under a

a*
nitrogen atmosphere. The obtained thermograms were examined

89.1 ± 0.2a,b,c

89.0 ± 0.2a,b,c
88.9 ± 0.0a,b,c
88.9 ± 0.1ª,b,c

88.5 ± 0.1ª,b,c
using the Universal V4.5A software (TA Instruments, Japan). The melt-

a,b

88.4 ± 0.4a,b
89.3 ± 0.6b,c

89.2 ± 0.1b,c

86.3 ± 0.03
88.3 ± 0.2
ing temperature (Tm) and melting enthalpy (Δh) values were
determined.

L*
2.3.2. Thermogravimetric analysis (TGA)

a,b,c,d,e

79.3 ± 8.3a,b,c,d,e
66.3 ± 6.9a,b,c,d
84.5 ± 8.3a,c,d,e

92.2 ± 13.4a,c,e
Thermogravimetric analysis (TGA) was performed using a thermo-

62.5 ± 4.8a,b,d
WVP × 10−11

0.27 ± 0.0g
38.5 ± 4.5f
40.8 ± 0.7f
gravimetric analyzer (TGA-60, Shimadzu, Kyoto, Japan) by following

72.5 ± 5.8
the methodology described by Liu et al. [31]. The percentage weight
loss (WL) and thermal degradation temperature (TD) values of the chito-


san films were determined over the temperature range of 20–600 °C,

13.8 ± 1.72a,d,e

2.46 ± 0.07c,g,h

2.77 ± 0.15c,g,h
a,c,d

7.72 ± 0.13a,c,g
using a heating rate of 20 °C/min, under a nitrogen stream. The MC of

1.50 ± 0.06g,h
15.5 ± 0.81d,e
42.4 ± 3.98b

26.5 ± 2.62f

deviation (n = 3 except for thickness where n = 20). Different superscript letters in the same column mean significantly differences (p b 0.05).
8.51 ± 0.60
the films was determined from the weight loss corresponding to the
first step of the curves.

EB (%)


2.4. Statistical analysis

309 ± 33.2b,d
351 ± 9.1b,c,d

372 ± 22.9b,c

210 ± 13.0e
231 ± 10.8e
398 ± 7.1b,c

656 ± 26.6f
a

107 ± 5.3a
120 ± 4.7
A 32 full-factorial design was applied to evaluate the effects of the TS (MPa)

chitosan concentration and drying temperature on the properties of


the edible films. The mean and standard deviation of the results ob-


tained for the three replicates under each treatment were reported.
a,c,d

181 ± 22.2a,c,d
184 ± 6.58a,c,d

204 ± 13.1a,c,d

214 ± 4.96a,c,d
167 ± 21.5a,b,c

224 ± 20.9a,d

236 ± 22.3a,d
113 ± 6.57b,c
Analysis of variance (ANOVA) at the 95% confidence level was used to
198 ± 29.3

determine whether there were any statistically significant differences


SP (%)

between the mean results associated with each property and treatment.
Multiple range tests (MRT), using the Bonferroni method, were used to

determine which mean results were significantly different from others.
25.1 ± 1.64b

24.6 ± 1.62b

24.4 ± 0.35b
a

32.3 ± 3.46a

19.6 ± 1.66c
19.2 ± 0.68c
19.3 ± 0.41c

18.4 ± 0.15c
30.0 ± 0.22

ANOVA and MRT were performed using the “anova1” and


“multcompare” functions of the Statistic Toolbox of MATLAB® R2017b
Physicochemical properties of chitosan edible films obtained at different temperatures.

S (%)

(The MathWorks Inc., Natick, MA, USA), respectively.



Thickness (μm)

3. Results and discussion


a

24.7 ± 14.9a
39.3 ± 14.1a
34.5 ± 9.95a

30.6 ± 11.4a
32.8 ± 4.23a
23.9 ± 4.98a
30.8 ± 7.81a
28.4 ± 9.93ª
32.4 ± 6.10

3.1. Moisture content, swelling power, and solubility


The moisture-binding abilities of biodegradable edible films can af-


17.7 ± 0.75b,c,f

19.4 ± 0.23c,e,f

fect significantly their physical and barrier properties; therefore, the


16.2 ± 1.30b,c

16.5 ± 0.08b.c
21.3 ± 0.23e,f
12.7 ± 1.10d

10.8 ± 0.69d

10.8 ± 0.53d
a
31.3 ± 0.74

7.73 ± 0.07

water content of biopolymeric materials is a key parameter with regard


to their applications [32]. The MC values of the chitosan edible films
MC (%)

ranged between 10.8 and 31.3% (Table 1). The chitosan concentration
did not affect the MC. The drying temperature had an inversely propor-
Denomination

tional effect on the MC values of the chitosan edible films. Thus, the sam-
ples that were dried at low temperatures presented higher MC values.
CH05T02
CH10T02
CH15T02
CH05T25
CH10T25
CH15T25
CH05T40
CH10T40
CH15T40
Chitosan

This result can be explained in terms of the drying time; a longer drying
time was required when a low temperature was used during the pro-
cess. Such extended drying times resulted in reorganization of the solu-
CH (%w/v)

tion structure; this is owed to the long-term interaction, due to


hydrogen bonds and Van der Waals forces, between the water mole-
0.5
1.0
1.5
0.5
1.0
1.5
0.5
1.0
1.5

cules and hydrophilic tails of the chitosan. Owing to this interaction,


an ordered agglomerate was formed; this resulted in the development
T (°C)
Table 1

of a gel-net structure and the subsequent formation of the film. In con-


25

40
2

trast, the transport phenomena that occur at higher drying


1236 A. Homez-Jara et al. / International Journal of Biological Macromolecules 113 (2018) 1233–1240

temperatures limit the gel-formation step; this is owing to rapid water 1.0% chitosan that were dried at 2 °C (CH10T02). The temperature had
evaporation. Therefore, the structure of the resultant film is disordered an inversely proportional effect on the EB; specifically, the EB decreased
[6,32]. The obtained results correlate with those reported by Bourbon as the drying temperature increased. This behavior could be related to
et al. [33], which were obtained for a chitosan edible film, with a chito- the water inclusion phenomena that were discussed in Section 3.1.
san concentration of 2% that was dried at 35 °C. However, they were When long drying times are used, water is incorporated into the struc-
lower than those reported by Pereda et al. [34] for films produced ture of the edible films. The water then functions as a plasticizer; this ex-
with the same parameters. These differences can be explained by the plains the high EB values obtained for the films that were dried at low
absence of Tween 80® in the latter formulation. temperatures. These results are comparable with those reported by
The SP values of the chitosan edible films ranged between 113 and Baron et al. [24] for chitosan films with a chitosan concentration of 2%
236%; the sample that was prepared with the highest chitosan concen- that were dried at 30 °C. According to Lima et al. [36], the obtained re-
tration (1.5%) and dried at 40 °C (CH15T40) showed the highest SP sults suggest that the chitosan edible films exhibit suitable characteris-
value (p b 0.05) (Table 1). In general, under all chitosan concentrations, tics to support food-based applications because they possess high TS
the samples that were prepared at 40 °C presented the highest SP values, and slightly deform before breaking.
values. These were followed by the films that were prepared at the low-
est temperature (2 °C). In addition, the chitosan concentration only 3.3. Water vapor permeability
shows a direct correlation with the SP in the case of the samples pre-
pared at 25 °C (R2 0.98); here, an increase in the chitosan concentration The WVP is an important feature with regard to the application of ed-
resulted in the production of films with increased SP. The SP refers to the ible films as food packaging materials. It reflects the ability of the film to
ability of a film to hold water in its matrix; this is related to the presence prevent or reduce the transfer of moisture between food and its sur-
of hydrophilic groups in its structure, such as carboxylic and hydroxyl roundings [27]. As shown in Table 1, the WVP values ranged between
groups, which can easily interact with water [35]. In addition to chito- 0.27 × 10−11 g/m s Pa and 92.2 × 10−11 g/m s Pa (p b 0.05). The lowest
san, the edible films also contain glycerol; glycerol has a hydrophilic na- WVP value (0.27 × 10−11 g/m s Pa) was obtained for the edible film with
ture, which can result in the swelling of the samples. Moreover, the a chitosan concentration of 1.5% that was dried at 40 °C (CH15T40). The
interaction between the polymer chains and plasticizer (glycerol) can chitosan concentration did not have a consistent effect on the WVP
be improved when the film is processed at elevated temperature. This values. The edible films that were dried at lower temperatures showed
is due to an increase in the mobility of the matrix, which enhances its af- the highest WVP values; this could be associated with the MC of the film.
finity with water molecules, rendering it more susceptible to swelling. The presence of water molecules within the structural matrix of the
Shahbazi [35] reported lower SP values than those obtained in this film, owed to the plasticization process performed under a high-RH at-
study. This could be associated with the comparatively high molecular mosphere, results in the formation of amorphous areas. This promotes
weight of the chitosan used for the film preparation. internal rearrangement and allows water to permeate [39,40]. The
S is an inverse measure of the resistance of a film towards water [34]. WVP results obtained for the chitosan edible films in this study were su-
The S values of the chitosan edible films ranged from 19 to 32% perior to those reported for chitosan and chitosan/Carboxymethyl cellu-
(Table 1). As expected, the films that were dried at the lowest temper- lose (CMC) films (in the order of 4.3 × 10−3 g/m s Pa) [27], and for 2%
ature showed the highest S values (p b 0.05); this could be owed to chitosan films prepared at 35 °C (1.6 × 10−10 g/m s Pa) Bourbon et al.
the greater inclusion of water molecules within the film matrix, as ex- [33]. These chitosan edible films are considered suitable to provide a
plained in the above section. Lima et al. [36] obtained S values of 22% protective barrier against the surrounding environment and to prevent
for chitosan films with a chitosan concentration of 1.5% (w/v) that the loss of water from the packaged material.
were dried at 38 °C. Baron et al. [24] obtained S values of 7.6% for chito-
san films that were dried at 30 °C. In this study, we obtained greater S 3.4. Optical properties
values than those typically reported in literature. This could be ex-
plained by the comparatively high plasticizer content used in this The color and transparency of edible films, as a packaging material,
study, and its hydrophilic nature. are crucial factors with regard to consumer acceptance. Highly transpar-
ent and glossy packages help to improve the visual characteristics of
3.2. Mechanical properties food products [41]. The effects of the chitosan concentration and drying
temperature on the optical properties of the films are shown in Table 1.
It is important to evaluate the mechanical properties of edible films The obtained L* values ranged from 88.3 to 89.3. The highest L* value
because they indicate the durability of such films, and their ability to was achieved for the film with a 1.0% chitosan concentration that was
preserve food integrity during handling, shipping, and storage [35]. dried at 25 °C (CH10T25). It is noteworthy that under the lowest drying
The obtained TS and EB values of the chitosan edible films are shown temperature, the chitosan concentration has a positive effect on the lu-
in Table 1. The TS refers to the ability of a material to resist fracture minosity of the material. However, when the film is processed at a dry-
under tensile stress [37]. The TS values of the chitosan edible films ing temperature of 40 °C, the opposite effect is observed; the L* values
ranged between 107 and 655 MPa (p b 0.05). The drying temperatures decrease as the chitosan concentration increases. Increased interaction
and chitosan concentration had positive effects on the TS; specifically, between the polymer chains and plasticizer occur at the higher drying
the TS values increased when the drying temperatures and chitosan temperatures; this synergy leads to an increase in the level of polymeric
concentrations increased. Elevated drying temperatures promote chain compaction, which can modify the refractive index and restrict
greater interaction between the polymer chains; this, in combination the passage of light through the film matrix [42]. This phenomenon
with increased chitosan concentrations, resulted in greater TS values. can explain the above results. The a* (green–red) values were between
In the case of chitosan films with a chitosan concentration of 2% w/v −1.52 and −0.73. This indicates that the chitosan edible films show a
that were dried at 35 °C, Pereda et al. [34] obtained TS values of slight tendency towards greenness. The chitosan concentration had a
17.3 MPa. This difference can be attributed to the degree of positive effect on the a* values of the edible films dried at 2 °C and 25
deacetylation of the chitosan used in the preparation of the edible films. °C; when the polymer concentration increased, the a* values increased.
The EB value of an edible film represents its capability to resist shape The chitosan concentration did not show a significant effect on the films
changes without crack formation (there is more flexibility when the that were dried at 40 °C. However, the b* values ranged between 5.66
sample is subjected to tension and mechanical stresses) [38]. The EB and 7.81, indicating that the chitosan edible films show a tendency to-
values of the chitosan edible films ranged between 1.5 and 42% (p b wards yellowness. The presence of a yellow color may be considered a
0.05). The highest EB value was measured for the edible films with natural characteristic of chitosan films, since this color is associated
A. Homez-Jara et al. / International Journal of Biological Macromolecules 113 (2018) 1233–1240 1237

with the presence of repeat units of β-(1-4)-2 amino-2-deoxy-D-


glucopyranose [43]. As the drying temperature was increased, there
was a negative effect on the b* values; amongst the samples, the lowest
yellowness values were obtained for the films that were dried at 40 °C.
In addition, in the case of the films that were processed with a drying
temperature of 40 °C, the chitosan concentration exhibited a positive ef-
fect on the b* values. The lowest yellowness value was measured for the
edible films that were produced using a chitosan concentration of 1.0%
and dried at 25 °C (CH10T25). Zhang, et al. [44] measured the color pa-
rameters of chitosan edible films with chitosan concentrations of 2% w/v
that were dried at 40 °C. Results of L* = 65.3, a* = − 0.84, and b* = 5.54
were obtained. However, López-Mata et al. [43] obtained values of L* =
59.5, a* = −1.30, and b* = 14.7 for chitosan films with concentrations
of 2% that were dried at 37 °C.
The OP is an inverse measure of transparency; thus, the transparency
of the film decreases as its OP increases. The OP values of the films in this
study ranged between 1.72 and 5.13 (Table 1). The highest OP value was
obtained for the film with the lowest chitosan concentration that was
dried at 2 °C. The MC appears to have a positive effect on the OP; the
OP value of the film increases as its MC increases. The presence of
water prevents light transmission through the resultant film [45]. The
results obtained in this study were greater than those reported by Hu
et al. [27] and Liu et al. [46] for chitosan edible films dried at 50 °C and
60 °C, respectively.

3.5. Thermal and thermogravimetric properties

To determine the thermal transitions of the chitosan film samples,


DSC was performed. The DSC results are shown in Table 2. An endother-
mic peak, below 190 °C, was observed (Fig. 1) within the thermograms;
this can be attributed to the Tm of the chitosan edible films. The Tm
values ranged between 155 °C and 181 °C. Amongst the films, the film
with a chitosan concentration of 0.5% that was dried at 2 °C
(CH05T02) presented the highest Tm value. It is interesting to note the
lowest Tm values were obtained for the films with chitosan concentra- Fig. 1. DSC thermogram of chitosan edible films and pure chitosan (powder).

tions of 0.5% that were dried at 25 °C and 40 °C (CH05T25 and


CH05T40), respectively. These results could be explained in terms of chitosan, a polymer, is responsible for the formation of crystals in the
the drying time (explained in Section 3.1); a film with an ordered struc- system [47]. The results obtained in this study were greater than those
ture is more difficult to melt than a film with a disordered net. This is be- reported by Lima et al. [36] for edible films with chitosan concentrations
cause the chains have lower mobility due to the occurrence of of 1.5% that were dried at 38 °C.
electrostatic interactions [36]. However, in the case of the films that TGA was used to evaluate the thermal stability and weight degrada-
were dried at 2 °C, the chitosan concentration exhibits an inverse corre- tion of the chitosan edible films. Fig. 2 shows the weight-loss (WL) and
lation with the Tm values; the Tm values decrease as the chitosan con- first derivative (DTG) curves obtained for the chitosan edible films. In
centration increases. An identical phenomenon was observed by general, chitosan edible films exhibit four weight-loss stages. The first
Fundo et al. [45]. A greater chitosan concentration results in an increase stage was presented between 40 °C and 74 °C, with WL values of 2 to
in the polymer/plasticizer ratio, which affects the free volume between 10%; this peak may correspond to the water adsorbed by the chitosan
the polymer chains. Consequently, there is a reduction in the level of in- edible films. According to de Morais Lima et al. [48], in the case of chitosan
termolecular forces between the polymers. The Tm values of the films edible films, water evaporation can occur up to a temperature of 120 °C.
decrease and approach the Tm value for pure chitosan (165 °C). This ex- The second stage is associated with the degradation of glycerol, which has
plains the above results. In contrast, the chitosan concentration exhib- a lower TD than the pure component. The TD of the chitosan edible films
ited a positive effect on the Δh; this result was expected since ranged between 151 °C and 190 °C, with WL values of 14% to 32%; pure
glycerol has a TD of 208 °C. The same behavior was observed during the
Table 2 third phase of weight loss. This transition indicates the polymer degrada-
Melting temperature (Tm) and melting enthalpy (Δh) of pure chitosan (powder) and chi- tion of the edible film, and ranged between 288 °C and 299 °C, with WL
tosan edible films. values of 40% to 54%. However, the pure chitosan had a TD of 302 °C. In
Sample Tm (°C) Δh these stages, the low TD values are associated with the dehydroxylation
of the organic components, which are connected by hydrogen bonds. It
CH05T02 181 ± 3.14 205 ± 9.51
CH10T02 177 ± 2.10 205 ± 7.65 is believed that the elimination of the adsorbed water can approximates
CH15T02 156 ± 12.5 267 ± 32.1 the chains of the organic components in the films. This induces the dehy-
CH05T25 159 ± 20.3 205 ± 30.3 droxylation process, thus lowering the transition temperature [36]. How-
CH10T25 162 ± 11.3 176 ± 33.2
ever, the fourth weight-loss stage is related to the decomposition of the
CH15T25 161 ± 15.2 151 ± 23.4
CH05T40 155 ± 14.9 199 ± 25.1 Tween 80®. The chitosan edible films showed a TD between 412 °C and
CH10T40 166 ± 24.2 189 ± 27.5 437 °C, with WL values ranging between 50% and 74%. In contrast, the
CH15T40 158 ± 23.5 236 ± 26.9 pure component showed a TD of 406 °C. According to Lima et al. [36],
Chitosan (powder) 165.9 156.4 this increase in the TD indicates the high level of interaction between
Values are expressed as means ± standard deviation (n = 3). the film components.
1238 A. Homez-Jara et al. / International Journal of Biological Macromolecules 113 (2018) 1233–1240

Fig. 2. DTG curves of chitosan edible films and pure chitosan (powder), Tween 80 and
glycerol.
Fig. 3. FTIR spectrums of chitosan edible films and pure chitosan (powder), Tween 80 and
glycerol.
3.6. Fourier-transform infrared spectroscopy

FTIR was used to observe the chemical interactions of the chitosan


edible films. Fig. 3 shows the FTIR spectra of the chitosan edible films. barrier, mechanical, and optical properties [49]. At a magnification of
The majority of the IR spectra of the chitosan edible films exhibited a 200×, all the chitosan films showed a clear, smooth, and uniform struc-
broad band between 3360 and 3406 cm−1, which can be attributed to ture. In addition, the presence of small, solid particles could be observed
the stretching of hydroxyl groups (\\O\\H) and amines (N\\H). How- at the surface of the film. However, when the magnification was in-
ever, in the same region, signals corresponding to the stretching of creased to 20,000×, several differences could be observed. The films
O\\H can be observed; therefore, it is possible that overlapping oc- dried at 2 °C appeared to be scaly with deep pores (Fig. 4), which can ex-
curred amongst the signals [48]. The IR spectra also showed two plain the high WVP values obtained from these films. However, the films
bands between 3000 and 2800 cm−1, corresponding to the asymmetri- dried at 25 °C presented a homogeneous surface, suggesting the pres-
cal and symmetrical stretching of C\\H bonds. However, the bands pre- ence of a cohesive matrix, which is owed to the polymer and glycerol in-
sented at 1638 cm−1, 1555 cm−1, and 1322 cm−1 are representative of teractions [36]. The films that were dried at 40 °C showed surface cracks
the chitosan spectra, which can be assigned to the C_O stretching and irregularities, which was owed to the lack of glycerol molecules
modes (amide I), N\\H deformation modes (amide II), and C\\N within the film matrix, suggesting the existence of a polymer–polymer
stretching modes (amide III), respectively [48]. The registered bands structure [50].
at 1407 cm−1 and 1377 cm−1 represent the deformation modes of
CH2 and CH3 bonds [36], respectively, whereas the band near 4. Conclusions
1240 cm−1 represents the vibrations of the C\\N and N\\H groups
[32]. The bands between 1150 and 1027 cm−1 can be assigned to the This study attempted to illustrate the properties of edible films
C\\O stretching modes of the linkages of the alcohol groups [32,48]. produced using conditions that food-packaging products are typically
subjected to on industrial or commercial scales. Thus, the drying temper-
atures were selected to represent refrigeration conditions (2 °C), the
3.7. Scanning electron microscopy average temperature of the environment (25 °C), and the typical temper-
ature used in conventional air-forced drying (40 °C). Thereby, we suc-
The morphologies of the edible films reflect the level of interaction cessfully produced chitosan edible films with the use of various
between the film components. This can be influenced by the drying polymer concentrations and drying temperatures. It can be concluded
temperature, which affects the physical properties of the resultant ma- that the lower drying temperatures had a positive effect on the properties
terial. Thus, the microstructure of the film can be associated with its of the films, namely the MC, S, WVP, and OP. This is associated with the
A. Homez-Jara et al. / International Journal of Biological Macromolecules 113 (2018) 1233–1240 1239

Fig. 4. SEM micrographs of the surface of chitosan edible films (2000×). A, CH05T02; B, CH10T02; C, CH15T02; D, CH05T25; E, CH10T25; F, CH15T25; G, CH05T40; H, CH10T40 and I,
CH15T40.

use of longer drying periods, which promotes the incorporation of water [4] A.B. de Aquino, A.F. Blank, L.C.L. de Aquino Santana, Impact of edible chitosan–
cassava starch coatings enriched with Lippia gracilis Schauer genotype mixtures
molecules within the chitosan-film structure. However, the use of greater on the shelf life of guavas (Psidium guajava L.) during storage at room temperature,
drying temperatures (40 °C), combined with a higher chitosan Food Chem. 171 (2015) 108–116.
concentration, enhanced the TS, SP, and b* values of the films but dimin- [5] N.E. Suyatma, et al., Effects of hydrophilic plasticizers on mechanical, thermal,
and surface properties of chitosan films, J. Agric. Food Chem. 53 (10) (2005)
ished their L* values. The EB of the films decreased as the drying temper- 3950–3957.
ature increased. The use of the lowest drying temperature (2 °C), in [6] G. Aguirre-Alvarez, et al., The effect of drying temperature on mechanical properties
combination with an increase in the chitosan concentration, resulted in of pig skin gelatin films, CyTA J. Food 9 (3) (2011) 243–249.
[7] W. Choi, J. Han, Film-forming mechanism and heat denaturation effects on the phys-
the achievement of greater L* and a* values; it also decreased the Tm of ical and chemical properties of pea-protein-isolate edible films, J. Food Sci. 67 (4)
the films. The chitosan films developed in this study showed many desir- (2002) 1399–1406.
able characteristics, such as their low WVP and EB values, and elevated TS [8] L. Palou, S.A. Valencia-Chamorro, M.B. Pérez-Gago, Antifungal edible coatings for
fresh citrus fruit: a review, Coatings 5 (4) (2015) 962–986.
and L* values. Therefore, they could be used as packaging or coating ma-
[9] M. Aider, Chitosan application for active bio-based films production and potential in
terials for food products in the future. the food industry: review, LWT- Food Sci. Technol. 43 (6) (2010) 837–842.
[10] T. Sun, et al., Preparation and preservation properties of the chitosan coatings mod-
ified with the in situ synthesized nano SiOx, Food Hydrocoll. 54 (2016) 130–138.
Acknowledgments [11] J.V. Tezotto-Uliana, et al., Chitosan applications pre-or postharvest prolong rasp-
berry shelf-life quality, Postharvest Biol. Technol. 91 (2014) 72–77.
[12] A.S.K. Gardesh, et al., Effect of nanochitosan based coating on climacteric behavior
The authors acknowledge financial support from the COLCIENCIAS and postharvest shelf-life extension of apple cv. Golab Kohanz, LWT- Food Sci.
(Colombia), through project number 2309-669-45371, and from the Technol. 70 (2016) 33–40.
Universidad del Tolima, through project number 750115. [13] M. Kaya, et al., Chitosan coating of red kiwifruit (Actinidia melanandra) for extending
of the shelf life, Int. J. Biol. Macromol. 85 (2016) 355–360.
[14] M. Duran, et al., Potential of antimicrobial active packaging ‘containing
References natamycin, nisin, pomegranate and grape seed extract in chitosan coating’
to extend shelf life of fresh strawberry, Food Bioprod. Process. 98 (2016)
[1] M. Spagnuolo, et al., Migration test of bisphenol A from polycarbonate cups using 354–363.
excitation-emission fluorescence data with parallel factor analysis, Talanta 167 [15] M.Z. Treviño-Garza, et al., Edible active coatings based on pectin, pullulan, and chi-
(2017) 367–378. tosan increase quality and shelf life of strawberries (Fragaria ananassa), J. Food Sci.
[2] M.Z. Elsabee, E.S. Abdou, Chitosan based edible films and coatings: a review, Mater. 80 (8) (2015) M1823–M1830.
Sci. Eng. C 33 (4) (2013) 1819–1841. [16] K. Liu, et al., Influence of postharvest citric acid and chitosan coating treatment on
[3] Y. Tan, et al., Functional chitosan-based grapefruit seed extract composite films for ripening attributes and expression of cell wall related genes in cherimoya (Annona
applications in food packaging technology, Mater. Res. Bull. 69 (2015) 142–146. cherimola mill.) fruit, Sci. Hortic. 198 (2016) 1–11.
1240 A. Homez-Jara et al. / International Journal of Biological Macromolecules 113 (2018) 1233–1240

[17] M. Petriccione, et al., Influence of a chitosan coating on the quality and nutraceutical [33] A.I. Bourbon, et al., Physico-chemical characterization of chitosan-based edible films
traits of loquat fruit during postharvest life, Sci. Hortic. 197 (2015) 287–296. incorporating bioactive compounds of different molecular weight, J. Food Eng. 106
[18] S.S. Shekarforoush, et al., Effect of chitosan on spoilage bacteria, Escherichia coli and (2) (2011) 111–118.
Listeria monocytogenes in cured chicken meat, Int. J. Biol. Macromol. 76 (2015) [34] M. Pereda, et al., Chitosan-gelatin composites and bi-layer films with potential anti-
303–309. microbial activity, Food Hydrocoll. 25 (5) (2011) 1372–1381.
[19] G. Vasilatos, I. Savvaidis, Chitosan or rosemary oil treatments, singly or combined to [35] Y. Shahbazi, The properties of chitosan and gelatin films incorporated with ethanolic
increase Turkey meat shelf-life, Int. J. Food Microbiol. 166 (1) (2013) 54–58. red grape seed extract and Ziziphora clinopodioides essential oil as biodegradable
[20] D. Dehnad, et al., Thermal and antimicrobial properties of chitosan–nanocellulose materials for active food packaging, Int. J. Biol. Macromol. 99 (2017) 746–753.
films for extending shelf life of ground meat, Carbohydr. Polym. 109 (2014) [36] M. Lima, et al., Structural, thermal, physical, mechanical, and barrier properties of
148–154. chitosan films with the addition of xanthan gum, J. Food Sci. 82 (3) (2017) 698–705.
[21] M. Vargas, A. Albors, A. Chiralt, Application of chitosan-sunflower oil edible films to [37] M.A. Cerqueira, et al., Effect of glycerol and corn oil on physicochemical properties of
pork meat hamburgers, Procedia Food Sci. 1 (2011) 39–43. polysaccharide films–a comparative study, Food Hydrocoll. 27 (1) (2012) 175–184.
[22] S. Vimaladevi, et al., Packaging performance of organic acid incorporated [38] L. Sperling, Introduction to Polymer Physical Science, John Wiley & Sons, New Jersey,
chitosan films on dried anchovy (Stolephorus indicus), Carbohydr. Polym. 2006.
127 (2015) 189–194. [39] R.Y. Aguirre-Loredo, et al., Effect of equilibrium moisture content on barrier, me-
[23] M.J. Costa, et al., Use of wheat bran arabinoxylans in chitosan-based films: effect on chanical and thermal properties of chitosan films, Food Chem. 196 (2016) 560–566.
physicochemical properties, Ind. Crop. Prod. 66 (2015) 305–311. [40] G. Kerch, V. Korkhov, Effect of storage time and temperature on structure, mechan-
[24] R.D. Baron, et al., Production and characterization of films based on blends of chito- ical and barrier properties of chitosan-based films, Eur. Food Res. Technol. 232 (1)
san from blue crab (Callinectes sapidus) waste and pectin from Orange (Citrus (2011) 17–22.
sinensis Osbeck) peel, Int. J. Biol. Macromol. 98 (2017) 676–683. [41] A. Nawab, et al., Mango kernel starch-gum composite films: physical, mechanical
[25] Materials, A.S.f.T.a, Standard Test Method for Tensile Properties of Thin Plastic and barrier properties, Int. J. Biol. Macromol. 98 (2017) 869–876.
Sheeting, ASTM International, 2010. [42] B. Saberi, et al., Optimization of physical and optical properties of biodegradable ed-
[26] A. Standard, E96-00. Standard test methods for water vapour transmission of mate- ible films based on pea starch and guar gum, Ind. Crop. Prod. 86 (2016) 342–352.
rials, Annual Book of ASTM Standards 2000, p. 4. [43] M.A. López-Mata, et al., Physicochemical, antimicrobial and antioxidant properties
[27] D. Hu, H. Wang, L. Wang, Physical properties and antibacterial activity of of chitosan films incorporated with carvacrol, Molecules 18 (11) (2013)
quaternized chitosan/carboxymethyl cellulose blend films, LWT- Food Sci. Technol. 13735–13753.
65 (2016) 398–405. [44] Z.-H. Zhang, et al., Enhancing mechanical properties of chitosan films via modifica-
[28] A.E.C. Fai, et al., Development and evaluation of biodegradable films and coatings tion with vanillin, Int. J. Biol. Macromol. 81 (2015) 638–643.
obtained from fruit and vegetable residues applied to fresh-cut carrot (Daucus [45] M. Pereda, G. Amica, N.E. Marcovich, Development and characterization of edible
carota L.), Postharvest Biol. Technol. 112 (2016) 194–204. chitosan/olive oil emulsion films, Carbohydr. Polym. 87 (2) (2012) 1318–1325.
[29] J.R. Rodríguez-Núñez, et al., Chitosan/hydrophilic plasticizer-based films: prepara- [46] Y. Liu, et al., Molecular interactions, characterization and antimicrobial activity of
tion, physicochemical and antimicrobial properties, J. Polym. Environ. 22 (1) curcumin–chitosan blend films, Food Hydrocoll. 52 (2016) 564–572.
(2014) 41–51. [47] J.F. Fundo, et al., The effect of polymer/plasticiser ratio in film forming solutions on
[30] V. Rizzi, et al., Molecular interactions, characterization and photoactivity of chloro- the properties of chitosan films, Food Biophys. 10 (3) (2015) 324–333.
phyll a/chitosan/2-HP-β-cyclodextrin composite films as functional and active sur- [48] M. de Morais Lima, et al., Biodegradable films based on chitosan, xanthan gum, and
faces for ROS production, Food Hydrocoll. 58 (2016) 98–112. fish protein hydrolysate, J. Appl. Polym. Sci. 134 (23) (2017).
[31] C. Liu, et al., Characterization of edible corn starch nanocomposite films: the effect of [49] S. Acosta, et al., Antifungal films based on starch-gelatin blend, containing essential
self-assembled starch nanoparticles, Starch-Starke 68 (2015) 239–248. oils, Food Hydrocoll. 61 (2016) 233–240.
[32] D. Kowalczyk, et al., Characterization of films based on chitosan lactate and its [50] M. Shahbazi, et al., Kinetic study of κ-carrageenan degradation and its impact on
blends with oxidized starch and gelatin, Int. J. Biol. Macromol. 77 (2015) mechanical and structural properties of chitosan/κ-carrageenan film, Carbohydr.
350–359. Polym. 142 (2016) 167–176.

View publication stats

You might also like