You are on page 1of 12

European Polymer Journal 89 (2017) 162–173

Contents lists available at ScienceDirect

European Polymer Journal


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

Incorporation of tetraethylorthosilicate (TEOS) in biodegradable


films based on bean starch (Phaseolus vulgaris)
MARK

Karina Oliveira Limaa, , Bárbara Biduskib, Wyller Max Ferreira da Silvab,
Silvia Moreira Ferreiraa, Lara Machado Pereira Montenegroa,
Alvaro Renato Guerra Diasb, Daniela Bianchinia
a
Centro de Ciências Químicas, Farmacêuticas e de Alimentos, Universidade Federal de Pelotas, 96010-900 Pelotas, Brazil
b
Departamento de Ciência e Tecnologia Agroindustrial, Universidade Federal de Pelotas, 96010-900 Pelotas, Brazil

AR TI CLE I NF O AB S T R A CT

Keywords: The starch has been investigated to obtain biodegradable films, but has disadvantages such as
Bean starch hydrophilicity and poor mechanical properties. The addition of inorganic materials can improve
TEOS mechanical properties by the synergism of the components. The aim of this study was to obtain
Hybrid films hybrid films of bean starch and different tetraethylorthosilicate (TEOS) content (5, 20 and
Sol-gel method
40 g⋅100 g−1 starch) as well as to evaluate the influence in the properties of films. The films were
prepared by casting method with hydrolysis and condensation of TEOS in situ under a controlled
pH. The films were characterized after 3 and 15 days of storage, presenting a higher value of
Young’s modulus for the film containing 40 g TEOS, lower melting temperature, more
hydrophobic surface, and a higher resistance to degradation than starch film. Furthermore,
barrier properties improved when stored for 15 days. The relative crystallinity was not affected
by adding the precursor.

1. Introduction

The effect of pollution generated by the disposal and accumulation of petroleum based plastic on the environment has encouraged
research to develop biodegradable packaging obtained from renewable sources [1]. Starch has been investigated for the preparation
of biodegradable films, due to its low cost, wide availability, and biodegradability [2,1]. Starch is a semicrystalline material
consisting of two macromolecules, amylose which consists mainly of a straight chain forming an amorphous region in the structure,
and amylopectin which gives a crystalline region [3].
The application of starch in the film production is due to the ability of amylose to form a strong and stable network in a solution,
with the strong orientation of the chains through hydrogen bonds [4]. The starch bean has 24–65% amylose content, it can vary
according to the botanical origin and cultivar. This way, due the high amylose content in bean starch, this grain becomes interesting
for developing film [5]. Carioca bean (Phaseolus vulgaris L.) is a Brazilian grain, widely consumed in Brazil, characterized by a light
brown tegument with brown stripes [6]. The bean is a legume which has been known as healthy food due to its high contents of
protein, dietary fiber as well as slowly digestible. Because of this, different researchers have studied the physicochemical and
processing properties as well as the digestibility of the grains [7,8] and bean starch [9–11]. However, the starch of legumes is not as


Corresponding author.
E-mail addresses: karinah_ol@hotmail.com (K.O. Lima), babi_biduski@hotmail.com (B. Biduski), wyller_m@hotmail.com (W.M.F.d. Silva),
silviaferreira92@gmail.com (S.M. Ferreira), mp_lara@yahoo.com.br (L.M.P. Montenegro), alvaro.guerradias@gmail.com (A.R.G. Dias),
danielabianchini.ufpel@gmail.com (D. Bianchini).

http://dx.doi.org/10.1016/j.eurpolymj.2017.02.008
Received 27 September 2016; Received in revised form 30 January 2017; Accepted 6 February 2017
Available online 08 February 2017
0014-3057/ © 2017 Elsevier Ltd. All rights reserved.
K.O. Lima et al. European Polymer Journal 89 (2017) 162–173

thoroughly studied as the cereals and tubers starches. Therefore, legume starches need to be better studied to facilitate the
developments of new applications in both food and non-food products [12,9].
Starch films may be obtained after a disruption of the granular structure of starch by the gelatinization process, and addition of a
plasticizer. The application of native starch for developing films is limited due to its hydrophilic character, which results in a film
with both poor mechanical properties and a water barrier when compared to synthetic polymers [13–17]. Thus, the use of additives,
such as glycerol (plasticizer), cellulose, and chitosan have been reported to improve the properties of the starch films [18].
Currently there is great interest in the synthesis of hybrid films, consisting of the combination of organic and inorganic
components in order to obtain specific properties by the synergism of the components [19], for instance, the association of the
flexibility of organic compounds with improved mechanical properties of an inorganic component. Thus, this enables the formation
of a hybrid network. The most common method used for the preparation of hybrid materials is the sol-gel process, which allows
control of the chemical composition at mild temperatures by hydrolysis and condensation processes [20].
The precursors usually employed as an inorganic component for the sol-gel process are organoalkoxysilane [20]; among these we
highlight tetraethylorthosilicate (TEOS), which may form small particles of silica with a high specific surface area where there are
unsaturated chemical bonds. The hydroxyl groups on the surface facilitate the dispersion of macromolecular chains in hybrid films
[21]. Some studies about a starch-inorganic hybrid materials have been carried out [21–23]. However, relevant studies on water
vapor permeability, the mechanical, and thermal characterization of hybrid films, containing only starch and TEOS, and synthesized
by sol-gel method, have not been reported in the literature. To our knowledge, this is the first report showing a complete
characterization and study of the storage time effect. The aim of this study was to investigate the incorporation of TEOS in
biodegradable films based on carioca bean starch on the structural, morphological, thermal, opacity, surface, barrier, and mechanical
properties.

2. Materials and methods

2.1. Materials

Carioca bean grains (Phaseolus vulgaris L.), which were purchased from a local market (Pelotas, RS), were used for starch isolation.
Tetraethylorthosilicate 98% (Si(OC2H5)4) was obtained from Sigma-Aldrich (131903); and all the chemical reagents used in this work
had an analytical grade.

2.2. Isolation of starch

The carioca bean starch was isolated according to Rupollo et al. [10], with some modifications, such as the centrifugation of the
starch layer suspended in distilled water at 4000 rpm for 10 min. The brownish top layer was discarded; the underlayer was
resuspended in distilled water and recentrifuged at 4000 rpm for 10 min. This process was repeated once more. The resulting material
was dried in an oven with air circulation at 40 °C for 16 h until the moisture content was around 9%. The dried starch was milled
using basic analytical mill (IKA, A11, USA) and stored at 17 ± 2 °C in a sealed recipient. The starch had approximately 99% purity
(0.35% protein, 0.17% lipid and 0.055% ash).

2.3. Characterization of starch

Analysis of the amylose content, morphology, X-ray diffraction pattern, thermal analysis (gelatinization) and nature of the surface
of the carioca bean starch was performed such as describe in item 2.4, 2.5, 2.6, 2.7, 2.8, respectively.

2.4. Amylose

The starch amylose content was determined by a colorimetric method with iodine according to the method described by McGrane;
Rix and Cornell [24]. A suspension with 20 mg of defatted carioca bean starch (d.b.) and 8 mL of dimethylsulfoxide (DMSO) at 90%
was stirred and then heated in bath at 85 °C for 2 h. After cooling, the samples were transferred to a volumetric flask of 25 mL,
homogenized and adjusted volume. An aliquot of 1 mL of the solution was added 5 mL of I2/KI (0.0025/0.0065 mol/L) solution and
the volume was completed to 50 mL. The resulting solution was homogenized and allowed to stand for 15 min prior to reading
absorbance at 600 nm. To perform the standard amylose curve was used 20 mg of pure potato amylose (Sigma) and realized the same
procedure described previous, using aliquots of 0; 0.1; 0.2; 0.4; 0.6; 0.8 and 1.0 mL to determine the absorbance in a
spectrophotometer (Jenway, 6705, UK) and construction of the standard curve.

2.5. Scanning electron microscopy (SEM)

The starch morphology was visualized by using a microscope (JEOL, JSM 6610LV, USA). It focused on the sample with an electron
beam, with an accelerating voltage of 15 kV. The samples were deposited on copper tape at the aluminum stub. The sample was
covered with an Au layer using a Denton Vacuum (Desk V, USA), with the sputtering method for 110 s and 20 mA. The starch images
were performed with a magnification of 500×.

163
K.O. Lima et al. European Polymer Journal 89 (2017) 162–173

2.6. X-ray diffraction (XRD)

The structural properties of the starch granule and films were determined in the X-ray diffractometer (Advance Brukers, D8,
Germany) using Cu Kα (λ = 1.5418 Å) as their source. The diffraction angle (2θ) ranged between 5 and 50°, with a target voltage of
40 kV, a current of 40 mA, a scan increment of 0.0205, and a rate of 1° min−1. The relative crystallinity (RC) was calculated as
described by Rabek [25] by the equation RC (%) = (Ac/(Aa + Ac) ∗ 100), where Ac is the area of crystalline phase, and Aa is the
area of the amorphous phase in the X-ray diffractogram.

2.7. Differential scanning calorimetry (DSC)

The gelatinization characteristics of starch were determined by differential scanning calorimetry (TA-60WS, Shimadzu, Japan)
according to the method described by Vanier et al. [11]. Approximately 2.5 mg of starch were weighed in an aluminum pan to which
was added distilled water (1:3 w/w). After that the pan was hermetically sealed and allowed to stabilize for 24 h before analysis. The
sample was analyzed from 30 to 120 °C, with a heating rate of 10 °C·min−1 in a nitrogen atmosphere. An empty pan was used as a
reference. The glass transition temperature, melting temperature, and the enthalpy change were determined for the films.
Approximately 10 mg of film were weighed, the sample were analyzed from −100 to 250 °C.

2.8. Attenuated total reflectance Fourier transform infrared spectroscopy (ATR-FTIR)

The nature of the starch and films surface were analyzed by using a Shimadzu spectrometer (IRAffinity-1, Japan), coupled to an
Attenuated Total Reflectance accessory (ATR) (Pike Tech, Madison, WI.). The samples were placed on the zinc selenide crystal (ZnSe)
and the analysis were performed in the range of 4000–500 cm−1, co-adding 32 scans and spectral resolution of 4 cm−1.

2.9. Synthesis of the films

Films were prepared by casting method according to Biduski et al. [26], with some modifications. The solutions were prepared
with the following proportion: 3.0 g starch:0.3 g glycerol:100 mL distilled water. The solutions were stirred at 300 rpm for 10 min at
90 °C using a thermostatized bath (Quimis, Q214M2). After this time, the filmogenic solution was cooled to 40 °C, followed by a pH
adjustment to 9.0 and a slow addition of TEOS at a concentration of 0.0 g, 5.0 g, 20.0 g and 40.0 g TEOS.100 g−1 starch. The
temperature was maintained at 40 °C for 1 h and then the solution was homogenized in an ultraturrax (IKA, T18B, Werke, Germany)
at 15.500 rpm for 10 min at room temperature. Afterwards, the solution was heated to 40 °C for 1 additional hour. Lastly, 20 g of the
solution were spread on an acrylic plate and dried in an oven with forced air circulation (Ethik, Brazil) for 20 h at 30 °C. Before
characterization, the films were stored according ASTM D882-12 [27] for 3 and 15 days in a desiccators at 25 °C ± 3 °C with a
relative humidity (RH) of 50% ± 3 using a saturated solution of magnesium nitrate hexahydrate (133, VETEC). The RH and
temperature were controlled using thermohygrometer. The films were named as ST0, ST5, ST20 and ST40, according to the TEOS
content.

2.10. Scanning electron microscopy with energy dispersive X-ray spectroscopy (SEM-EDS)

The surface and cross-section morphology of the films were visualized by a scanning electron microscope (SEM) (JEOL, JSM-
6610LV, USA). For the surface analysis, the samples were adhered to the surface of a double sided copper tape at the stub; and for the
analysis of the cross section, the films were fractured under liquid N2 and adhered to the colorless glaze at the stub. The films were
covered with a gold layer and analyzed by a SEM with an accelerating voltage of 10 kV. The micrographs of the surface and cross
section of the films were performed using 1000× and 500× magnification, respectively. An elemental analysis by EDS was
performed at different positions on the cross section film.

2.11. Thermogravimetric analysis (TGA)

The degradation temperature of the samples was obtained using a thermogravimetric analyzer (TA-60WS, Shimadzu, Kyoto,
Japan). Samples (4–6 mg) were heated between 30 and 550 °C at a heating rate of 10 °C.min−1 and a nitrogen flow of 50 mL·min−1.

2.12. Opacity of the films

The opacity of the films were obtained using a Minolta colorimeter (CR, 400, Japan), and was calculated as the ratio between the
value obtained from films overlapping the black standard and the value obtained from the films overlapping the white standard,
multiplied by 100 [28].

2.13. Thickness, water solubility and water vapor permeability (WVP) of the films

The film thickness was obtained by a digital micrometer (INSIZE, IP-54), according to the ASTM method F2251-13 [29], by the
average of eight measurements in random positions for each film.

164
K.O. Lima et al. European Polymer Journal 89 (2017) 162–173

The water solubility was calculated as the percentage of dry matter of the film solubilized after immersion in water at 25 °C for
24 h according to Gontard et al. [30].
The WVP was performed according to the method ASTM E-96/E96M-14 [31] at 25 °C. The samples were sealed with paraffin films
on aluminum permeation cells (PVA-4, REGMED, Brazil) containing calcium chloride (0% RH). Permeation cells were stored in
vacuum desiccators with saturated sodium chloride solution 75% RH. The mass gain of the system was measured for 7 days.

2.14. Mechanical properties of the films

The tensile strength, elongation and Young’s modulus of the films were evaluated in a texturometer (TA.TXplus Texture
Analyzer), according to ASTM D882-12 [27]. Eight film samples (80 mm × 25 mm) from each film were evaluated. The tensile
strength was determined by the maximum stress point from a stress-strain curve. The elongation (strain-specific) was measured as the
difference between the distance run until the film ruptured and the initial separation distance (40 mm), multiplied by 100. The
Young's modulus was calculated from the inclination of the initial linear region of the stress-strain curve.

2.15. Contact angles

Contact angle films were measured using a Drop Shape Analyzer (Kruss) and software- DSA4. The sessile drop method was used
with deionized water; a 3 µL droplet of deionized water was gently deposited on the film using a microsyringe and observed with a
digital microscope. The values reported are averages of three measurements performed in different areas of sample surface.

2.16. Statistical analysis

The analytical determination of the samples was performed in triplicate. The average values and standard errors were
demonstrated for each result, except for the relative crystallinity and thermal analysis of the starch and films. At first, the results were
evaluated by a Q test for rejection of data. After that, the results were compared by Tukey test at a 5% level of significance by an
analysis of variance (ANOVA) or a t test to compare the two means, when necessary.

3. Results and discussion

3.1. Characterization of starch

The carioca bean starch showed 34.1% of amylose. According to Singh [5], bean starch presents a high content of amylose,
between 24 and 65%. Therefore, due to its amylose content, the bean starch is an interesting carbohydrate for the development of
films. Amylose is a biopolymer with linear chains and when in a solution, it tends to form a strong and stable network through
hydrogen bonds between the hydroxyl groups [4].
Fig. 1A shows the morphology of the starch granules of the carioca bean obtained by SEM. The granules have oval and spherical
morphologies with the size range between 12 and 36 µm, as well as smooth surfaces without cracks, as reported by Vanier et al.[11].
According to XRD pattern, starches can be classified as A-type (cereals), B-type (tubers), or the mix of both, C-type (leguminous
and some seeds). Carioca bean starch presents a C-type diffraction pattern, which is very common in legume plants. Peaks observed in
the diffraction angles (2θ) 5.6, 15, 17 and 23° (Fig. 1B) are typical for this type of starch. Similar peaks were observed in the study
carried out by Rupollo et al. [10] with carioca bean starch stored in different atmospheric conditions. The carioca bean starch used in
this study had 34.0% of RC.
The characterization of the carioca bean starch by DSC showed that the onset temperature, the peak temperature and the endset
temperature of gelatinization were 66.9, 73.7 and 81.2 °C, respectively. Besides, the enthalpy change of gelatinization of 9.9 J·g−1
was calculated.
Fig. 2 shows the ATR-FTIR spectra of bean starch (Fig. 2A). According to the results shown in Fig. 2A, the spectrum of bean starch
has a broad band centered at 3300 cm−1, which is assigned to OeH stretching modes. Bands between 3000 and 2800 cm−1 are
assigned to CeH antisymmetric and symmetric stretching modes of methylene groups. Due to the hydrophilic character of the bean
starch, the spectrum also shows a band at 1640 cm−1, which is assigned to HeOeH deformation modes of water adsorbed on the
surface. The highlighted region between 1160 and 854 cm−1 is assigned to CeOeC vibrational modes of glycosidic bonds of starch
[32].

3.2. Characterization of films

3.2.1. Attenuated total reflectance Fourier transform infrared spectroscopy (ATR-FTIR)


Fig. 2 shows the ATR-FTIR spectra of the films of bean starch with different TEOS content (Fig. 2B). The spectra of bean starch
film (Fig. 2B a) show the same band profile as the spectrum of native bean starch (Fig. 2A). However, the spectra of the films
containing the inorganic precursor (Fig. 2B b–d) show an enlargement of the bands between 1180 and 850 cm−1, compared to film
ST0, mainly in film ST40 (Fig. 2B d). This suggests the presence of new vibrational modes, asymmetric and symmetric SieOeSi
stretching modes of silica obtained from the hydrolysis and condensation of TEOS. He et al. [33] reported similar results. More
specifically, 1100–1000, and 950–900 cm−1 was attributable respectively, to ν(SieOeSi) and ν(SieOH) [19].

165
K.O. Lima et al. European Polymer Journal 89 (2017) 162–173

Fig. 1. Image of carioca bean starch granules (500×) (A), X-ray diffractogram of bean starch (B) and films (C) ST0 (a), ST5 (b), ST20 (c) and ST40 (d) with highlighted
peaks.

During the reaction and aging time is expect to occur the condensation of the biopolymer matrix and the inorganic precursor, the
broad band assigned to OeH stretching modes between 3500 and 3000 cm−1 suggests a large extension of hydrogen bonds between
the components of the film [32].

Fig. 2. ATR-FTIR spectra of bean starch (A) and films (B) ST0 (a), ST5 (b), ST20 (c) and ST40 (d).

166
K.O. Lima et al. European Polymer Journal 89 (2017) 162–173

Fig. 3. Images of the surface (A, C, E, G) and the cross section (B, D, F, H) of the ST0 (A, B), ST5 (C, D), ST20 (E, F) and ST40 (G, H) films.

3.2.2. X-ray diffraction (XRD)


The relative crystallinity of films seems not to be affected by the addition of the precursor TEOS, presenting RC of 10.9, 10.4, 10.0
and 10.5% for ST0, ST5 ST20 and ST40, respectively. It should be noted that RC of native bean starch is 34.0% indicating that the
initial crystalline structure was destroyed in the films, after adding glycerol. This new structure was formed during the cooling by

167
K.O. Lima et al. European Polymer Journal 89 (2017) 162–173

Table 1
Distribution of Si in the films with different TEOS content.

Films

ST0 ST5 ST20 ST40

Upper border Si (%) – 48.14 76.14 97.65


Center – 55.77 52.80 89.11
Bottom border – 55.78 71.72 97.24

recrystallization of amylose in single helices involving glycerol [34]. Peaks in the same 2θ angles, 5.6, 15, 17, 19.5, 22, and 24°, were
observed for all films (Fig. 1C). It should be noted that the low ordering degree in the starch films (around 10%) hinders observation
of any difference in the crystallinity of the films by XRD analysis.
The formation of stable crystalline regions in polymeric materials requires that an economical close packed arrangement of the
chains can be achieved in three dimensions, imposes restrictions on the type of chain that can crystallize most readily. Linear and
symmetrical chains allow the formation of regular close packing. In addition, chains possessing groups that encourage strong
intermolecular attraction thereby stabilizing the alignment of chains [35]. Starch chains have hydroxyl groups that can interact by
hydrogen bonds each other and stabilize the structure of biopolymer. It should be reminded that starch is formed by both linear and
branched chains, which results in polymeric material with lower degree of crystallinity. The crystallinity of starch films is lower than
that observed in the native starch due the addition of components with lower molecular weight, such as glycerol. Glycerol also
interact by hydrogen bonds and can destroy the interactions among starch chains.

3.2.3. Scanning electron microscopy with energy dispersive X-ray spectroscopy (SEM-EDS)
Initially, through a visual subjective evaluation all bean starch films had a homogeneous appearance, without blisters. The side of the
film in contact with the Petri plate presented a higher brilliance in all cases. The global appearance of the ST5 and ST20 films was not
affected by the addition of TEOS. On the other hand, the ST40 film was the most opaque, with different degree of opacity on the same film.
The images of the surface and the cross section of films prepared with different TEOS content are shown in Fig. 3 (A, C, E, G) and
(B, D, F, H), respectively. The numbers inside the images of the cross section indicates the region where EDS analyses were
performed. The surface of films containing the inorganic precursor TEOS is more compact than that observed for the film synthesized
without TEOS. Xiong et al. [21] have studied the addition of nano-SiO2 in films of corn starch and polyvinyl acetate. The authors have
reported that the addition of nano-SiO2 results in films with smooth and compact surface, suggesting that an increase in the
miscibility and compatibility of components had occurred in the film. In addition, the authors explained that nanoparticles of silica
have unsaturated bonds on the surface, which favors the hydrogen bonds of silica with hydroxyl groups of starch chains and oxygen
atoms of vinyl polyacetate, as well as the formation of SieOeC bonds. This may explain the compact morphology of the films in this
study, suggesting the presence of interactions between starch and an inorganic precursor.
The EDS analysis, Fig. 3(B, D, F and H), obtained a semi quantitative analysis of different points in the cross section of the films. If
we consider that the organic components of films are richer in C and O atoms, while the inorganic component is richer in Si and O
atoms, the EDS analysis can show the distribution of these components in the film. The EDS analyses showed only C and O atoms for
the film ST0. On the other hand, the films containing TEOS showed C, O and Si atoms distributed throughout the film.
The EDS analyses (Table 1) showed that the distribution of Si was not homogeneous for the films with more than 5% of TEOS.
Whereas the other films was observed the Si content is higher in the border than the center for films containing higher TEOS content.
The SEM images of surface, Fig. 3(E and G) showed some concentration of white dots on the surface, which is not observed in the ST5
film (Fig. 3C). These results suggest the segregation of inorganic material in the border of films due to an excess of the inorganic
precursor. A probably saturating SieOH bonds between starch and silica could be occurred, decreasing the intensity band around
900 cm−1 as observed in the FTIR analysis.

3.2.4. Differential scanning calorimetry (DSC)


The melting enthalpy change indicates how much heat is needed in the melting process. Results in Table 2 show that the films
containing TEOS have an enthalpy change (ΔmH) higher than that obtained for the ST0 film. In this work, the increase of the TEOS
content in filmogenic solutions resulted in a decrease of the onset temperature, the peak temperature, and the endset temperature of
melting of the films (Table 2). The endothermic peak is associated with the melting of the crystalline phase of starch, which is
reorganized during the retrogradation process in the synthesis of film [36]. A broadening of peak (Table 2, ΔTm) was also observed
when the TEOS content was increased. The displacement of the peak to lower temperatures, associated with the enlargement of the
peak, indicate heterogeneity in the crystalline phase of starch dispersed in the polymeric matrix. In addition, the melting of the
crystalline phase occurs in a wide temperature range. It should be noted that the displacement of the melting temperature to lower
values and the broadening of the peak are related to the organization of the chains, and consequently, to the size and crystallinity of
crystalline phase.
The addition of the inorganic precursor seems to affect the organization of chains. It is worth mentioning that the thermogram of
the ST0 film showed an abrupt melting at the beginning, with the narrowest peak, suggesting that the crystalline phase is more
homogeneous in this film.
The Tg is related to the amorphous phase present in the material and indicates a reduction in the movement of the molecular

168
K.O. Lima et al. European Polymer Journal 89 (2017) 162–173

Table 2
Thermal properties of bean starch films with different TEOS content.

TEOS content (%) Tm onset (°C) Tm peak (°C) Tm endset (°C) ΔTm (endset- onset) (oC) ΔmH (J·g−1) Tg (°C)

0 165.5 169.0 185.0 19.5 176.3 −30


5 153.3 163.7 183.1 29.8 249.6 15–36
20 113.4 139.0 163.5 50.0 268.4 20–23
40 117.1 133.0 159.4 42.3 223.5 7–20

Tm onset: initial crystalline melting temperature; Tm peak: peak crystalline melting temperature; Tm endset: final crystalline melting temperature; ΔTm: temperature
variation; ΔmH: enthalpy change; Tg: glass transition temperature.

chains with decreasing the temperature analysis. Polymeric materials show a glassy behavior below the Tg. On the other hand, the
chains of the polymeric material are flexible above the Tg [37]. According to Frone, Nicolae, Gabor, and Panaitescu [38], certain
factors can influence the increase of the characteristic glass transition temperature and melting temperature, such as crosslinking,
interactions between chains, and the release of water. The increase in Tg with the addition of TEOS (Table 2, Tg) may be associated
with reduced mobility of the polymer chains, which may suggest a strong interaction between the starch chains and the precursor
through inter- and intramolecular hydrogen bonds, which gives a greater rigidity for these films. Jia et al. [19] found similar results
when studied preparation and properties of poly(vinyl alcohol)/silica nanocomposites derived from copolymerization of vinyl silica
nanoparticles and vinyl acetate.

3.2.5. Thermogravimetric analysis (TGA)


The thermal analysis curves shown in Fig. 4 present an initial loss of weight up 120 °C, which is related to the loss of adsorbed
water molecules by the film due to the hydrophilic character. The films ST0, ST5 to ST20 have a similar profile between 120 and
170 °C, the presence of the plateau. This is almost not observed in the film ST40, showing a continuous profile of weight loss. The
high silica content in the film, may indicate the slow dehydroxylation of this component during the analysis and possibly a strong
interaction with the organic components. Thus, it shows a higher initial resistance to degradation by temperature, as well as a slow
and continuous loss. This profile is followed by a weight loss between 170 and 290 °C, which is probably due to the dehydroxylation
of the film components associated by hydrogen bonds. According to Lawal et al. [39], the decomposition of glycerol occurs in the
range of 120 and 300 °C, which can be occurring in this work.
The intense weight loss in only one stage between 290 and 350 °C corresponds to the thermal decomposition of organic matter of
the film components. According to Guinesi et al. [40] in this temperature range occur the degradation of starch in non-oxidative
conditions, giving mainly CO2, CO, water, acetaldehyde, furan. The ST5, ST20 and ST40 films exhibited a weight loss slightly lower
than ST0 film, which is in agreement with the increasing levels of inorganic material added to the films.

3.2.6. Thickness, opacity and mechanical properties of the films


The addition of TEOS results in a slight increase in the thickness of the films due to the presence of small particles of silica in the
films (Table 3). The ST40 film, stored for 15 days, showed a higher opacity compared to other films (Table 3). However, this property
was not affected during the storage time of 3 days, which may be related to the aging process [16] as condensation reactions of the
TEOS that occur during storage [41]. In this storage step condensation reactions between starch and silica species could also occur,
giving rise to SieOeC bonds. However, it is not possible to conclude that this type of chemical bond is present in our films.
On the other hand, the opacity can be attributed to a phase separation of the organic and inorganic constituents as observed in the
EDS analysis, since optical transparency is a first criterion for the formation of a homogeneous phase [42]. In addition it can be
attributed the greater thickness presented, because higher thickness values result in more opaque samples [43] or the insertion of the
silica particles in the intermediate spaces of the starch film, preventing the transmission [13].

Fig. 4. Decomposition temperature curves of the ST0 (a), ST5 (b), ST20 (c), and ST40 (d) films.

169
K.O. Lima et al. European Polymer Journal 89 (2017) 162–173

Table 3
Color, opacity and mechanical properties of bean starch films with different TEOS content, stored for 3 and 15 days.

Parameters TEOS content (%) Storage time (days)

3 15

b ns
Thickness (mm) 0 0.103 ± 0.00 0.099 ± 0.00bc
5 0.106 ± 0.00b* 0.097 ± 0.00c
20 0.107 ± 0.00b* 0.104 ± 0.00b
40 0.114 ± 0.00a ns 0.118 ± 0.00a

Opacity 0 10.13 ± 0.32a ns 9.98 ± 0.28b


5 10.17 ± 0.57a ns 9.33 ± 0.50b
20 10.43 ± 0.67a ns 9.46 ± 0.47b
40 10.49 ± 0.27a* 11.81 ± 0.57ª

Water solubility (%) 0 14.35 ± 1.07b* 13.35 ± 1.15ª


5 15.53 ± 0.97ab ns 14.70 ± 1.05ª
20 16.03 ± 0.11ab* 13.83 ± 0.06ª
40 16.53 ± 0.38a* 14.35 ± 0.20ª

Tensile strength (MPa) 0 3.55 ± 0.31b* 5.03 ± 0.35b


5 3.46 ± 0.48b* 4.73 ± 0.40b
20 3.39 ± 0.59b* 4.80 ± 0.28b
40 6.66 ± 0.31ans 6.38 ± 0.31a

Young’s modulus (MPa) 0 48.13 ± 6.85b* 90.29 ± 6.85b


5 42.77 ± 13.06b* 64.57 ± 13.06b
20 45.84 ± 12.16b* 70.87 ± 12.16b
40 226.14 ± 43.70ans 226.76 ± 46.31a

Elongantion (%) 0 36.70 ± 0.06ans 40.45 ± 0.08a


5 43.47 ± 0.07a* 32.41 ± 0.05a
20 35.90 ± 0.14a* 31.83 ± 0.14ab
40 30.31 ± 0.08a* 20.58 ± 0.07b

Different lowercase letters in the same column (for each parameter) represent significant differences between the averages submitted to the Tukey test (p < 0.05). *
and ns are significant and not significant, respectively, for values in the same line submitted to the t test at 5% probability of error.

Table 3 shows a gradual increase of the water solubility at 3 days of storage as the TEOS content is increased, presumably by
forming hydrophilic species, such as unsaturated silica particles and small organic molecules containing hydroxyl groups, which are
leached from the films during solubility analysis. The water solubility decreased at 15 days of storage, probably due to an improved
interaction between the components of the film during the aging process, which convert silanol groups (SieOH) of silica and
alcoholic groups (CeOH) of starch and glycerol in less available species for interactions with water, such as SieOeSi and SieOeC.
Besides, strong hydrogen bonds between all components of the film could act as crosslinking points between chains and difficult the
solubility of films.

3.2.7. Mechanical properties of the films


The tensile strength, the elongation percentage, and the Young's modulus of the films with different TEOS content and storage
times are shown in Table 3. The higher tensile strength and Young's modulus, which are related to the rigidity of the films, were
observed for the ST40 film stored for both days (3 and 15) when compared to other TEOS content (ST0, ST5 and ST20), probably due
to its greater rigidity. However, the ST20 and ST40 films showed a lower elongation than the ST0 and ST5 films for 15 days of
storage, which reinforce that, besides the storage time, the TEOS content affected significantly the mechanical properties of the films.
These properties can be correlated with increasing silicon content in the center of the film, as observed by EDS. The inorganic
precursor added to the film did not act in the loss of mobility of the chains for 3 days of storage. On the other hand, it is possible to
notice a decrease in elongation value for 15 days of storage, suggesting strong hydrogen bonds between the TEOS, the starch and the
glycerol, which can induce a loss of macromolecular mobility and avoid the sliding of a starch chains over the others.
This effect was also observed by Lim et al. [14] being reported by the authors an increase in tensile strength and a decrease in
elongation percentage of the films, according to the increase of boric acid content. Thus, the authors suggested the formation of a
strong intermolecular crosslinking between the two components with covalent bonding, leading to a more rigid structure and
hindrance of the sliding of chains [14]. The tensile strength and Young's modulus of the ST0, ST5 to ST20 films increase when
different storage times are compared, a fact that occurred in the opposite way in the ST40. This is related probably to the aging
process during the storage time, when the rearrangement of starch chains and the condensation of TEOS occur. Thus, films with lower
amounts of TEOS become stronger after many days, while the ST40 film seems to be more resistant in a few days. The amount of
TEOS in the ST40 film seems to be an excess considering the migration of inorganic component for the borders, which can accelerate
the aging process by condensation of TEOS on the surface of the film, as shown by the white dots in Fig. 3. On the other hand, high
amounts of the inorganic component seem to act as reinforcement on mechanical properties of the films.
Thus, the silica synthesized in situ during the aging process acts as reinforcement in the polymeric material. Moreover, lower

170
K.O. Lima et al. European Polymer Journal 89 (2017) 162–173

Fig. 5. Water vapor permeability (WVP) values of bean starch films with different TEOS contents stored for 3 and 15 days and analyzed during 7 days. The heights of
columns represent the mean ± deviation values, different letters (a,b,c,d) on the columns indicate significant difference between ST0, ST5, ST20, ST40, and different
letters (A,B,C,D,E) on the columns indicate significant difference between the days to same films submitted to the Tukey test (p < 0.05).

melting temperature of films can be associated to this difficulty in ordering the chains.

3.2.8. Water vapor permeability of the films


The water vapor permeability is a measure of the ease with which the water vapor can permeate the material under controlled
conditions. According to Fig. 5, films stored for 3 days showed lower values of WVP when compared with those stored for 15 days,
except the ST40. ST5 film stored for 3 days showed a decrease in WVP values over 7 days of analyses, the same result was observed
for the ST20 film. However, the ST20 film showed a decrease of WVP also for 15 days of storage. The differences in WVP values can
be attributed to changes in the film structure during the aging process [44].
Regarding the ST40 film, no significant difference was observed during 7 days of analysis. On the other hand, the storage time
seems to be an important parameter to obtain lower WVP values. Films stored for 15 days showed lower WVP values, evidencing that
higher concentrations of the inorganic precursor need a longer storage time for the aging process to occur. The water diffusivity in
materials and the water solubility coefficient of the film are related directly to the permeability of water vapor in a polymer matrix
[45]. Thus, reducing the water transport by diffusion in the polymer, by incorporating a reagent, and decreasing the solubility
coefficient, a reduction in values of permeability to water vapor can be achieved. This fact may suggest that, if this material were
applied in packaging with a longer shelf life, the interactions between components would reduce the available hydroxyl groups and
would decrease the solubility coefficient of the films [46]. The time of storage and addition of inorganic precursor TEOS have an
important role to obtain packaging with lower WVP.

3.2.9. Contact angles (CA)


The surface hydrophobicity properties of hybrid films were investigated by analysis of the contact angle. The images obtained are
illustrated in Fig. 6, the average contact angle of the films were 56.7 ± 0.62d, 58.7 ± 0.05c, 62.3 ± 1.04b, 73.3 ± 0.58a for the

171
K.O. Lima et al. European Polymer Journal 89 (2017) 162–173

Fig. 6. Images of the contact angle (CA) of the ST0 (A), ST5 (B), ST20 (C) and ST40 (D) films.

ST0, ST5, ST20 and ST40, respectively. It is known that an increase in the contact angle indicates an increase in film hydrophobicity.
According to Slavutsky and Bertuzzi [47] the formation of hydrogen bonds between chains reduces the interaction between water and
the film surface, which corroborates the increase in hydrogen bonds observed in FTIR results.
Thus, a correlation between TEOS content and contact angle can be observed, since the contact angle increases as the inorganic
precursor is added. Organic components containing hydroxyl groups and inorganic precursor probably interact by strong hydrogen
bonds enough so that act as non-covalent crosslinking points on the chains presenting more hydrophobic surface.

4. Conclusion

Homogeneous, bubble-free and transparent hybrid films of carioca bean starch and TEOS were synthesized successfully. The TEOS
content and storage time of films develop an important role in the mechanical and barrier properties of the films. The mechanical
performance of the hybrid films were improved by adding 40% of the TEOS, which suggests a strong interaction between the
components. The water vapor permeability and solubility of films in water were improved when the films were stored for 15 days,
allowing the aging process for both starch and the inorganic precursor. Films with a higher TEOS content show more hydrophobic
surface, as well as a higher opacity and migration of silica to the border of the film. The migration of silica to the border could avoid
the reorganization of the starch chains, and lead to a lower melting temperature for the films.

Acknowledgements

We would like to thank CAPES, CNPQ (Ação transversal n° 06/2011 – Casadinho/Procad – Process n° 552197/2011-4) and
FAPERGS for project financing, CEME-SUL for the microscopy and XRD analyses, and professor Daniel Eduardo Weibel of IQ/UFRGS
for the contact angle analyses.

References

[1] F. Xie, E. Pollet, P.J. Halley, L. Avérous, Starch-based nano-biocomposites, Prog. Polym. Sci. 38 (2013) 1590–1628.
[2] A. Cano, A. Jiménez, M. Cháfer, C. Gónzalez, A. Chiralt, Effect of amylose:amylopectin ratio and rice bran addition on starch films properties, Carbohydr. Polym.
111 (2014) 543–555.
[3] J.M.V. Blanshard, Starch granule structure and function: a physicochemical approach, in: T. Galliard (Ed.), Starch: Properties and Potentials, Society of Chemical
Industry, London, 1987, pp. 16–54.
[4] C. Menzel, M. Andersson, R. Andersson, J.L. Vázquez-Gutiérrez, G. Daniel, M. Langton, M. Gällstedt, K. Koch, Improved material properties of solution-cast
starch films: Effect of varying amylopectin structure and amylose content of starch from genetically modified potatoes, Carbohydr. Polym. 130 (2015) 388–397.
[5] N. Singh, Functional and physicochemical properties of pulse starch, Pulse Foods: Processing, Quality and Technological Applications, Academic press, 2011, pp.
91–120.
[6] B.dos S. Siqueira, P.Z. Bassinello, S.C. Santos, G. Malgaresi, P.H. Ferri, A.G. Rodriguez, K.F. Fernandes, Do enzymatic or non-enzymatic pathways drive the

172
K.O. Lima et al. European Polymer Journal 89 (2017) 162–173

postharvest darkening phenomenon in carioca bean tegument? LWT – Food Sci. Technol. 69 (2016) 593–600.
[7] B.S. Siqueira, P.Z. Bassinello, G. Malgaresi, W.J. Pereira, K.F. Fernandes, Analyses of technological and biochemical parameters related to the HTC phenomenon
in carioca bean genotypes by the use of PCA, LWT – Food Sci. Technol. 65 (2016) 939–945.
[8] E.M.M. da Silva, J.L.R. Ascheri, C.W.P. de Carvalho, C.Y. Takeiti, J.de J. Berrios, Physical characteristics of extrudates from corn flour and dehulled carioca bean
flour blend, LWT – Food Sci. Technol. 58 (2014) 620–626.
[9] I.M. Demiate, A.M. Figueroa, M.E.B. Zortéa Guidolin, T.P. Rodrigues dos Santos, H. Yangcheng, F. Chang, J. lin Jane, Physicochemical characterization of
starches from dry beans cultivated in Brazil, Food Hydrocolloids 61 (2016) 812–820.
[10] G. Rupollo, N.L. Vanier, E.R. Zavareze, M. Oliveira, J.M. Pereira, R.T. Paraginski, A.R.G. Dias, M.C. Elias, Pasting, morphological, thermal and crystallinity
properties of starch isolated from beans stored under different atmospheric conditions, Carbohydr. Polym. 86 (2011) 1403–1409.
[11] N.L. Vanier, E.R. Zavareze, V.Z. Pinto, B. Klein, F.T. Botelho, A.R.G. Dias, M.C. Elias, Physicochemical, crystallinity, pasting and morphological properties of
bean starch oxidised by different concentrations of sodium hypochlorite, Food Chem. 131 (2012) 1255–1262.
[12] R. Hoover, T. Hughes, H.J. Chung, Q. Liu, Composition, molecular structure, properties, and modification of pulse starches: a review, Food Res. Int. 43 (2010)
399–413.
[13] L. Dai, C. Qiu, L. Xiong, Q. Sun, Characterisation of corn starch-based films reinforced with taro starch nanoparticles, Food Chem. 174 (2015) 82–88.
[14] M. Lim, H. Kwon, D. Kim, J. Seo, H. Han, S.B. Khan, Highly-enhanced water resistant and oxygen barrier properties of cross-linked poly(vinyl alcohol) hybrid
films for packaging applications, Prog. Org. Coat. 85 (2015) 68–75.
[15] Y. Zhang, Z. Liu. Starch-based edible films, in: E. Chiellini (Ed). Environmentally compatible food packaging, Cambridge, 2009, pp 108–136.
[16] D. Muscat, M.J. Tobin, Q. Guo, B. Adhikari, Understanding the distribution of natural wax in starch–wax films using synchrotron-based FTIR (S-FTIR),
Carbohydr. Polym. 102 (2014) 125–135.
[17] B. Priya, V.K. Gupta, D. Pathania, A.S. Singha, Synthesis, characterization and antibacterial activity of biodegradable starch/PVA composite films reinforced with
cellulosic fibre, Carbohydr. Polym. 109 (2014) 171–179.
[18] A.M. Shi, L.J. Wang, D. Li, B. Adhikari, Characterization of starch films containing starch nanoparticles part 1: physical and mechanical properties, Carbohydr.
Polym. 96 (2013) 593–601.
[19] X. Jia, Y. Li, Q. Cheng, S. Zhang, B. Zhang, Preparation and properties of poly(vinyl alcohol)/silica nanocomposites derived from copolymerization of vinyl silica
nanoparticles and vinyl acetate, Eur. Polymer J. 43 (2007) 1123–1131.
[20] M.D. Morales-Acosta, C.G. Alvarado-Beltrán, M.A. Quevedo-López, B.E. Gnade, B.E. Gnade, A. Mendoza-Galván, R. Ramírez-Bom, Adjustable structural, optical
and dielectric characteristics in sol-gel PMMA-SiO2 hybrid films, J. Non-Cryst. Solids 362 (2013) 124–135.
[21] H. Xiong, S. Tang, H. Tang, P. Zou, The structure and properties of a starch-based biodegradable film, Carbohydr. Polym. 71 (2008) 263–268.
[22] V. Singh, S.K. Singh, S. Pandey, P. Kumar, Sol-gel synthesis and characterization of adsorbent and photoluminescent nanocomposites of starch and silica, J. Non-
Cryst. Solids 357 (2011) 194–201.
[23] K. Yao, J. Cai, M. Liu, Y. Yu, H. Xiong, S. Tang, S. Ding, Structure and properties of starch/PVA/nano-SiO2 hybrid films, Carbohydr. Polym. 86 (2011)
1784–1789.
[24] S.J. McGrane, H.J. Cornell, C.J. Rix, A simple and rapid colourimetric method for determination of amylose in starch products, Starch/Stärke 50 (1998)
158–163.
[25] J.F. Rabek, Experimental Methods in Polymer Chemistry: Applications of Wide-Angle X-ray Diffraction (WAXD) to the Study of the Structure of Polymers, Wiley
Interscience, Chichester, 1980, pp. 505–508.
[26] B. Biduski, F.T. Silva, W.M. Silva, S.L.M. El Halal, V.Z. Pinto, A.R.G. Dias, E.R. Zavareze, Impact of acid and oxidative modifications, single or dual, of sorghum
starch on biodegradable films, Food Chem. 214 (2017) 53–60.
[27] ASTM D882-02. Standard Test Method for Tensile Properties of Thin Plastic Sheeting, ASTM International, West Conshohocken, PA. < www.astm.org > , 2002.
[28] Hunterlab. The color management company. Universal software, version 3.2. Reston, 1997.
[29] ASTM F2251-13. Standard Test Method for Thickness Measurement of Flexible Packaging Material, ASTM International, West Conshohocken, PA, < www.astm.
org > , 2013.
[30] N. Gontard, C. Duchez, J.L. Cuq, S. Guilbert, Edible composite films of wheat gluten and lipids: water vapor permeability and other physical properties, Int. J.
Food Sci. Technol. 29 (1994) 39–50.
[31] ASTM E96/E96M-14. Standard Test Methods for Water Vapor Transmission of Materials, ASTM International, West Conshohocken, PA, < www.astm.org > ,
2014.
[32] N.B. Colthup, L.H. Daly, S.E. Wiberley, Introduction to Infrared and Raman Spectroscopy, thirth ed., Longman, New York, 1990.
[33] X. He, M. Du, H. Li, T. Zhou, Removal of direct dyes from aqueous solution by oxidized starch cross-linked chitosan/silica hybrid membrane, Int. J. Biol.
Macromol. 82 (2016) 174–181.
[34] D.M. Panaitescu, A.N. Frone, M. Ghiurea, I. Chiulan, Influence of storage conditions on starch/PVA films containing cellulose nanofibers, Ind. Crops Prod. 70
(2015) 170–177.
[35] J.M.G. Cowie, Polymer: Chemistry and Physics of Modern Materials, Chapman & Hall, London, 1991.
[36] W. Tongdeesoontorn, L.J. Mauer, S. Wongruong, P. Rachtanapun, Water vapour permeability and sorption isotherms of cassava starch based films blended with
gelatin and carboxymethyl cellulose, Asian J. Food Agro-Ind. 2 (2009) 501–514.
[37] W.D. Callister Jr. Ciência e engenharia de materiais: uma introdução. (fifth ed.) Rio de Janeiro: Livros técnicos e científicos- LTC, 2002 (Chapter 16).
[38] A.N. Frone, C.A. Nicolae, R.A. Gabor, D.M. Panaitescu, Thermal properties of water-resistant starch – polyvinyl alcohol films modified with cellulose nanofibers,
Polym. Degrad. Stab. 121 (2015) 385–397.
[39] O.S. Lawal, K.O. Adebowale, B.M. Ogunsanwo, L.L. Barba, N.S. Ilo, Oxidized and acid thinned starch derivatives of hybrid maize: functional characteristics,
wide-angle X-ray diffractometry and thermal properties, Int. J. Biol. Macromol. 35 (2005) 71–79.
[40] L.S. Guinesi, A.L. Roz, E. Corradini, L.H.C. Mattoso, E.M. Teixeira, A.A.S. Curvelo, Kinetics of thermal degradation applied to starches from different botanical
origins by non-isothermal procedures, Thermochim. Acta 447 (2006) 190–196.
[41] D.S.F. Gay, The effects of temperature of condensation on the thermal stability and morphology of 1,4-phenylenediamine-1-propylsilica xerogels, J. Sol-Gel. Sci.
Technol. 34 (2005) 189–195.
[42] F. Mammeri, E.L. Bourhis, L. Rozes, C. Sanchez, Elaboration and mechanical characterization of nanocomposites thin films. Part I: determination of the
mechanical properties of thin films prepared by in situ polymerisation of tetraethoxysilane in poly(methyl methacrylate), J. Eur. Ceram. Soc. 26 (2006) 259–266.
[43] S. Mali, M.V.E. Grossmann, M.A. García, M.N. Martino, N.E. Zaritzky, Barrier, mechanical and optical properties of plasticized yam starch films, Carbohydr.
Polym. 56 (2004) 129–135.
[44] C.A. Romero-Bastida, L.A. Bello-Pérez, M.A. García, M.N. Martino, J. Solorza-Feria, N.E. Zaritzky, Physicochemical and microstructural characterization of films
prepared by thermal and cold gelatinization from non-conventional sources of starches, Carbohydr. Polym. 60 (2005) 235–244.
[45] C.M. Muller, F. Yamashita, J.B. Laurindo, Evaluation of the effects of glycerol and sorbitol concentration and water activity on the water barrier properties of
cassava starch films through a solubility approach, Carbohydr. Polym. 72 (2008) 82–87.
[46] N.J. Morales, R. Candal, L. Famá, S. Goyanes, G.H. Rubiolo, Improving the physical properties of starch using a new kind of water dispersible nano-hybrid
reinforcement, Carbohydr. Polym. 127 (2015) 291–299.
[47] A.M. Slavutsky, M. Bertuzzi, Water barrier properties of starch films reinforced with cellulose nanocrystals obtained from sugarcane bagasse, Carbohydr. Polym.
110 (2014) 53–61.

173

You might also like