You are on page 1of 10

pubs.acs.

org/acsapm Article

Biodegradable Films Derived from Corn and Potato Starch and


Study of the Effect of Silicate Extracted from Sugarcane Waste Ash
Luciana C. de Azevedo,* Suzimara Rovani,* Jonnatan J. Santos, Djalma B. Dias, Sandi S. Nascimento,
Fab́ io F. Oliveira, Leonardo G. A. Silva, and Denise A. Fungaro
Cite This: ACS Appl. Polym. Mater. 2020, 2, 2160−2169 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via NATL INST OF TECH KOZHIKODE on September 4, 2023 at 12:22:04 (UTC).

ABSTRACT: The growing concern with the amount of plastic materials found in
the oceans makes it necessary to develop biodegradable materials that have low
toxicity to marine animals and humans, but at the same time are resistant to the
actions of microorganisms such as fungi or bacteria. On the other hand, agricultural
waste rich in inorganic materials (such as silica) is often discarded, while it could
be reused as a source of raw material. Considering these points, sodium silicate
solution extracted from sugarcane waste ash was utilized to prepare biodegradable
bioplastics based on corn starch and potato starch. The starch-based bioplastics
were produced by casting and characterized by several physical-chemical
techniques evaluating tensile strength, elongation at break, color analyses,
transparency, opacity, moisture, and biodegradation assay. Bioplastics prepared
with corn starch presented better physical, mechanical, and thermal properties and
optical quality than bioplastics based on potato starch. The samples called CS3 and
PS3, with 5.0% glycerol, were the most resistant to tensile strengths of 0.73 and
0.36 MPa, respectively. On the other hand, the highest elongation at break values were found for the samples with 7.5% glycerol
(CS9, 52.90% and PS9, 49.33%). Corn starch-based bioplastics were more thermally resistant (CS3, 152.86 °C and CS9, 135.20 °C)
when compared to potato starch-based bioplastics (PS3, 140.39 °C and PS9, 127.57 °C). In general, the addition of sodium silicate
solution improved the mechanical and thermal properties of both types of bioplastics. The potato starch-based bioplastics were
biodegraded in 5 days, while those made from corn starch took almost 40 days. The inclusion of sodium silicate inhibited fungal
growth for both corn starch and potato starch bioplastics. The results suggest that sodium silicate solution obtained from renewable
sources can be incorporated into starch-based bioplastics for production of biodegradable packaging with antifungal activity.
KEYWORDS: starch-based bioplastics, sugarcane waste ash, corn starch, potato starch, sodium silicate

1. INTRODUCTION obtained from starch show broad utility in food packaging,


drug transport, and bioengineering.
Growing concern for the environment has stimulated the
The gelation of starch is one of its most vital properties,
development of biodegradable bioplastics produced from
involving the irreversible transformation of the granular starch
renewable sources.1,2 Natural polymers such as starch, chitin, into a viscoelastic paste. When granular starch is subjected to
cellulose, gelatin, and lignin are excellent alternatives to heating in an aqueous medium, some internal hydrogen
decrease the use of non-biodegradable substances,3 and bondsresponsible for the stabilization of the crystalline
many studies have been reported on starch-based bioplastics structure of the granuleare broken, and water enters its
since 1950. 4,5 Recent research has demonstrated the molecular structure, causing grain rupture and loss of
importance of starch-based bioplastics as substitutes for birefringence. At this time, the gelatinization of the starch
conventional plastics;6−10 however, the technology for occurs.13
obtaining these materials still requires further adjustments to The plasticizers most commonly used in starch-based
provide mechanical and barrier properties compatible with bioplastics are polyols, such as glycerol and sorbitol. Glycerol
those of petroleum-derived polymers.
Starch, in particular, has been considered one of the most Received: February 4, 2020
promising materials due to its availability from a variety of Accepted: May 14, 2020
renewable sources (including wheat, rice, corn, potato, and Published: May 14, 2020
cassava), low cost, and good performance; further, it has
excellent biocompatibility and is utterly biodegradable in soil
and water.10−12 According to the literature, bioplastics

© 2020 American Chemical Society https://dx.doi.org/10.1021/acsapm.0c00124


2160 ACS Appl. Polym. Mater. 2020, 2, 2160−2169
ACS Applied Polymer Materials pubs.acs.org/acsapm Article

is a very useful plasticizer because it presents a smaller capacity tape and covered with gold in the BAL-TEC sample coater/sputterer,
for interaction with the starch molecules, giving greater model SCD 050.
wettability to the chains of the polymeric matrix, destroying X-ray diffraction (XRD) analyses were performed using a Rigaku
the crystalline structure of starch; it also improves the thermal Multiflex diffractometer with a Cu anode using Co Kα radiation at
40.0 kV and 20.0 mA over the 2θ range of 5−80° with a scan rate of
stability and thermoplasticity of the starch bioplastics.14,15 The 0.5° min−1.
simplest and most straightforward method for preparing Fourier transform infrared (FTIR) spectroscopy was performed
starch-based bioplastics is casting, which involves the using an Alpha spectrometer from Bruker, operating in attenuated
preparation of a film-forming solution, its deposition in a total reflectance (ATR) mode. The spectra were obtained using 200
suitable support, and subsequent drying of the solvent at room cumulative scans in the range 375−4000 cm−1.
temperature.16−19 Thermogravimetric analysis (TGA) was recorded on a TGA/
Some of the technological applications of starch-based SDTA thermogravimetric analyzer from Mettler Toledo. The sample
bioplastics,20 for example, bioplastics for food packaging, weighed ∼10.0 mg and was analyzed under a nitrogen atmosphere
require specific functions that cannot be satisfied by the starch with a flow of 65.0 mL min−1, using an alumina-port sample heated to
600 °C with a heating rate of 20 °C min−1.
alone. Sodium silicate solution can be added to the film- Differential scanning calorimetry (DSC) was performed on a DSC
forming solution because, according to many studies reported 822e instrument from Mettler Toledo. Dry samples (∼40.0 mg) were
in the literature,21−23 the silicate has good antifungal analyzed under a nitrogen atmosphere with a flow of 30.0 mL min−1,
properties, which is considered very important for the using an aluminum-port sample heated 250 °C with a heating rate of
development of food packaging. Silicate is a very inexpensive 10 °C min−1.
reagent of great abundance that is usually extracted from 2.4. Mechanical Properties. The thickness of the samples
sand,24 but its extraction from renewable sources, in particular (conditioned at 25 °C and 46% relative humidity (RH) for 72 h) was
plant waste, such as that from rice,25 bamboo,26 wheat,27 and measured with an INSIZE 3109-25 micrometer (São Paulo, Brazil)
sugarcane,28 has gained significant importance as the before the tests. The measurements were taken at five different
positions for each sample, and the average value of these
generation of waste from the agriculture industry grows determinations was calculated. This average value was used to
every year. calculate the cross-sectional area of the samples (the area is equal to
Therefore, the objectives of this study based, on a literature the thickness multiplied by the width of each sample).
review, were to develop a process for the preparation of starch- The mechanical properties, tensile strength, and percent elongation
based bioplastics, evaluate the ability to form bioplastics with at break were performed with three or more specimens (size of 73 ×
two different sources of starch, and investigate the addition of 12 mm) cut from each bioplastic. The tensile strength and percent
sodium silicate solution extracted from sugarcane waste ash. elongation at break were measured using an Instron 5567 electro-
Other objectives were to evaluate some physical and mechanical universal testing machine (São José dos Pinhais, Paraná,
mechanical properties of the products obtained and also Brazil), using a calibrated 1 kN (100 kg) load cell, according to
ASTM Standard Methods ASTM D882-12 and ASTM D638-14, with
perform a biodegradability assay of starch-based bioplastics.
some adaptations. The initial grip separation was 45 mm, and the
cross-head speed was 50 mm/min. The tensile strength was calculated
2. EXPERIMENTAL SECTION by dividing the maximum force exerted on the bioplastic during
2.1. Materials. All aqueous solutions were prepared using fracture by the cross-sectional areas. Percent elongation at break was
deionized water (resistivity >18.2 MΩ·cm) obtained from a Milli-Q expressed as the percentage of change of the original length of a
deionizer (Elix Millipore). Sugarcane waste ash (SWA) was donated specimen between grips at the break.34,35
by COSAN S.A. Sodium hydroxide micropearls (>99%) and 2.5. Transparency and Opacity. Color analyses of bioplastics
hydrochloric acid (35−37%) were purchased from Synth, Brazil. were determined using a Miniscan EZ colorimeter from Hunterlab
Glycerol (≥99.5%) was obtained from Sigma-Aldrich, Brazil. Corn (Murnau, Germany), adopting L* (luminosity), a* (red-green), and
starch was purchased from Unilever Brazil Industrial, and potato b* (yellow-blue), which were analyzed by using the CIELab scale.1 In
starch was purchased from Dinâmica Quimica ́ Contemporânea Ltd., the case of bioplastics, the attributes a* and b* represent redness and
Brazil. yellowing, respectively.
2.2. Procedures. 2.2.1. Extraction of Sodium Silicate Solution Transparency and opacity were determined by cutting rectangular
from Sugarcane Waste Ash. Sodium silicate (Na2SiO3) solution was strips of the bioplastic, with size of 0.8 × 30 mm, which were then
obtained as previously published,29,30 wherein SWA was mixed with adhered to the inner wall of a quartz cuvette, to remain perpendicular
NaOH solid (1:1.5, w/w) and heated to 550 °C over 1 h in a muffle to the light beam to read the transmittance in the range from 200 to
oven. After 1 h, the mixture was cooled, and deionized water was 800 nm (model Cary 1E UV−vis spectrophotometer from Varian,
added to the molten mixture, which was then refluxed for 4 h. Finally, California, USA), using the empty cuvette as a control.7,36 The
the mixture was filtered to remove the solid residue from the sodium transparency was calculated according to eq 1:
silicate solution.
2.2.2. Casting of Bioplastics. The filmogenic solution was prepared transparency = 1/(A500 /t ) (1)
by dissolving corn starch or potato starch in 0.10 mol L−1 HCl
aqueous solution under stirring and heating for 30 min. Glycerol and where A500 is the absorption at 500 nm and t is the bioplastic
sodium silicate solution were added to the mixture when the thickness in mm.
temperature reached at 55 and 65 °C, respectively (Table S1).31,32 Apparent opacity was determined on the same spectrophotometer,
According to the casting technique, for each formulation, a specific according to the methodology described in ref 37. The absorbance at
content of the filmogenic solution was poured onto rectangular plates, 500 nm was read for each bioplastic, and the opacity was calculated
followed by drying at ambient temperature for approximately 96 according to eq 2:
h.1,6,33 A flow diagram of the procedure used to obtain the
biodegradable films is shown in Figure S1. opacity = A500 /t (2)
2.3. Characterization. Scanning electron microscopy (SEM)
images were recorded using a Quanta FEG 650 tabletop microscope where A500 is the absorption at 500 nm and t is the bioplastic
from Thermo Fischer Scientific (Oregon, USA) and a model TM thickness in mm.
3000 tabletop microscope from Hitachi (Tokyo, Japan). Before the The moisture was determined in an oven at 105 °C (∼4 h),
SEM analysis, the sample was fixed in the sample holder with carbon according to the methodology suggested by Rocha et al. (2014).37

2161 https://dx.doi.org/10.1021/acsapm.0c00124
ACS Appl. Polym. Mater. 2020, 2, 2160−2169
ACS Applied Polymer Materials pubs.acs.org/acsapm Article

2.6. Biodegradation Assay. A biodegradation assay of the


starch-based bioplastics was performed as follows. The bioplastics
were cut into 4 × 10 cm pieces (40 cm2 area) and packed into foil
bags with four pieces of each sample per bag (n = 4), as shown in
Figure S2. They were then buried in a garden area in the IF SERTÃ O-
PE, Petrolina-PE campus, at a depth of 20 cm under soil (Figure S3).
Samples were dug up and observed periodically, every 48 h, for 2
weeks. From the third week, the frequency of observations was
reduced to weekly. The total period of observation was 40 days, and
the term was defined according to the degradation state of the
bioplastics. The observations were made based on the bioplastic
sample’s area and completed when the samples decreased their area Figure 1. Infrared spectra of starch-based bioplastics: (a) CS samples
and degraded. Evaluation of the decomposition of the residues and (b) PS samples.
followed the simple exponential model used by Thomas and Asakawa
(1993),38 using eq 3:
cm−1, associated with OH stretching, attributed to the
X t = X 0 e−kt (3) hydrogen bonds (H−O···H) between the plasticizer molecules
where Xt is the amount of dry matter remaining after a period of time and the starch molecules and also to the presence of water in
t, X0 is the initial amount of dry matter, k is the decomposition the starch-based bioplastics.40 Two bands of low intensity
constant, and t is the time in days. By rearranging the terms of this located at 2930 and 2885 cm−1, attributed to the C−H
equation, it is possible to calculate the value of the decomposition stretching, and a band at 1650 cm−1, corresponding to the
constant k, using eq 4: δ(O−H) flexion due to the tightly bound water, were observed
in the spectra.41 A medium-intensity band was observed at
k = ln(X t /X 0)/t (4)
1370 cm−1, attributed to bending of the CH2 and C−OH
Half-life is another important parameter in the evaluation of the groups. The 1150 and 1079 cm−1 bands observed correspond
decomposition of plant residues, expressing the time, in days, to the C−O stretch vibration in C−O−H groups, and the 1110
necessary for half of the material to decompose, or for half of the cm−1 band observed only for starch-based bioplastic samples is
nutrients contained in the residues to be released. According to attributed to the C−O stretch vibration in C−O−C groups.41
Rezende et al. (1999),39 it is possible to calculate the half-life through These results corroborate with those reported by Dang and
eq 5:
Yoksan (2015).1 According to Shi et al. (2007),42 the bands at
t 1/2 = ln(2)/k (5) 1040 and 1025 cm−1 are sensitive to the amount of ordered or
crystalline starch and related to amorphous starch, respectively.
The two low-intensity bands at 925 and 850 cm−1 observed
3. RESULTS in the IR spectra are characteristic absorption bands of
3.1. Description of Samples. In the present study, we anhydroglucose ring stretching vibrations of the starch
evaluated various conditions for the formation of biodegrad- structure.41 The band at 925 cm−1 is attributed to a skeletal
able bioplastics with fungicidal properties from corn starch and mode involving an α-1,4-glycosidic linkage (C−O−C), and the
potato starch prepared with sodium silicate solution extracted band at 570 cm−1 is attributed to skeletal modes of the
from sugarcane waste ash. A total of 16 samples were prepared, pyranose ring.43
and the four best results are presented, as well as a bioplastics These results demonstrate that the implicit chemical
control. All conditions investigated are described in Table S1. properties of starch are not altered, and the presence of
These results were obtained after optimization of the process glycerolresponsible for maintaining the adhesion between
for preparation of the bioplastics, in which the quantities of the polymeric starch chains and for the plastic flexibilityat
starch, hydrochloric acid, glycerol, and sodium silicate solution different concentrations (5% for CS3 and PS3 and 7.5% for
were varied. The percentage values of the components were CS9 and PS9) does not change the chemical bonds/groups.
calculated in relation to the volume of the filmogenic solution. The silicate peaks in turn are not observed, since they are in
The samples prepared with corn starch and potato starch were low concentration.
named as CS and PS, respectively, where the samples CS1 and The XRD patterns of native corn starch (CS), native potato
PS1 do not contain silicate, but to each was added 0.10 mol starch (PS), and starch bioplastics modified with silicate show
L−1 NaOH solution (control bioplastics). The complete similar results, as can be seen in Figure 2.
characterization of these samples is given below. As can be seen from Figure 2, CS and PS present very
3.2. Structural Characterization and Mechanical similar spectra, with main diffraction peaks at 15.2°, 17.5°,
Properties. The formation of bioplastics occurs through a 23.0° and 14.9°, 17.4°, 23.0°, respectively. Generally, the
process of starch organization during drying. The chemical native starch crystallinity is of types A, B, and C, as an
influence of sodium silicate and glycerol concentration was intermediate form crystalline.44−46 The peaks exhibited by CS
evaluated by infrared (IR) spectroscopy and XRD. and PS are the A-type diffraction patterns typical of corn and
The possible bond formation or modification was assessed potato starch.47 The peak at 2θ = 17.1° and 17.5° of some
by IR spectroscopy. The spectra obtained for the starch-based starch-based bioplastics, including the control bioplastics CS1
bioplastic samples before the addition of sodium silicate and PS1, is attributed to typical peaks of amylose complexed
solution (CS1 and PS1) and with sodium silicate and two with glycerol.48
different contents of glycerol (CS3, CS9 and PS3, PS9) are 3.3. Morphological Characteristics and Thermal
shown in Figure 1. Stability. IR spectroscopy and XRD do not show significant
In general, no changes are observed in the bioplastics spectra alterations in the chemical characteristics of the bioplastics. On
after addition of the sodium silicate solution. All IR spectra of the other hand, SEM images present real changes in the
the starch-based bioplastics present an intense band at 3285 physical characteristics, as can be seen in Figure 3.
2162 https://dx.doi.org/10.1021/acsapm.0c00124
ACS Appl. Polym. Mater. 2020, 2, 2160−2169
ACS Applied Polymer Materials pubs.acs.org/acsapm Article

resistant to the tensile strength when compared to samples


CS9, PS3, and PS9. This is probably due to the higher amylose
content in corn starch than in potato starch,6 corroborating the
results obtained by Luchese et al. (2018),3 where the corn
starch-based films also presented higher tensile strength than
potato, cassava, and wheat starch-based films.
Comparing the different contents of glycerol for the starch-
based bioplastics, the highest E values (Table S2) were
obtained for the samples with 7.5% glycerol (CS9 and PS9).
The addition of sodium silicate in the starch-based bioplastics
decreased the elongation at break and the tensile strength; on
the other hand, the E values found in this study for corn and
potato starch-based bioplastics were about 3 and 10 times
higher, respectively, than those found by Basiak et al. (2017)51
for films obtained with the same starches.
Thermogravimetric analysis was used to evaluate the thermal
stability of the starch-based bioplastics, in order to know the
maximum temperature at which these materials could be used
without degradation or loss of their characteristics, as well as to
compare the thermal resistance between the two types of
starch used in the study. Thermogravimetric (TG) and
differential thermogravimetric (DTG) curves of some plasti-
Figure 2. X-ray diffraction patterns of starch-based bioplastics: (a) CS cized samples are shown in Figure 5.
samples and (b) PS samples. From the TG and DTG curves, five and four weight loss
stages can be seen for corn and potato starch-based bioplastics,
The SEM images confirm that plastification and processing respectively (Figure 5 and Table S3). The first mass loss,
completely modified the structure of the starches, as the between 25 and ∼110 °C for all starch-based bioplastics,
resulting bioplastics differ completely from the native starches ranged from ∼5 to ∼10%, except for the PS1 sample (∼16%);
(Figure S4A,B). The morphology of the bioplastics CS1 and this is related to evaporation of free water.3,36 The PS1 sample
PS1 (Figure 3A1,B1) was similar to those of a corn starch/ presented higher free water loss compared to the other samples
chitosan film obtained by Ren et al. (2017)49 and sweet potato since PS1 had the lowest thickness (Table S3).
starch film obtained by Li et al. (2015).50 In the samples with The second weight loss, between ∼110 and ∼200 °C for all
the addition of sodium silicate (CS3, CS9, PS3, and PS9) the starch-based bioplastics, ranged from ∼14 to ∼22%, except for
formation of fibrous silicate was observed, which indicates a PS1 (∼11%) and PS3 (∼40%); this is related to loss of
higher inorganic character for silicate-doped bioplastics. moisture and/or evaporation of water.48 Corroborating the
The results from the measurements of the tensile strength results of moisture analysis (Table S2), PS1 and PS3 have the
and percent elongation at break of the starch-based bioplastics, lowest and highest moisture percent, respectively, among the
obtained for the samples CS/PS1, CS/PS3, and CS/PS9, are studied samples.
shown in Figure 4, corroborating the SEM observations. It can The third stage of mass loss, between ∼200 and ∼300 °C for
be seen that the CS3 sample (0.73 MPa) was the most all starch-based bioplastics, ranged from ∼26 to ∼33%, except

Figure 3. SEM images of starch-based bioplastics: (A1) CS1, (A2) CS3, and (A3) CS9, and (B1) PS1, (B2) PS3, and (B3) PS9.

2163 https://dx.doi.org/10.1021/acsapm.0c00124
ACS Appl. Polym. Mater. 2020, 2, 2160−2169
ACS Applied Polymer Materials pubs.acs.org/acsapm Article

Figure 4. Mechanical properties of starch-based bioplastics: (a) tensile strength (MPa) of CS samples, (b) elongation at break (%) of CS samples,
(c) tensile strength (MPa) of PS samples, and (d) elongation at break (%) of PS samples.

for PS1 (∼54%) and PS3 (∼22%), which demonstrates 76.1 °C and ∼80 °C for rice starch-based biodegradable films
thermal expansion of the material and degradation of glycerol and gelatin/potato starch edible biocomposite films, respec-
(boiling point 290 °C). These results were similar to those tively.
reported by Shi et al. (2007)42 and by Luchese et al. (2018).3 Comparing bioplastics with 5% of glycerol without the
The fourth weight loss, between ∼300 and ∼360 °C for addition of sodium silicate (CS1 and PS1) to bioplastics with
samples with sodium silicate (CS3 and CS9), ranged from ∼4 the addition of 25% sodium silicate solution (obtained from
to ∼10%. This was associated with low-molecular-weight sugarcane waste ash), it was possible to observe an increase of
reaction products.3 The fifth and final weight loss, between 15 °C and a decrease of 5 °C (see Table S4) in the melting
∼360 and ∼500 °C for corn starch-based bioplastics with temperature of most of the crystals of PS3 and CS3,
sodium silicate (CS3 and CS9), ranged from ∼8 to ∼12%, respectively.
corresponding to the degradation of starch and loss of the 3.4. Biodegradability Assay. The biodegradability assay
chemically adsorbed water bonded to Si−OH through the of the bioplastics included observation of the visual aspects
hydrogen bond.51,52 For CS1 (without sodium silicate) the during the period in which the samples remained buried as well
fourth and fifth weight losses (∼15%) occurred until ∼465 and as measurement of the area of the material. The initial proposal
∼500 °C, respectively, and for potato starch-based bioplastics was that the visual analysis would be accompanied by the loss
the fourth and fifth weight losses occurred until ∼465 °C (PS1, of mass of the bioplastics; however, this variable was affected
∼4% and PS3, ∼10%) and until ∼500 °C for the other samples
by soil moisture and aggregation of sand grains and other
and ranged from ∼12 to ∼15%. These samples showed only
residues, and thus it was discarded. Figure 7 shows the CS and
four weight losses, and this fourth corresponds to the
PS bioplastics at the beginning and end of the biodegradability
degradation of starch and loss of the chemically adsorbed
test.
water bonded to Si−OH (from sodium silicate) through the
hydrogen bond.51,52 Above these temperatures no further Degradation curves for corn and potato starch-based
weight loss occurred; only inorganic components remained. bioplastics shown in Figure S6 represent the bioplastic area
Thermal analyses performed by DSC are shown in Figure 6. reduction curve in the test period. It is possible to perceive the
The thermograms exhibited a single endothermic peak, which fragility of the potato starch-based bioplastics (PS), which
indicated homogeneity of the bioplastics.3,53 The Tpeak values degraded in only 5 days, which implies that its use for
of most of the crystals obtained in this work ranged from commercial purposes is unviable, although it is very good from
135.20 to 157.61 °C for CS samples and from 125.12 to an environmental point of view. For corn starch-based
140.39 °C for PS samples (Table S4). These results bioplastics (CS), however, a higher resistance to soil
demonstrate that the starch-based bioplastics have good conditions is observed relative to PS bioplastics. The time
resistance at high temperatures. Similar results were found by required for degradation of CS was approximately 40 days,
Tongdeesoontorn et al. (2011),54 who reported crystallization which can be considered an excellent degradation time when
temperatures of 96−165 °C for cassava starch films. compared to synthetic commercial polymers, like PVC films.
The values found in this study were higher than those Luchese et al. (2018) evaluated the biodegradability of cassava,
reported by Woggum et al. (2014)6 and Podshivalov et al. wheat, and corn-based films and commercial PVC films and
(2017),36 who reported crystallization temperatures of 70.7− observed that after 56 days the starch-based films were
2164 https://dx.doi.org/10.1021/acsapm.0c00124
ACS Appl. Polym. Mater. 2020, 2, 2160−2169
ACS Applied Polymer Materials pubs.acs.org/acsapm Article

Figure 5. TG curves of starch-based bioplastics: (a) CS samples and (e) PS samples. TG/DTG curves of starch-based bioplastics: (b) CS1, (c)
CS3, (d) CS9, (f) PS1, (g) PS3, and (h) PS9. Samples were analyzed under a nitrogen flow of 65 mL min−1 with a heating rate of 20 °C min−1.

Figure 6. DSC curves of starch-based bioplastics: (a) CS samples and (b) PS samples.

2165 https://dx.doi.org/10.1021/acsapm.0c00124
ACS Appl. Polym. Mater. 2020, 2, 2160−2169
ACS Applied Polymer Materials pubs.acs.org/acsapm Article

2015,55 Gliricidia sepium (t1/2 = 14 days), Pueraria phaseoloides


(t1/2 = 23 days), and Panicum maximum (t1/2 = 20 days).
3.5. Antifungal Activity. Silicate has antifungal activity, as
shown by Capeletti et al. (2014)56 and Derbalah et al.
(2018).57 In this study, it was also observed that the addition
of sodium silicate in the starch-based bioplastics inhibited the
growth of fungi. All starch-based bioplastics produced have
starch, glycerol, NaOH, and acid solution. However, in
bioplastics made without sodium silicate, fungal growth was
observed (Figure 8).

Figure 8. Photographs of starch-based bioplastics: (a) CS samples


and (b) PS samples. Fungal growth in control films (CS1 and PS1).

Figure 7. Biodegradability assay images of corn starch-based


bioplastics, (a) day 0 and (b) at the end of the degradation period In the control bioplastics (CS1 and PS1), black dots related
(day 39), and potato starch-based bioplastics, (c) day 0 and (d) at the to fungal growth were observed, and Figure S7 shows different
end of the degradation period (day 5). colors in starch-based bioplastics also associated with fungal
growth. This result agrees with those found in the literature:
practically all bio-disintegrated, while the commercial PVC- Ge et al. (2017)21 observed in their study that sodium silicate
based film remained intact.3 expressively inhibited mycelial growth and spore germination
Table 1 shows the half-life times of the starch-based of Trichothecium roseum (fungus); Li et al. (2009)22 observed
bioplastics (CS and PS) submitted to the biodegradability that sodium silicate had direct fungitoxic activity against
Fusarium sulphureum (fungal plant pathogen) in potato tubers;
Table 1. Half-Life Values of Corn and Potato Starch-Based Bi et al. (2006)58 observed that sodium silicate inhibited
Bioplastics pathogen growth and thus induced disease resistance in Hami
bioplastics t1/2 (days) melons; Ellepola et al. (2011)59 observed that toothpaste with
CS1 22.52 sodium silicate as one of the ingredients presented anti-
CS3 22.34 candidal activity against Candida albicans and Candida
CS9 17.64 dubliniensis.
PS1 3.05
PS3 2.59
4. CONCLUSION
PS9 2.59 The results obtained in this study show that corn starch-based
bioplastics are more colorless, more transparent, and less
opaque and present lower moisture content than potato starch-
test. These values corroborate with the results presented in based bioplastics. However, the thickness of the corn starch
Figure 7. The half-life time of the CS3 sample (t1/2 = 22.34 bioplastics was higher than that of the potato starch bioplastics,
days) approached that of the CS1 sample, whereas for the CS9 but the mechanical and thermal properties remained higher for
sample, containing higher glycerin content, it was lower (t1/2 = the corn starch-based bioplastics.
17.64 days). For the potato starch-based bioplastics, the half- In general, the addition of sodium silicate solution, obtained
life time was short for all samples, ranging from 2.59 to 3.05 from a renewable source, improved the mechanical and
days. These results corroborate with the DSC thermal analyses thermal properties of both types of bioplastics.
(Figure 6), which showed that the corn starch-based The corn starch-based bioplastic obtained with 5% glycerol
bioplastics were more thermally resistant when compared to and 25% sodium silicate solution (CS3) showed the highest
potato starch-based bioplasticsthis is an indication that CS tensile strength.
bioplastics are less biodegradable than PS bioplastics. In the biodegradability assay, although the potato starch-
These results of half-life time obtained for CS samples were based bioplastics had worse physical, mechanical, and thermal
equivalent to that of some organic materials found naturally in properties, they presented higher biodegradability than corn
the soil, such as some plant species studied by de Paula et al. starch-based bioplastics. The potato starch bioplastics were
2166 https://dx.doi.org/10.1021/acsapm.0c00124
ACS Appl. Polym. Mater. 2020, 2, 2160−2169
ACS Applied Polymer Materials


pubs.acs.org/acsapm Article

biodegraded in 5 days, whereas those of corn starch took ACKNOWLEDGMENTS


almost 40 days. This study was financed in part by the Coordenaçaõ de
Finally, the addition of sodium silicate solution from a Aperfeiçoamento de Pessoal de Nivel ́ Superior - Brazil
renewable source inhibited fungal growth for both corn starch (CAPES) - Finance Code 001, in part by the Conselho
and potato starch bioplastics. These starch-based bioplastics ́ e Tecnológico - Brazil
Nacional de Desenvolvimento Cientifico
can be considered for use in biodegradable packaging with (CNPq), and in part by the Fundaçaõ de Amparo à Pesquisa
antifungal activity. However, further research is recommended do Estado de São Paulo - Brazil (FAPESP). The authors also
to improve the main properties of these bioplastics in order for are grateful to COSAN S.A. (São Paulo, Brazil) for supplying
them to withstand external stress and maintain integrity. the sugarcane waste ash.


*
ASSOCIATED CONTENT
sı Supporting Information
■ REFERENCES
(1) Dang, K. M.; Yoksan, R. Development of Thermoplastic Starch
The Supporting Information is available free of charge at Blown Film by Incorporating Plasticized Chitosan. Carbohydr. Polym.
2015, 115, 575−81.
https://pubs.acs.org/doi/10.1021/acsapm.0c00124. (2) Gheribi, R.; Puchot, L.; Verge, P.; Jaoued-Grayaa, N.; Mezni, M.;
Photographs of bioplastics packed in foil bags for Habibi, Y.; Khwaldia, K. Development of Plasticized Edible Films
biodegradability analysis; flow diagram of the procedure from Opuntia Ficus-Indica Mucilage: A Comparative Study of Various
Polyol Plasticizers. Carbohydr. Polym. 2018, 190, 204−211.
used to produce of the biodegradable films; preliminary
(3) Luchese, C. L.; Benelli, P.; Spada, J. C.; Tessaro, I. C. Impact of
test formulations for obtaining starch-based bioplastics the Starch Source on the Physicochemical Properties and Biodegrad-
(bioplastics preparation: selection of samples); SEM ability of Different Starch-Based Films. J. Appl. Polym. Sci. 2018, 135
images of native corn and potato starch; physicochem- (33), 46564.
ical characterization and mechanical and thermal (4) Wolff, I. A.; Davis, H. A.; Cluskey, J. E.; Gundrum, L. J.; Rist, C.
properties of the starch-based bioplastics; and degrada- E. Preparation of Films from Amylose. Ind. Eng. Chem. 1951, 43 (4),
tion curves for starch-based bioplastics (PDF) 915−919.
(5) Treadway, R. H. Uses of Potato Starch and Potato Flour in the

■ AUTHOR INFORMATION
Corresponding Authors
United States. Am. Potato J. 1952, 29 (4), 79−84.
(6) Woggum, T.; Sirivongpaisal, P.; Wittaya, T. Properties and
Characteristics of Dual-Modified Rice Starch Based Biodegradable
Films. Int. J. Biol. Macromol. 2014, 67, 490−502.
Luciana C. de Azevedo − Department of Food Technology, (7) El Miri, N.; Abdelouahdi, K.; Barakat, A.; Zahouily, M.; Fihri, A.;
Federal Institute - IF SERTAO-PE, CEP: 56314-520 Petrolina, Solhy, A.; El Achaby, M. Bio-Nanocomposite Films Reinforced with
Pernambuco, Brazil; Instituto de Pesquisas Energéticas e Cellulose Nanocrystals: Rheology of Film-Forming Solutions, Trans-
Nucleares (IPEN-CNEN/SP), CEP: 05508-000 São Paulo, parency, Water Vapor Barrier and Tensile Properties of Films.
SP, Brazil; Email: luciana.cavalcanti@ifsertao-pe.edu.br Carbohydr. Polym. 2015, 129, 156−67.
Suzimara Rovani − Instituto de Pesquisas Energéticas e (8) Fu, L.; Zhu, J.; Zhang, S.; Li, X.; Zhang, B.; Pu, H.; Li, L.; Wang,
Nucleares (IPEN-CNEN/SP), CEP: 05508-000 São Paulo, Q. Hierarchical Structure and Thermal Behavior of Hydrophobic
Starch-Based Films with Different Amylose Contents. Carbohydr.
SP, Brazil; orcid.org/0000-0002-7802-1050;
Polym. 2018, 181, 528−535.
Email: suzimara.rovani@ipen.br, suzirovani@gmail.com (9) Sessini, V.; Arrieta, M. P.; Fernandez-Torres, A.; Peponi, L.
Humidity-Activated Shape Memory Effect on Plasticized Starch-Based
Authors
́ Biomaterials. Carbohydr. Polym. 2018, 179, 93−99.
Jonnatan J. Santos − Instituto de Quimica, Universidade de São (10) Ceseracciu, L.; Heredia-Guerrero, J. A.; Dante, S.; Athanassiou,
Paulo, CEP: 05508-000 São Paulo, SP, Brazil; orcid.org/ A.; Bayer, I. S. Robust and Biodegradable Elastomers Based on Corn
0000-0003-3789-6229 Starch and Polydimethylsiloxane (PDMS). ACS Appl. Mater. Interfaces
Djalma B. Dias − Instituto de Pesquisas Energéticas e Nucleares 2015, 7 (6), 3742−3753.
(IPEN-CNEN/SP), CEP: 05508-000 São Paulo, SP, Brazil (11) Abdillahi, H.; Chabrat, E.; Rouilly, A.; Rigal, L. Influence of
Sandi S. Nascimento − Department of Food Technology, Citric Acid on Thermoplastic Wheat Flour/Poly(lactic Acid) Blends.
Federal Institute - IF SERTAO-PE, CEP: 56314-520 Petrolina, II. Barrier Properties and Water Vapor Sorption Isotherms. Ind. Crops
Pernambuco, Brazil Prod. 2013, 50, 104−111.
Fábio F. Oliveira − Departamento de Ciências Agrárias/Solos, (12) Liew, C. W.; Ramesh, S. Electrical, Structural, Thermal and
Federal Institute - IF SERTAO-PE, CEP: 56302-970 Petrolina, Electrochemical Properties of Corn Starch-Based Biopolymer Electro-
lytes. Carbohydr. Polym. 2015, 124, 222−8.
Pernambuco, Brazil
(13) Ratnayake, W. S.; Jackson, D. S. Gelatinization and Solubility of
Leonardo G. A. Silva − Instituto de Pesquisas Energéticas e Corn Starch During Heating in Excess Water: New Insights. J. Agric.
Nucleares (IPEN-CNEN/SP), CEP: 05508-000 São Paulo, Food Chem. 2006, 54 (10), 3712−6.
SP, Brazil (14) Bertuzzi, M. A.; Armada, M.; Gottifredi, J. C. Physicochemical
Denise A. Fungaro − Instituto de Pesquisas Energéticas e Characterization of Starch Based Films. J. Food Eng. 2007, 82 (1),
Nucleares (IPEN-CNEN/SP), CEP: 05508-000 São Paulo, 17−25.
SP, Brazil (15) Souza, A. C.; Benze, R.; Ferrão, E. S.; Ditchfield, C.; Coelho, A.
Complete contact information is available at: C. V.; Tadini, C. C. Cassava Starch Biodegradable Films: Influence of
Glycerol and Clay Nanoparticles Content on Tensile and Barrier
https://pubs.acs.org/10.1021/acsapm.0c00124
Properties and Gass Transition Temperature. LWT - Food Sci.
Technol. 2012, 46 (1), 110−117.
Notes (16) Ali, A.; Yu, L.; Liu, H.; Khalid, S.; Meng, L.; Chen, L.
The authors declare no competing financial interest. Preparation and Characterization of Starch-Based Composite Films

2167 https://dx.doi.org/10.1021/acsapm.0c00124
ACS Appl. Polym. Mater. 2020, 2, 2160−2169
ACS Applied Polymer Materials pubs.acs.org/acsapm Article

Reinforced by Corn and Wheat Hulls. J. Appl. Polym. Sci. 2017, 134 (37) Rocha, G. O.; Farias, M. G.; Carvalho, C. W. P. d.; Ascheri, J. L.
(32), 45159. R.; Galdeano, M. C. Filmes compostos biodegradáveis a base de
(17) Bertuzzi, M. A.; Gottifredi, J. C.; Armada, M. Mechanical amido de mandioca e proteina ́ de soja. Polim.: Cienc. Tecnol. 2014, 24,
Properties of a High Amylose Content Corn Starch Based Film, 587−595.
Gelatinized at Low Temperature. Braz. J. Food Technol. 2012, 15, (38) Thomas, R. J.; Asakawa, N. M. Decomposition of Leaf Litter
219−227. from Tropical Forage Grasses and Legumes. Soil Biol. Biochem. 1993,
(18) Claro, P. I. C.; Correa, A. C.; de Campos, A.; Rodrigues, V. B.; 25 (10), 1351−1361.
Luchesi, B. R.; Silva, L. E.; Mattoso, L. H. C.; Marconcini, J. M. (39) Rezende, C. d. P.; Cantarutti, R. B.; Braga, J. M.; Gomide, J. A.;
Curaua and Eucalyptus Nanofibers Films by Continuous Casting: Pereira, J. M.; Ferreira, E.; Tarré, R.; Macedo, R.; Alves, B. J. R.;
Mechanical and Thermal Properties. Carbohydr. Polym. 2018, 181, Urquiaga, S.; Cadisch, G.; Giller, K. E.; Boddey, R. M. Litter
1093−1101. Deposition and Disappearance in Brachiaria Pastures in the Atlantic
(19) Wittaya, T. Rice Starch-Based Biodegradable Films: Properties Forest Region of the South of Bahia, Brazil. Nutr. Cycling Agroecosyst.
Enhancement. In Structure and Function of Food Engineering; Eissa, A. 1999, 54 (2), 99−112.
A., Ed.; IntechOpen: 2012; pp 103−134. DOI: 10.5772/47751 (40) Wang, C.-z.; Li, F.-y.; Wang, L.-m.; Li, J.-f.; Guo, A.-f.; Zhang,
(20) Zhang, B.; Selway, N.; Shelat, K. J.; Dhital, S.; Stokes, J. R.; C.-w.; Liu, P. Research on Thermoplastic Starch and Different Fiber
Gidley, M. J. Tribology of Swollen Starch Granule Suspensions from Reinforced Biomass Composites. RSC Adv. 2015, 5 (62), 49824−
Maize and Potato. Carbohydr. Polym. 2017, 155, 128−135. 49830.
(21) Ge, Y.; Chen, Y.; Li, C.; Wei, M.; Lv, J.; Meng, K. Inhibitory (41) Pankaj, S. K.; Bueno-Ferrer, C.; Misra, N. N.; O’Neill, L.;
Effects of Sodium Silicate on the Fungal Growth and Secretion of Cell Tiwari, B. K.; Bourke, P.; Cullen, P. J. Dielectric Barrier Discharge
Wall-Degrading Enzymes by Trichothecium Roseum. J. Phytopathol. Atmospheric Air Plasma Treatment of High Amylose Corn Starch
2017, 165 (9), 620−625. Films. LWT - Food Sci. Technol. 2015, 63 (2), 1076−1082.
(22) Li, Y. C.; Bi, Y.; Ge, Y. H.; Sun, X. J.; Wang, Y. Antifungal (42) Shi, R.; Liu, Q.; Ding, T.; Han, Y.; Zhang, L.; Chen, D.; Tian,
Activity of Sodium Silicate on Fusarium Sulphureum and its Effect on W. Ageing of Soft Thermoplastic Starch with High Glycerol Content.
Dry Rot of Potato Tubers. J. Food Sci. 2009, 74 (5), M213−8. J. Appl. Polym. Sci. 2007, 103 (1), 574−586.
(23) Zhou, X.; Shen, Y.; Fu, X.; Wu, F. Application of Sodium (43) Wiącek, A. E. Effect of Surface Modification on Starch
Silicate Enhances Cucumber Resistance to Fusarium Wilt and Alters Biopolymer Wettability. Food Hydrocolloids 2015, 48, 228−237.
Soil Microbial Communities. Front. Plant Sci. 2018, 9, 624. (44) Shaw, G. M.; Levy, P. C.; Lobuglio, A. F. Re-Examination of the
(24) Haldimann, M.; Luible, A.; Overend, M. Structural Use of Glass; EA Rosette Assay (Ripley) for Fc Rceptor Leucocytes. Clin. Exp.
International Association for Bridge and Structural Engineering: 2008. Immunol. 1979, 36 (3), 496−501.
(25) Della, V. P.; Kühn, I.; Hotza, D. Rice Husk Ash as an Alternate (45) Li, Y. D.; Xu, T. C.; Xiao, J. X.; Zong, A. Z.; Qiu, B.; Jia, M.;
Source for Active Silica Production. Mater. Lett. 2002, 57 (4), 818− Liu, L. N.; Liu, W. Efficacy of Potato Resistant Starch Prepared by
821. Microwave-Toughening Treatment. Carbohydr. Polym. 2018, 192,
(26) Kow, K.-W.; Yusoff, R.; Aziz, A. R. A.; Abdullah, E. C. From 299−307.
Bamboo Leaf to Aerogel: Preparation of Water Glass as a Precursor. J. (46) Nguyen Vu, H. P.; Lumdubwong, N. Starch Behaviors and
Non-Cryst. Solids 2014, 386, 76−84. Mechanical Properties of Starch Blend Films with Different
(27) Liu, S.-W.; Wei, Q.; Cui, S.-P.; Nie, Z.-R.; Du, M.-H.; Li, Q.-Y. Plasticizers. Carbohydr. Polym. 2016, 154, 112−20.
Hydrophobic Silica Aerogel Derived from Wheat Husk Ash by (47) Singh, V.; Urs, R. G.; Somashekarappa, H.; Ali, S. Z.;
Ambient Pressure Drying. J. Sol-Gel Sci. Technol. 2016, 78 (1), 60−67. Somashekar, R. X-Ray Analysis of Different Starch Granules. Bull.
(28) Alves, R. H.; Reis, T. V. d. S.; Rovani, S.; Fungaro, D. A. Green Mater. Sci. 1995, 18 (5), 549−555.
Synthesis and Characterization of Biosilica Produced from Sugarcane (48) Mendes, J. F.; Paschoalin, R. T.; Carmona, V. B.; Sena Neto, A.
Waste Ash. J. Chem. 2017, 2017, 1−9. R.; Marques, A. C. P.; Marconcini, J. M.; Mattoso, L. H. C.; Medeiros,
(29) Rovani, S.; Santos, J. J.; Corio, P.; Fungaro, D. A. Highly Pure E. S.; Oliveira, J. E. Biodegradable Polymer Blends Based on Corn
Silica Nanoparticles with High Adsorption Capacity Obtained from Starch and Thermoplastic Chitosan Processed by Extrusion.
Sugarcane Waste Ash. ACS Omega 2018, 3 (3), 2618−2627. Carbohydr. Polym. 2016, 137, 452−458.
(30) Rovani, S.; Santos, J. J.; Corio, P.; Fungaro, D. A. An (49) Ren, L.; Yan, X.; Zhou, J.; Tong, J.; Su, X. Influence of Chitosan
Alternative and Simple Method for the Preparation of Bare Silica Concentration on Mechanical and Barrier Properties of Corn Starch/
Nanoparticles Using Sugarcane Waste Ash, an Abundant and Chitosan Films. Int. J. Biol. Macromol. 2017, 105 (Pt 3), 1636−1643.
Despised Residue in the Brazilian Industry. J. Braz. Chem. Soc. (50) Li, J.; Ye, F.; Liu, J.; Zhao, G. Effects of Octenylsuccination on
2019, 30, 1524−1533. Physical, Mechanical and Moisture-Proof Properties of Stretchable
(31) Weber, F. H.; Collares-Queiroz, F. P.; Chang, Y. K. Sweet Potato Starch Film. Food Hydrocolloids 2015, 46, 226−232.
́
Caracterizaçaõ Fisico-Qui ́
mica, Reológica, Morfológica e Térmica (51) Basiak, E.; Lenart, A.; Debeaufort, F. Effect of Starch Type on
dos Amidos de Milho Normal, Ceroso e com Alto Teor de Amilose. the Physico-Chemical Properties of Edible Films. Int. J. Biol.
Cienc. Tecnol. Aliment. (Campinas, Braz.) 2009, 29, 748−753. Macromol. 2017, 98, 348−356.
(32) Souza, R. C. R.; Andrade, C. T. Investigaçaõ dos Processos de (52) Girsova, M.; Golovina, G. F.; Drozdova, I. A.; Polyakova, I.;
Gelatinizaçaõ e Extrusão de Amido de Milho. Polim.: Cienc. Tecnol. Antropova, T. V. Infrared Studies and Spectral Properties of
2000, 10, 24−30. Photochromic High Silica Glasses. Opt. Appl. 2014, 44 (2), 337−344.
(33) Versino, F.; García, M. A. Cassava (Manihot esculenta) Starch (53) Luchese, C. L.; Frick, J. M.; Patzer, V. L.; Spada, J. C.; Tessaro,
Films Reinforced with Natural Fibrous Filler. Ind. Crops Prod. 2014, I. C. Synthesis and Characterization of Biofilms Using Native and
58, 305−314. Modified Pinhão Starch. Food Hydrocolloids 2015, 45, 203−210.
(34) Schroder, K. E. Treatment of Diabetes Mellitus. Principles, (54) Tongdeesoontorn, W.; Mauer, L. J.; Wongruong, S.; Sriburi, P.;
Possibilities, Limits. ZFA (Stuttgart) 1976, 52 (24), 1213−9. Rachtanapun, P. Effect of Carboxymethyl Cellulose Concentration on
(35) Maeda, E. A.; Santos, A. F.; Silva, L. G. A.; Schön, C. G. Physical Properties of Biodegradable Cassava Starch-Based Films.
Chemical, Physical, and Mechanical Properties Evolution in Electron Chem. Cent. J. 2011, 5 (1), 6.
Beam Irradiated Isotactic Polypropylene. Mater. Chem. Phys. 2016, (55) de Paula, P. D.; Campello, E. F. C.; Guerra, J. G. M.; Santos, G.
169, 55−61. D. A.; Resende, A. S. d. Decomposiçaõ das Podas das Leguminosas
(36) Podshivalov, A.; Zakharova, M.; Glazacheva, E.; Uspenskaya, Arbóreas Gliricidia Sepium e Acacia Angustissima em um Sistema
M. Gelatin/Potato Starch Edible Biocomposite Films: Correlation Agroflorestal. Ciencia Florestal 2015, 25 (3), 10.
between Morphology and Physical Properties. Carbohydr. Polym. (56) Capeletti, L. B.; de Oliveira, L. F.; Goncalves, K. d. A.; de
2017, 157, 1162−1172. Oliveira, J. F.; Saito, A.; Kobarg, J.; dos Santos, J. H.; Cardoso, M. B.

2168 https://dx.doi.org/10.1021/acsapm.0c00124
ACS Appl. Polym. Mater. 2020, 2, 2160−2169
ACS Applied Polymer Materials pubs.acs.org/acsapm Article

Tailored Silica-Antibiotic Nanoparticles: Overcoming Bacterial


Resistance with Low Cytotoxicity. Langmuir 2014, 30 (25), 7456−64.
(57) Derbalah, A.; Shenashen, M.; Hamza, A.; Mohamed, A.; El
Safty, S. Antifungal Activity of Fabricated Mesoporous Silica
Nanoparticles Against Early Blight of Tomato. Egypt. J. Basic Applied
Sci. 2018, 5 (2), 145−150.
(58) Bi, Y.; Tian, S. P.; Guo, Y. R.; Ge, Y. H.; Qin, G. Z. Sodium
Silicate Reduces Postharvest Decay on Hami Melons: Induced
Resistance and Fungistatic Effects. Plant Dis. 2006, 90 (3), 279−283.
(59) Ellepola, A. N.; Khan, Z. U.; Chandy, R.; Philip, L. A
Comparison of the Antifungal Activity of Herbal Toothpastes Against
other Brands of Toothpastes on Clinical Isolates of Candida albicans
and Candida Dubliniensis. Med. Princ. Pract. 2011, 20 (2), 112−7.

2169 https://dx.doi.org/10.1021/acsapm.0c00124
ACS Appl. Polym. Mater. 2020, 2, 2160−2169

You might also like