Buscar

Ultrasound-assisted extraction of starch nanoparticles from breadfruit (Artocarpus altilis (Parkinson) Fosberg)

Esta é uma pré-visualização de arquivo. Entre para ver o arquivo original

Contents lists available at ScienceDirect
Colloids and Surfaces A
journal homepage: www.elsevier.com/locate/colsurfa
Ultrasound-assisted extraction of starch nanoparticles from breadfruit
(Artocarpus altilis (Parkinson) Fosberg)
Ivo H.P. Andradea, Caio G. Otonib, Thaís S. Amorima, Geany P. Camillotoc, Renato S. Cruzc,*
aGraduate Program in Food Science, School of Pharmacy, Federal University of Bahia, Rua Barão de Jeremoabo, s/n, Salvador, BA, 40170-115, Brazil
b Institute of Chemistry, University of Campinas, Rua Josué de Castro, 126, Campinas, SP, 13083-861, Brazil
c Department of Technology, State University of Feira de Santana, Av. Transnordestina, s/n, Feira de Santana, BA, 44036-900, Brazil
A R T I C L E I N F O
Keywords:
Biocolloid
Nanostarch
Mechanochemistry
Ultrasonic homogenization
Cavitation
A B S T R A C T
In this contribution, we report on the characterization of starch nanoparticles (SNP) extracted from breadfruit by
high-intensity ultrasound. These biocolloids were characterized regarding their spectroscopic, morphological,
rheological, and thermal properties in addition to ζ potential, light transmission, and X-ray diffraction (XRD)
measurements. After submitting a 0.5 % (w/v) aqueous suspension to 75min of ultrasound treatment, particles
of ca. 145 nm in diameter were obtained. The relatively low ζ potential (ca. −17mV) did not prevent SNP
aggregation. This was corroborated by transmission electron microscopy images. The viscosity of SNP suspen-
sion (2.7mPa s) was lower than that of native breadfruit starch (4.2mPa s). Thermogravimetry indicated that
SNP were more thermally unstable than native starch. Infrared spectroscopy indicated that CeO groups were the
most weakened by the ultrasound treatment. XRD revealed the rupture of the crystalline structure of native
starch, providing SNP with an amorphous character. Altogether, this set of characterization techniques de-
monstrates the feasibility of producing SNP in a rapid fashion and in the absence of chemical modifications,
characteristics that are in line with the contemporary trends towards greener products and processes. The SNP
produced herein through a mechanochemical approach have enormous potential for colloidal-related applica-
tions, including food, cosmetic, pharmaceutical, and biomedical systems.
1. Introduction
Starch is an abundant naturally occurring, rapidly renewable, and
biodegradable polysaccharide [1–4]. Starch is produced during photo-
synthesis as semi-crystalline granules ranging in diameter from 1 to
100 μm [5,6]. Starch granules are hierarchically structured into alter-
nate and concentric crystalline and amorphous layers [2]. Corn, potato,
wheat, and cassava stand out as the most common sources of starch [7].
Breadfruit (Artocarpus altilis (Parkinson) Fosberg) – a fruit originated
from Indonesia and Malaysia and favorably grown in warm and humid
regions – denotes an alternative amylaceous source because of its high
starch contents, typically between 53 and 76 % [8].
In addition to micro-sized granules, starch may fall within the col-
loidal regime after downsizing native granules into starch nanoparticles
(SNP). SNP feature the intrinsic characteristics of native starch (i.e.,
biodegradability, non-toxicity, and biocompatibility) as well as the
advantage of having large specific surface areas (surface area-to-weight
ratio), the latter providing SNP with unique properties such as
increased adsorption capacity and decreased diffusion-related limita-
tions. Furthermore, the production of SNP often involves low operating
costs [9–11]. Commercial applications of SNP include – but are not
limited to – pharmaceuticals, for controlled drug release [12,13];
dentistry, in formulations of materials with improved physical proper-
ties [14]; polymer-based materials, for mechanical reinforcement pur-
poses [15,16]; and foods [17].
The extraction of SNP from native granules typically involves the
combination of high energy inputs and complex chemical modifica-
tions. Ultrasound, however, is a top-down method that allows obtaining
SNP with high yields, in short periods, and without the need for addi-
tional chemical treatments, e.g. acid hydrolysis, which requires long
processing periods and provides low yields [18]. Ultrasound is a top-
down isolation approach of SNP and conveys energy through acoustic
waves as frequencies typically higher than the human threshold of
hearing, i.e. above 20 kHz. Ultrasonic waves have been demonstrated to
cause physicochemical transformations in matter by means of acoustic
cavitation, which is the collapse of bubbles of gases that propagate as a
https://doi.org/10.1016/j.colsurfa.2019.124277
Received 10 October 2019; Received in revised form 22 November 2019; Accepted 24 November 2019
⁎ Corresponding author.
E-mail addresses: ivo_henriquee@hotmail.com (I.H.P. Andrade), cgotoni@gmail.com (C.G. Otoni), thaisouzamorim@gmail.com (T.S. Amorim),
geanyperuch@yahoo.com.br (G.P. Camilloto), cruz.rs@uefs.br (R.S. Cruz).
Colloids and Surfaces A 586 (2020) 124277
Available online 25 November 2019
0927-7757/ © 2019 Elsevier B.V. All rights reserved.
T
http://www.sciencedirect.com/science/journal/09277757
https://www.elsevier.com/locate/colsurfa
https://doi.org/10.1016/j.colsurfa.2019.124277
https://doi.org/10.1016/j.colsurfa.2019.124277
mailto:ivo_henriquee@hotmail.com
mailto:cgotoni@gmail.com
mailto:thaisouzamorim@gmail.com
mailto:geanyperuch@yahoo.com.br
mailto:cruz.rs@uefs.br
https://doi.org/10.1016/j.colsurfa.2019.124277
http://crossmark.crossref.org/dialog/?doi=10.1016/j.colsurfa.2019.124277&domain=pdf
sound wave passes in a suspension medium [14,19,20].
To date, no previous study has focused on the exploitation of
breadfruit as a botanical source of SNP. In this sense, this contribution
set out not only to debut the production and characterization of SNP
from breadfruit, but also to shed light on a mechanochemical strategy
for achieving such an outcome through an efficient and sustainable
protocol. An extensive characterization of this biocolloid was aimed in
order to allow establishing correlations among its chemical structure,
extraction protocol, and final properties in an effort to pave the route
for commercial applications of SNP.
2. Materials and methods
2.1. Source of biological macromolecule
Breadfruit starch of the apyrena variety (50 ± 5 % amylose con-
tent; 1.47 ± 0.03 (g/ml) density; 0.030 ± 0.001 % phosphorus con-
tent), extracted according to Adebowale et al. [21], was kindly pro-
vided by the State University of Feira de Santana (UEFS).
2.2. Preparation of breadfruit starch nanoparticles
Breadfruit SNP were prepared according with Bel Haaj et al. [22],
with some modifications. Briefly, 40ml of aqueous breadfruit starch
suspensions at 0.5 % (w/v) were submitted to ultrasound treatment at a
frequency of 20 KHz using a 50-W probe ultrasonic processor, model
Q55 (QSonica, LLC, USA) operating at 1.0 % power for 75min. Ultra-
sound treated SNP were either kept in suspension or frozen and freeze-
dried. The dried SNP were used for thermogravimetry, X-ray diffrac-
tion, and infrared spectroscopy analyses.
2.3. Particle size and ζ potential
Never-dried SNP suspensions at 0.5 % were analyzed by dynamic
light scattering on a Zetasizer Nano ZS (Malvern Instruments Ltd., UK)
at 25 °C using a light dispersion angle of 173°. From these data, average
particle size and particle size distribution, the latter indicated by
polydispersity index (PDI), were calculated. The same equipment was
used for electrophoretic mobility measurements to determine ζ poten-
tial. All analyses were carried out in triplicates.
2.4. Light transmission
The transmittances of the SNP suspensions were measured at wa-
velengths ranging from 350 to 550 nm on a ultraviolet-visible spec-
trophotometer, model SP-220 (Biospectro, Brazil). This analysis was
performed 24 h after preparation and rest at room temperature.
2.5. Morphology
Native starch granules were imaged by scanning electron micro-
scopy (SEM) on a JSM-6610LV (JEOL, Japan) microscope operating at
30 kV. Samples were previously dried and coated by a gold layer for ca.
2min on a Desk V (Denton Vacuum, USA) thin film deposition system.
SNP were imaged through transmission electron microscopy (TEM) on a
JEM-1230 (JEOL, Japan) operating at 80 kV. Prior to TEM imaging,
SNP were diluted up to 1:500 (volume) in distilled water and a drop of
the diluted suspension was deposited onto a copper grid, stained with a
2 % aqueous dispersion of uranyl acetate and finally dried at room
temperature for 24 h.
2.6. Rheological measurements
The steady shear viscosities of the native starch and the breadfruit
SNP suspensions (same suspensions used for the ultrasound treatment)
were determined at 25 °C on a DV-II + Pro viscometer (Brookfield,
USA) equipped with a #61 spindle rotating at 150 rpm. The measure-
ments were repeated five times.
2.7. Thermogravimetry
Native starch and SNP samples of approximately 5mg were con-
tinuously weighed while being heated from 25 to 600 °C at 10 °C/min
on a thermal analyzer, model Pyris 1 TGA (PerkinElmer, USA), to in-
vestigate the thermal degradation profile by means of thermogravi-
metric (TG) and derivative TG (dTG) curves.
2.8. Attenuated total reflectance Fourier-transform infrared spectroscopy
(ATR-FTIR)
Infrared spectra of freeze-dried native starch and SNP were obtained
on a vibration spectrometer, model Paragon 1000 (PerkinElmer, USA)
operating in the infrared region with Fourier transform and equipped
with attenuated total reflectance (ATR) module. Spectra were acquired
at wavenumbers between 4000 and 500 cm−1 using a resolution of
2 cm−1.
2.9. X-ray diffraction (XRD)
Freeze-dried native starch and SNP samples were analyzed on an X-
ray diffractometer, model 6000 (Shimadzu Co.) with a CuKα radiation,
and tension and current were 30 kV and 30mA. XRD patterns were
recorded at Bragg angles (2θ) varying from 4 to 35° at 2.0° min−1.
2.10. Statistical analysis
SNP suspensions were obtained in three repetitions. Particle size, ζ
potential, PDI, and viscosity values are reported as mean values ±
standard deviations. Particle size and steady shear viscosity were fur-
ther submitted to analysis of variance (ANOVA) at a significance level
of 5 %, followed by t-test at the same level. The OriginPro software,
version 8.0724 (OriginLab, USA) was used for the statistical treatment
of the obtained data.
3. Results and discussion
3.1. Size and stability of SNP
The obtained SNP presented a polymodal size distribution, as shown
in Fig. 1. The first population of particles (from left to right) shows a
size distribution range from 8.7–37.8 nm and corresponds to 35.2 % of
the total population. The second population corresponds to
43.8–64.0 nm and 23.3 % of the population. Finally, the third
Fig. 1. Particle size distribution of breadfruit SNP.
I.H.P. Andrade, et al. Colloids and Surfaces A 586 (2020) 124277
2
population ranges in diameter from 190.0–955.0 nm and comprises
38.1 % of the particles, therefore showing the highest intensity. A
fourth population was also observed, with particles showing an average
size of 4,836.7 nm, but this represents only 3.4 % of the total particles.
Hebeish et al. [23] pointed out that nanoparticles are likely to ag-
glomerate in an aqueous medium during size measurements. Conse-
quently, the measured size represents that of the particle aggregates
instead of the individual particles. Either particle aggregation or any
possible contamination could justify the fourth population. However,
the herein obtained SNP had an average diameter of 146 ± 51 nm.
The studied SNP presented a PDI value of 0.46 ± 0.06. PDI is a
measurement of the width of the distribution of particle sizes within a
population. It ranges from 0 to 1, values close to zero indicating a
homogeneous distribution, whereas values greater than 0.5 indicate
high heterogeneity [23,24]. As mentioned before, the SNP produced in
this study ranged in size from 43 to 955 nm and presented three main
populations, each of which comprised between 20 and 40 % of the total
population, therefore denoting a nonhomogeneous distribution and
justifying the encountered PDI value.
Bel Haaj et al. [22] pointed out that too intense ultrasound treat-
ments, i.e. above 80 % power, the strong cavitation tends to increase the
coalescence between the bubbles, which in turn decreases the me-
chanochemical effects provided by such a treatment. In other studies,
only part of the total power was actually used [25–28]. However, in our
preliminary studies, we could only obtain a translucent suspension –
characteristic of nanoparticle suspensions - using 1.0 % power.
Other parameters that can interfere in the size of nanoparticles
obtained through ultrasound-assisted procedures are processing time
[29] and the size of the native starch granules, which depends on the
botanical origin [30]. Bravo et al. [31] produced maize SNP larger than
400 nm. Silva et al. [28] obtained cassava SNP with 93.1 % of the
particles having an average diameter of 77.51 nm. In both studies, the
ultrasound apparatus and the sonication time and power were identical
to those used in the present study (75min and 50W). Even using the
same amylaceous source, different sizes of SNP can be obtained. For
instance, Bel Haaj et al. [32] as well as Bravo et al. [31] produced maize
SNP with average diameters of 37 and>400 nm using ultrasound
powers of 400 and 50W, respectively.
Concerning particle stability in suspension, the average ζ potential
of the produced SNP was -16.9 ± 0.5mV. High nominal ζ potential
values (higher than +30mV or lower than −30mV) indicate stable
systems due to electrostatic repulsion between charged groups, which
prevent particle aggregation and other destabilization mechanisms
[33]. The negative surface charge of SNP arises from hydroxyl radicals
that are deprotonated in aqueous media [18,23]. In addition to the
negative character, the obtained ζ potential value indicates insufficient
electrostatic repulsion among SNP, possibly leading to aggregation
[32]. This observation is in line with DLS data, which suggest SNP
aggregation. Similar outcomes with regards to SNP stability were ob-
served in other investigations on ultrasound-assisted SNP extraction.
Bravo et al. [31] produced cassava and maize SNP with ζ potential
values equal to -8.70 and -1.90 mV, respectively. Silva et al. [28] ob-
tained cassava SNP with a ζ potential of -8.67mV.
3.2. Morphology of SNP
SEM and TEM images of breadfruit native starch and SNP, respec-
tively, are presented in Figs. 2 and 3. Native starch granules were
mostly quasi-spherical, but few polyhedral granules with irregular
surfaces were also observed (Fig. 2A and B). Rincón and Padilla [34]
also found these characteristics in breadfruit starch granules and at-
tributed such irregularities to possible indentations caused by the
compression of small starch granules during their development and
extraction. Indentation arises from endosperm volume reduction and
wrinkling during plant maturation [35].
Breadfruit SNP also presented a round shape (Fig. 3A and B), with
some stretched, quasi-cylindrical exceptions (Fig. 3C and D). These
structures were similar to those obtained by Wei et al. [36]. The pre-
sence of some SNP clusters (as in Fig. 3E) corroborates the aggregation
allowed by the relatively low ζ potential value. García et al. [37] and
Lamanna et al. [38] obtained similarly aggregated structures and at-
tributed it to the abundance of hydroxyl groups on SNP surface,
creating strongly associate interactions that, in addition to a low surface
charge, favor inter-SNP association. Finally, the darker SNP may be
related to particle overlapping during fragmentation, as suggested by
Hebeish et al. [23].
3.3. Light transmission through SNP suspensions
Pictures of SNP suspensions are presented in Fig. 4. Changes in the
turbidity of SNP suspensions were clearly observed when comparing
samples sonicated from 15 to 25min,
with suspensions becoming vir-
tually transparent at the end of this interval – indeed, after 25min of
sonication, the average SNP diameter was 228 ± 71 nm, i.e., smaller
than the wavelength of visible light, avoiding its scattering. Because we
aimed to evaluate the influence of sonication time on suspension
transparency, different sonication periods were also evaluated
(Fig. 4A).
As mentioned before, light scattering is directly related to the size of
particles in suspension. When aiming light beams directly at colloidal
suspension, small particles – as in the case of this study – are compar-
able in size with the lengths of the incident waves. The scattering is
observed when each particle in the light path behaves as a source of
secondary light [22,39]. As colloidal particles are known to scatter
visible radiation, the path of the beam that is passing through the
Fig. 2. SEM micrographs of native breadfruit starch (A, 1000X; B, 3000X).
I.H.P. Andrade, et al. Colloids and Surfaces A 586 (2020) 124277
3
suspension can be seen by naked eye. This phenomenon, so called
Tyndall effect [40], was observed in the SNP suspensions upon aiming a
light beam towards them. Suspensions that experienced shorter soni-
cation times presented a lower transmittance as well as a higher dis-
persed light intensity (Fig. 4B and C).
According to Craig et al. [39], larger particles scatter more light.
Thus, one can establish a direct relationship between sonication time,
particle diameter, and light transmission. Suspension transparency
therefore denotes a parameter that confirms particle fragmentation up
to the nanoscale, which was also observed by Boufi et al. [41]. Light
scattering by SNP was also observed when aiming the light beam at the
suspension sonicated for 75min (Fig. 4D, left). In the case of water
(Fig. 4D, right), this phenomenon was not observed because of the
absence of colloidal particles in suspension.
The transmittance curves (Fig. 5) clearly corroborate the direct re-
lationship among suspension transparency and sonication time, with
75min of ultrasound treatment leading to a 96.6 % transmittance at
550 nm. It is worth mentioning that SNP concentration in all suspen-
sions was the same. A remarkable change in transmittance at 550 nm
was observed among suspensions sonicated for 15 (14.6 %), 20 (71.0
%), and 25min (95.8 %).
3.4. Viscosity of SNP suspensions
The steady shear viscosity of breadfruit starch suspension was
4.2 ± 0.3mPa s, whereas for SNP suspension at the same solid content
this parameter was 2.7 ± 0.1mPa s. The lower viscosity of the SNP
suspension is caused by the reduced molecular weight of the granule
due to the breakdown of the glycosidic bonds caused by the ultrasonic
waves [42]. Reduced particle size is related to a lower capacity of water
retention by the swollen granules. This leads to a lower viscosity of SNP
suspensions in comparison with its native counterpart [42,43]. Other
studies have also reported a reduction in the viscosity of starch sus-
pensions submitted to ultrasound [19,29].
The decreased size of breadfruit SNP as well as the reduced viscosity
of SNP suspensions make them potential candidates as binders in pa-
permaking, to increase the binding capacity among starch and cellu-
lose. Decreased viscosity is also beneficial for nanocomposites from the
processing standpoint, wherein solid contents in suspensions should be
as high as possible in order to avoid excessive dilution [44].
3.5. Thermal properties of SNP
The thermogravimetric (TG) and derivative TG (dTG) curves of
breadfruit native starch and SNP are shown in Fig. 6. The TG curves of
both starch particles presented the same pattern, including two stages
of weight loss, behavior which were also reported by Bravo et al. [31],
Lamanna et al. [38], and Le Corre et al. [45]. According to Hornung
et al. [46], the first weight loss is related to water evaporation and
release of volatile compounds. The second weight loss stage is attrib-
uted to the degradation of the organic matter.
The dTG curves allow for a better visualization of the weight loss
events through their decomposition rate as a function of temperature.
The first event in native starch was as evidenced by the peak centered at
65.41 °C, being the offset of this weight loss stage at 140.78 °C. SNP, in
turn, had the maximum weight loss rate at 50.56 °C, and the process
continued up to 127.75 °C. The second event was observed between
311.86 and 412.73 °C (with the maximum decomposition rate occur-
ring at 371.96 °C) in native starch and from 294.88 to 426.79 °C
(maximum weight loss rate at 369.65 °C) for SNP.
Native starch was hence demonstrated to have a higher thermal
stability than SNP. A similar behavior has also been observed for SNP
elsewhere [28,38,47]. The higher thermal stability of native starch
compared to SNP can also be observed by comparing their rates of
maximum thermal decomposition, with native starch presenting a rate
(in modulus) of 0.83mg/min while SNP 0.62mg/min. Native starch has
more compact surface and inner structures than the nanoparticles de-
rived from it. Denser structures require more energy for the phase
Fig. 3. TEM micrographs of SNP (A, B, C, & D, scale bar: 200 nm; E, scale bar: 100 nm).
I.H.P. Andrade, et al. Colloids and Surfaces A 586 (2020) 124277
4
transitions and thermal decompositions [48]. Due to its lower degree of
compaction, SNP have a high number of OH groups on its surface,
through which thermal degradation occurs in advance due to the higher
hydrophilicity. Thus, weight losses generally occur earlier in SNP when
compared to native starch [38,49].
3.6. Spectroscopic characteristics of SNP
FTIR spectra of native starch and breadfruit SNP (Fig. 7) were re-
markably similar. This behavior was also documented in other studies
on nanostarch production [9,50]. Some bands in SNP spectrum pre-
sented different intensities in relation to native starch, in particular
those within the fingerprint region (below 800 cm−1 and between 800
and 1500 cm−1), where the characteristic peaks of starch are observed
[51].
Fig. 4. Tyndall effect on SNP suspensions.
Fig. 5. UV–vis Transmittance curves of SNP suspensions.
Fig. 6. TG (A) and dTG (B) of native starch and starch nanoparticles.
Fig. 7. Spectra of native starch and starch nanoparticles.
I.H.P. Andrade, et al. Colloids and Surfaces A 586 (2020) 124277
5
The most intense peak was observed at 1003 cm−1 and corresponds
to the stretching vibrations of CeO groups, as well as the peak at
1079 cm-1 [52]. Since the band 1003 cm−1 was the most accentuated of
the SNP, the CeO groups were the most affected by the ultrasonic
waves, which caused the weakening and rupture of the chemical con-
nections. The broad peak at 3290 cm−1 also presented accentuation in
relation to native starch, but in a low intensity. This band is related to
the stretching of the OeH hydrogen bonds. The region is characterized
by the interval between 3000 and 3600 cm-1 [53,54]. The accentuation
of such peak is related to the weakening of the hydrogen bond network,
probably due to a possible reduction of the crystalline structure of the
starch caused by the ultrasound [55]. Still in relation to the SNP
spectrum, the peak at 1350 cm-1 suffered a slight accentuation com-
pared to native starch. According to Pavlovic and Brandão [56], it is
related to the deformations in the CeOeH groups. Some specific peaks
that did not suffer alterations by the ultrasonic treatment were observed
in both spectra, such as the one at 1640 cm-1, attributed to water mo-
lecules adsorbed by the starch, at 2930 cm-1, related to the vibrations of
the CeH bonds [51].
The breakdown of starch glycosidic bonds is the factor responsible
for the formation of SNP. According to Chang et al. [20], the break-
down of covalent bonds of polymeric materials is caused by the emis-
sion of ultrasonic waves into aqueous suspension. Ultrasonic waves
emitted in starch suspensions may form air bubbles that collapse and
generate
shear forces on starch granules. This energy leads to the
breakdown of native starch chains as well as to fragmented particles,
potentially within the nanoscale [14,20].
3.7. Crystalline structure of SNP
The XRD patterns of native starch and breadfruit SNP are presented
in Fig. 8. It was observed that the native starch presented the char-
acteristic peaks of B-type starch at Bragg angles (2θ) close to 5.6°, 15°,
17°, 22°, and 24° [57]. The presence of the characteristic V-type peak
was also observed at ca. 20.6°, being associated with complexes formed
by simple amyloid propellers and lipids [58]. The B-type pattern is
typical of starches with high amylose contents (60–73 %), such as the
breadfruit starch studied here [30]. The results are in agreement with
recent studies on breadfruit starch, in which B-type crystalline patterns
were also observed [59,60].
In relation to the SNP diffractogram, the diffraction peaks dis-
appeared completely after the 75min of sonication, suggesting that the
energy input arising from the ultrasound waves disrupted the original
crystalline arrangement of starch. In other words, native starch un-
derwent an amorphization process when converted into SNP. Liu [61]
pointed out that the crystalline decomposition of starch can be attrib-
uted to the decrease in particle size. This, in turn, results in the con-
version of the ordered structure into fractions that are extremely small.
The deformation of the crystalline region of different starches by ul-
trasound treatment resulting in amorphous nanoparticles was also ob-
served elsewhere [22,28,32,40]. Bernardo et al. [62] also observed a
reduction in diffraction peaks and the appearance of fissures on the
surface of some starch granules after applying ultrasound. They at-
tributed this observation to cavitation-induced damage to the crystal-
line structure of yam starch. However, the energy was not sufficient to
modify the crystallinity pattern of starch, which remained unchanged.
When used for reinforcement purposes in polymer-based film ma-
trices, SNP and nanocrystals have different characteristics. Crystalline
particles feature organized polymer chains in ordered arrays, forming a
network that more effectively diffuses water and improves polymer
matrix stiffness when compared to amorphous SNP, thus denoting
better reinforcing agents [32]. Since amorphous particles do not feature
such a degree of organization, the SNP produced in this study are not
envisaged for reinforcement purposes in composite materials. Their
poor dispersion and propensity to agglomeration also lead to a low
reinforcing effect.
4. Conclusions
In summary, ultrasound was herein demonstrated as a purely phy-
sical means of isolating starch nanoparticles from breadfruit in a fast,
straightforward fashion. The produced SNP presented an average size of
145.65 nm after 75min of sonication. The surface charge in SNP was
not sufficient to prevent aggregation, which was further boosted by the
high occurrence of hydroxyl groups, leading to SNP aggregates. The
nanoparticles in suspension had a lower viscosity when compared to the
native starch due to their lower capacity of water retention and less
intermolecular interaction between the polymeric chains.
The thermogravimetry indicated a lower thermal stability of the
SNP compared to the native starch with a loss of volatiles and a de-
gradation of the organic matter at lower temperatures. The chemical
groups C–O were the most weakened by the ultrasonic waves during the
process. The analysis of the XRD indicated the rupture of the crystalline
region of the nanoparticles caused by the ultrasonic treatment and that
led to the formation of amorphous particles.
Thinking of future applications, the breadfruit SNP produced in this
study have potential for a range of colloidal-related systems, including
food formulations and as binders. Due to the small size of the breadfruit
SNP produced here and also because they were produced through a
purely physical procedure, they can directly contact foodstuffs without
major concern regarding toxicity, therefore opening the possibility of
being applied as fat mimetics – compounds that can mimic the sensory
and physical properties of triglycerides, though not totally replacing fat
in a gram-for-gram basis [63–66]. The combination of SNP with other
food components may lead to a cream-like blend featuring similar
properties to fats. An advantage of the use of SNP as mimetic fats is the
reduction of calories [14]. Another potential food application of
breadfruit SNP would be in shelf life extension of bakery and con-
fectionery products, as modified starches are known to slow down the
processes through which their moisture contents are reduced and their
hardness is increased, therefore leading to impaired palatability [67].
CRediT authorship contribution statement
Ivo H.P. Andrade: Conceptualization, Data curation, Writing -
original draft. Caio G. Otoni: Data curation, Writing - review & editing.
Thaís S. Amorim: Data curation. Geany P. Camilloto: Supervision,
Validation. Renato S. Cruz: Conceptualization, Methodology.
Declaration of Competing Interest
The authors declare that they have no known competing financial
interests or personal relationships that could have appeared to influ-
ence the work reported in this paper.Fig. 8. XRD patterns for native breadfruit starch and starch nanoparticles.
I.H.P. Andrade, et al. Colloids and Surfaces A 586 (2020) 124277
6
Acknowledgments
The authors are thankful to Coordenação de Aperfeiçoamento de
Pessoal de Nível Superior (CAPES), for granting the scholarship to
Andrade, I. H. P.; to the Electronic Microscopy Service of the Institute
Gonçalo Moniz (Fiocruz Bahia), for their support in SEM and TEM
analyses; to the Federal University of Bahia (UFBA), for the support in
thermal analyses; and to Dr. Loverice, M. V. at Embrapa
Instrumentação, for the support in XRD runs.
References
[1] Y. Farrag, W. Ide, B. Montero, M. Rico, S. Rodríguez-Llamazares, L. Barral,
R. Bouza, Preparation of starch nanoparticles loaded with quercetin using nano-
precipitation technique, Int. J. Biol. Macromol. 114 (2018) 426–433, https://doi.
org/10.1016/j.ijbiomac.2018.03.134.
[2] J.H. Kim, D.H. Park, J.Y. Kim, Effect of heat-moisture treatment under mildly acidic
condition on fragmentation of waxy maize starch granules into nanoparticles, Food
Hydrocoll. 63 (2017) 59–66, https://doi.org/10.1016/j.foodhyd.2016.08.018.
[3] R. Klimaviciute, J. Bendoraitiene, E. Lekniute, A. Zemaitaitis, Non-stoichiometric
complexes of cationic starch and 4-sulfophthalic acid and their flocculation effi-
ciency, Colloids Surf. A Physicochem. Eng. Asp. 457 (2014) 180–188, https://doi.
org/10.1016/j.colsurfa.2014.05.074.
[4] N.M.C. Silva, P.R.C. Correia, J.I. Druzian, F.M. Fakhouri, R.L.L. Fialho,
E.C.M.C. Albuquerque, PBAT/TPS composite films reinforced with starch nano-
particles produced by ultrasound, Int. J. Polym. Sci. (2017) 1–10, https://doi.org/
10.1155/2017/4308261.
[5] H.Y. Kim, J.A. Han, D.K. Kweon, J.D. Park, S.T. Lim, Effect of ultrasonic treatments
on nanoparticle preparation of acid-hydrolyzed waxy maize starch, Carbohydr.
Polym. 93 (2) (2013) 582–588, https://doi.org/10.1016/j.carbpol.2012.12.050.
[6] R.F. Tester, J. Karkalas, X. Qi, Starch-composition, fine structure and architecture,
J. Cereal Sci. 39 (2) (2004) 151–165, https://doi.org/10.1016/j.jcs.2003.12.001.
[7] J. Waterschoot, S.V. Gomand, E. Fierens, J.A. Delcour, Production, structure,
physicochemical and functional properties of maize, cassava, wheat, potato and rice
starches, Starch‐Stärke 67 (1-2) (2015) 14–29, https://doi.org/10.1002/star.
201300238.
[8] X. Wang, L. Chen, X. Li, F. Xie, H. Liu, L. Yu, Thermal and rheological properties of
breadfruit starch, J. Food Sci. 76 (1) (2011) E55–E61, https://doi.org/10.1111/j.
1750-3841.2010.01888.x.
[9] M.A. El-Sheikh, New technique in starch nanoparticles synthesis, Carbohydr.
Polym. 176 (2017) 214–219, https://doi.org/10.1016/j.carbpol.2017.08.033.
[10] O. Jeong, M. Shin, Preparation and stability of resistant starch nanoparticles, using
acid hydrolysis and cross-linking of waxy rice starch, Food Chem. 256 (2018)
77–84, https://doi.org/10.1016/j.foodchem.2018.02.098.
[11] C. Liu, Y. Qin, X. Li, Q. Sun, L. Xiong, Z. Liu, Preparation and characterization of
starch nanoparticles viaself-assembly at moderate temperature, Int. J. Biol.
Macromol. 84 (2016) 354–360, https://doi.org/10.1016/j.ijbiomac.2015.12.040.
[12] M.E. El-Naggar, M.H. El-Rafie, M.A. El-Sheikh, G.S. El-Feky, A. Hebeish, Synthesis,
characterization, release kinetics and toxicity profile of drug-loaded starch nano-
particles, Int. J. Biol. Macromol. 81 (2015) 718–729, https://doi.org/10.1016/j.
ijbiomac.2015.09.005.
[13] J.N. Putro, S. Ismadji, C. Gunarto, F.E. Soetaredjo, Y.H. Ju, Effect of natural and
synthetic surfactants on polysaccharide nanoparticles: Hydrophobic drug loading,
release, and cytotoxic studies, Colloids Surf. A Physicochem. Eng. Asp. (2019)
123618, , https://doi.org/10.1016/j.colsurfa.2019.123618.
[14] H.Y. Kim, S.S. Park, S.T. Lim, Preparation, characterization and utilization of starch
nanoparticles, Colloids Surf. B Biointerfaces 126 (2015) 607–620, https://doi.org/
10.1016/j.colsurfb.2014.11.011.
[15] M.C. Condés, M.C. Añón, A.N. Mauri, A. Dufresne, Amaranth protein films re-
inforced with maize starch nanocrystals, Food Hydrocoll. 47 (2015) 146–157,
https://doi.org/10.1016/j.foodhyd.2015.01.026.
[16] S. Jiang, C. Liu, X. Wang, L. Xiong, Q. Sun, Physicochemical properties of starch
nanocomposite films enhanced by self-assembled potato starch nanoparticles, LWT-
Food Sci. Technol. 69 (2016) 251–257, https://doi.org/10.1016/j.lwt.2016.01.053.
[17] Y. Tan, K. Xu, C. Niu, C. Liu, Y. Li, P. Wang, B.P. Binks, Triglyceride–water emul-
sions stabilised by starch-based nanoparticles, Food Hydrocoll. 36 (2014) 70–75,
https://doi.org/10.1016/j.foodhyd.2013.08.032.
[18] H. Angellier, L. Choisnard, S. Molina-Boisseau, P. Ozil, A. Dufresne, Optimization of
the preparation of aqueous suspensions of waxy maize starch nanocrystals using a
response surface methodology, Biomacromolecules 5 (4) (2004) 1545–1551,
https://doi.org/10.1021/bm049914u.
[19] F.O. Ogutu, T.H. Mu, R. Elahi, M. Zhang, H.N. Sun, Ultrasonic modification of se-
lected polysaccharides-review, J. Food Process. Technol. 6 (5) (2015) 1, https://doi.
org/10.4172/2157-7110.1000446.
[20] Y. Chang, X. Yan, Q. Wang, L. Ren, J. Tong, J. Zhou, High efficiency and low cost
preparation of size controlled starch nanoparticles through ultrasonic treatment and
precipitation, Food Chem. 227 (2017) 369–375, https://doi.org/10.1016/j.
foodchem.2017.01.111.
[21] K.O. Adebowale, B.I. Olu-Owolabi, E. kehinde Olawumi, O.S. Lawal, Functional
properties of native, physically and chemically modified breadfruit (Artocarpus
artilis) starch, Ind. Crops Prod. 21 (3) (2005) 343–351, https://doi.org/10.1016/j.
indcrop.2004.05.002.
[22] S. Bel Haaj, A. Magnin, C. Pétrier, S. Boufi, Starch nanoparticles formation via high
power ultrasonication, Carbohydr. Polym. 92 (2) (2013) 1625–1632, https://doi.
org/10.1016/j.carbpol.2012.11.022.
[23] A. Hebeish, M.H. El-Rafie, M.A. El-Sheikh, M.E. El-Naggar, Ultra-fine characteristics
of starch nanoparticles prepared using native starch with and without surfactant, J.
Inorg. Organomet. Polym. Mater. 24 (3) (2014) 515–524, https://doi.org/10.1007/
s10904-013-0004-x.
[24] M.R. Avadi, A.M.M. Sadeghi, N. Mohammadpour, S. Abedin, F. Atyabi,
R. Dinarvand, M. Rafiee-Tehrani, Preparation and characterization of insulin na-
noparticles using chitosan and Arabic gum with ionic gelation method, Nanomed.
Nanotechnol. Biol. Med. 6 (1) (2010) 58–63, https://doi.org/10.1016/j.nano.2009.
04.007.
[25] P.M. Gonçalves, C.P.Z. Noreña, N.P. da Silveira, A. Brandelli, Characterization of
starch nanoparticles obtained from Araucaria angustifolia seeds by acid hydrolysis
and ultrasound, LWT-Food Sci. Technol. 58 (1) (2014) 21–27, https://doi.org/10.
1016/j.lwt.2014.03.015.
[26] J.S. Santana, E.K.C. Costa, P.R. Rodrigues, P.R.C. Correia, R.S. Cruz, J.I. Druzian,
Morphological, barrier, and mechanical properties of cassava starch films re-
inforced with cellulose and starch nanoparticles, J. Appl. Polym. Sci. (2018) 47001,
, https://doi.org/10.1002/app.47001.
[27] J.S. Santana, E.K.C. Costa, N.M.C. Silva, J.I. Druzian, Influência do método de
obtenção de nanopartículas de amido nas propriedades resultantes, Paper presented
in Annals of the Congresso Brasileiro de Engenharia Química, Fortaleza/CE, 2016.
[28] L.P. Silva, C.C. Bonatto, F.D.E.S. Pereira, L.D. Silva, V.L. Albernaz, V.L.P. Polez,
Nanotecnologia verde para síntese de nanopartículas metálicas, Biotecnologia
Aplicada à Agro&indústria 4 Blucher, São Paulo, 2017, pp. 967–1012.
[29] C.O. Bernardo, J.L.R. Ascheri, C.W.P.D. Carvalho, Effect of ultrasound on the ex-
traction and modification of starches, Ciência Rural 46 (4) (2016) 739–746, https://
doi.org/10.1590/0103-8478cr20150156.
[30] D. Le Corre, J. Bras, A. Dufresne, Starch nanoparticles: a review,
Biomacromolecules 11 (5) (2010) 1139–1153, https://doi.org/10.1021/
bm901428y.
[31] A.M.J. Bravo, N.M.C. Silva, R.L. Fialho, E.C.M. Albuquerque, Caracterização de
nanopartículas de amido de mandioca e de milho pela técnica de ultrassom, Paper
presented in Annals of the Congresso Brasileiro de Engenharia Química, Fortaleza/CE,
2016.
[32] S. Bel Haaj, W. Thielemans, A. Magnin, S. Boufi, Starch nanocrystals and starch
nanoparticles from waxy maize as nanoreinforcement: a comparative study,
Carbohydr. Polym. 143 (2016) 310–317, https://doi.org/10.1016/j.carbpol.2016.
01.061.
[33] E.I. Benitez, J.E. Lozano, Influence of the soluble solids on the zeta potential of a
cloudy apple juice, Latin Am. Appl. Res. 36 (3) (2006) 163–168.
[34] A.M. Rincón, F.C. Padilla, Physicochemical properties of breadfruit (Artocarpus
altilis) starch from Margarita island, Venezuela, Archivos latinoamericanos de nu-
tricion 54 (4) (2004) 449–456.
[35] M.P. Benedetti, Granulometria do milho de textura dentada ou dura em rações para
frangos de corte (Master’s thesis), (2009).
[36] B. Wei, X. Xu, Z. Jin, Y. Tian, Surface chemical compositions and dispersity of starch
nanocrystals formed by sulfuric and hydrochloric acid hydrolysis, PLoS One 9 (2)
(2014) e86024, , https://doi.org/10.1371/journal.pone.0086024.
[37] N.L. García, L. Ribba, A. Dufresne, M.I. Aranguren, S. Goyanes, Physico‐mechanical
properties of biodegradable starch nanocomposites, Macromol. Mater. Eng. 294 (3)
(2009) 169–177, https://doi.org/10.1002/mame.200800271.
[38] M. Lamanna, N.J. Morales, N.L. García, S. Goyanes, Development and character-
ization of starch nanoparticles by gamma radiation: potential application as starch
matrix filler, Carbohydr. Polym. 97 (1) (2013) 90–97, https://doi.org/10.1016/j.
carbpol.2013.04.081.
[39] S.A. Craig, C.C. Maningat, P.A. Seib, R.C. Hoseney, Starch paste clarity, Cereal
Chemistry, (1989) (USA).
[40] D.A. Skoog, D.M. West, F.J. Holler, S.R. Crouch, Fundamentos de Química
Analítica, Editora Thomson, 2006 tradução da 8ª edição.
[41] S. Boufi, S.B. Haaj, A. Magnin, F. Pignon, M. Impéror-Clerc, G. Mortha, Ultrasonic
assisted production of starch nanoparticles: structural characterization and me-
chanism of disintegration, Ultrason. Sonochem. 41 (2018) 327–336, https://doi.
org/10.1016/j.ultsonch.2017.09.033.
[42] J.Y. Zuo, K. Knoerzer, R. Mawson, S. Kentish, M. Ashokkumar, The pasting prop-
erties of sonicated waxy rice starch suspensions, Ultrason. Sonochem. 16 (4) (2009)
462–468, https://doi.org/10.1016/j.ultsonch.2009.01.002.
[43] L. Kaur, J. Singh, O.J. McCarthy, H. Singh, Physico-chemical, rheological and
structural properties of fractionated potato starches, J. Food Eng. 82 (3) (2007)
383–394, https://doi.org/10.1016/j.jfoodeng.2007.02.059.
[44] S. Bloembergen, I. McLennan, D.I. Lee, J.V. Leeuwen, Paper binder performance
with biobased nanoparticles. A starch based biolatex can replace petroleum based
latex binders
in paper making, Paper 360° (2008) 46–48.
[45] D. Le Corre, J. Bras, A. Dufresne, Influence of native starch’s properties on starch
nanocrystals thermal properties, Carbohydr. Polym. 87 (1) (2012) 658–666,
https://doi.org/10.1016/j.carbpol.2011.08.042.
[46] P.S. Hornung, S. Avila, M. Lazzarotto, S.R. da Silveira Lazzarotto, G.L.D.A. de
Siqueira, E. Schnitzler, R.H. Ribani, Enhancement of the functional properties of
Dioscoreaceas native starches: mixture as a green modification process,
Thermochim. Acta 649 (2017) 31–40, https://doi.org/10.1016/j.tca.2017.01.006.
[47] M.S. Costa, Produção e caracterização estrutural, molecular e morfológica de na-
nocristais a partir de diferentes amidos e sua aplicação em biofilmes (Doctoral
dissertation), (2017).
[48] Q. Sun, G. Li, L. Dai, N. Ji, L. Xiong, Green preparation and characterisation of waxy
maize starch nanoparticles through enzymolysis and recrystallisation, Food Chem.
I.H.P. Andrade, et al. Colloids and Surfaces A 586 (2020) 124277
7
https://doi.org/10.1016/j.ijbiomac.2018.03.134
https://doi.org/10.1016/j.ijbiomac.2018.03.134
https://doi.org/10.1016/j.foodhyd.2016.08.018
https://doi.org/10.1016/j.colsurfa.2014.05.074
https://doi.org/10.1016/j.colsurfa.2014.05.074
https://doi.org/10.1155/2017/4308261
https://doi.org/10.1155/2017/4308261
https://doi.org/10.1016/j.carbpol.2012.12.050
https://doi.org/10.1016/j.jcs.2003.12.001
https://doi.org/10.1002/star.201300238
https://doi.org/10.1002/star.201300238
https://doi.org/10.1111/j.1750-3841.2010.01888.x
https://doi.org/10.1111/j.1750-3841.2010.01888.x
https://doi.org/10.1016/j.carbpol.2017.08.033
https://doi.org/10.1016/j.foodchem.2018.02.098
https://doi.org/10.1016/j.ijbiomac.2015.12.040
https://doi.org/10.1016/j.ijbiomac.2015.09.005
https://doi.org/10.1016/j.ijbiomac.2015.09.005
https://doi.org/10.1016/j.colsurfa.2019.123618
https://doi.org/10.1016/j.colsurfb.2014.11.011
https://doi.org/10.1016/j.colsurfb.2014.11.011
https://doi.org/10.1016/j.foodhyd.2015.01.026
https://doi.org/10.1016/j.lwt.2016.01.053
https://doi.org/10.1016/j.foodhyd.2013.08.032
https://doi.org/10.1021/bm049914u
https://doi.org/10.4172/2157-7110.1000446
https://doi.org/10.4172/2157-7110.1000446
https://doi.org/10.1016/j.foodchem.2017.01.111
https://doi.org/10.1016/j.foodchem.2017.01.111
https://doi.org/10.1016/j.indcrop.2004.05.002
https://doi.org/10.1016/j.indcrop.2004.05.002
https://doi.org/10.1016/j.carbpol.2012.11.022
https://doi.org/10.1016/j.carbpol.2012.11.022
https://doi.org/10.1007/s10904-013-0004-x
https://doi.org/10.1007/s10904-013-0004-x
https://doi.org/10.1016/j.nano.2009.04.007
https://doi.org/10.1016/j.nano.2009.04.007
https://doi.org/10.1016/j.lwt.2014.03.015
https://doi.org/10.1016/j.lwt.2014.03.015
https://doi.org/10.1002/app.47001
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0135
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0135
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0135
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0140
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0140
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0140
https://doi.org/10.1590/0103-8478cr20150156
https://doi.org/10.1590/0103-8478cr20150156
https://doi.org/10.1021/bm901428y
https://doi.org/10.1021/bm901428y
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0155
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0155
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0155
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0155
https://doi.org/10.1016/j.carbpol.2016.01.061
https://doi.org/10.1016/j.carbpol.2016.01.061
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0165
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0165
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0170
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0170
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0170
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0175
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0175
https://doi.org/10.1371/journal.pone.0086024
https://doi.org/10.1002/mame.200800271
https://doi.org/10.1016/j.carbpol.2013.04.081
https://doi.org/10.1016/j.carbpol.2013.04.081
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0195
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0195
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0200
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0200
https://doi.org/10.1016/j.ultsonch.2017.09.033
https://doi.org/10.1016/j.ultsonch.2017.09.033
https://doi.org/10.1016/j.ultsonch.2009.01.002
https://doi.org/10.1016/j.jfoodeng.2007.02.059
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0220
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0220
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0220
https://doi.org/10.1016/j.carbpol.2011.08.042
https://doi.org/10.1016/j.tca.2017.01.006
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0235
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0235
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0235
162 (2014) 223–228, https://doi.org/10.1016/j.foodchem.2014.04.068.
[49] N.L. García, M. Lamanna, N. D’Accorso, A. Dufresne, M. Aranguren, S. Goyanes,
Biodegradable materials from grafting of modified PLA onto starch nanocrystals,
Polym. Degrad. Stab. 97 (10) (2012) 2021–2026, https://doi.org/10.1016/j.
polymdegradstab.2012.03.032.
[50] L. Acevedo-Guevara, L. Nieto-Suaza, L.T. Sanchez, M.I. Pinzon, C.C. Villa,
Development of native and modified banana starch nanoparticles as vehicles for
curcumin, Int. J. Biol. Macromol. 111 (2018) 498–504 https://10.1016/j.ijbiomac.
2018.01.063.
[51] R.J. Estrada-León, V.M. Moo-Huchin, C.R. Ríos-Soberanis, D. Betancur-Ancona,
L.H. May-Hernández, F.A. Carrillo-Sánchez, J.M. Cervantes-Uc, E. Pérez-Pacheco,
The effect of isolation method on properties of parota (Enterolobium cyclocarpum)
starch, Food Hydrocoll. 57 (2016) 1–9, https://doi.org/10.1016/j.foodhyd.2016.
01.008.
[52] G. Zhou, Z. Luo, X. Fu, Preparation of starch nanoparticles in a water-in-ionic liquid
microemulsion system and their drug loading and releasing properties, J. Agric.
Food Chem. 62 (32) (2014) 8214–8220, https://doi.org/10.1021/jf5018725.
[53] M. Joshi, P. Aldred, S. McKnight, J.F. Panozzo, S. Kasapis, R. Adhikari, B. Adhikari,
Physicochemical and functional characteristics of lentil starch, Carbohydr. Polym.
92 (2) (2013) 1484–1496, https://doi.org/10.1016/j.carbpol.2012.10.035.
[54] B.S. Teixeira, R.H. Garcia, P.Y. Takinami, N.L. del Mastro, Comparison of gamma
radiation effects on natural corn and potato starches and modified cassava starch,
Radiat. Phys. Chem. 142 (2018) 44–49, https://doi.org/10.1016/j.radphyschem.
2017.09.001.
[55] Q. Sun, H. Fan, L. Xiong, Preparation and characterization of starch nanoparticles
through ultrasonic-assisted oxidation methods, Carbohydr. Polym. 106 (2014)
359–364, https://doi.org/10.1016/j.carbpol.2014.02.067.
[56] S. Pavlovic, P.R.G. Brandão, Adsorption of starch, amylose, amylopectin and glu-
cose monomer and their effect on the flotation of hematite and quartz, Miner. Eng.
16 (11) (2003) 1117–1122, https://doi.org/10.1016/j.mineng.2003.06.011.
[57] J. Zhu, L. Li, L. Chen, X. Li, Study on supramolecular structural changes of
ultrasonic treated potato starch granules, Food Hydrocoll. 29 (1) (2012) 116–122,
https://doi.org/10.1016/j.foodhyd.2012.02.004.
[58] E. Corradini, C. Lotti, E.S.D. Medeiros, A.J. Carvalho, A.A. Curvelo, L.H. Mattoso,
Comparative studies of corn thermoplastic starches with different amylose content,
Polmeros 15 (4) (2005) 268–273, https://doi.org/10.1590/S0104-
14282005000400011.
[59] O.F. Cavallini, Caracterização físico-química do amido e da farinha da fruta-pão
(Artocarpus altilis) e aplicação em pães de forma (Master’s thesis), (2015).
[60] L.M. Nwokocha, P.A. Williams, Comparative study of physicochemical properties of
breadfruit (Artocarpus altilis) and white yam starches, Carbohydr. Polym. 85 (2)
(2011) 294–302, https://doi.org/10.1016/j.carbpol.2011.01.050.
[61] D. Liu, Q. Wu, H. Chen, P.R. Chang, Transitional properties of starch colloid with
particle size reduction from micro-to nanometer, J. Colloid Interface Sci. 339 (1)
(2009) 117–124, https://doi.org/10.1016/j.jcis.2009.07.035.
[62] C.O. Bernardo, J.L.R. Ascheri, D.W.H. Chávez, C.W.P. Carvalho, Ultrasound assisted
extraction of yam (Dioscorea bulbífera) starch: effect on morphology and functional
properties, Starch‐Stärke 70 (5-6) (2018) 1700185, , https://doi.org/10.1002/star.
201700185.
[63] C.C. Akoh, Fat replacers, Food Technology, (1998) (USA).
[64] V.R. Diamantino, F.A. Beraldo, T.N. Sunakozawa, A.L.B. Penna, Effect of octenyl
succinylated waxy starch as a fat mimetic on texture, microstructure and physico-
chemical properties of Minas fresh cheese, LWT-Food Sci. Technol. 56 (2) (2014)
356–362, https://doi.org/10.1016/j.lwt.2013.12.001.
[65] J. Kaur, Cereal Starch Nanoparticles: Preparation, Characterization and Utilization
(Doctoral dissertation), Punjab Agricultural University, Ludhiana, 2016.
[66] W.R. Mason, Starch use in foods, Starch, Academic Press, 2009, pp. 745–795,
https://doi.org/10.1016/B978-0-12-746275-2.00020-3.
[67] M.M. Karaoğlu, H.G.R.Z. Kotancilar, İ. Çelik, Effects of utilization of modified
starches on the cake quality, Starch‐Stärke 53 (3-4) (2001) 162–169, https://doi.
org/10.1002/1521-379X(200104)53:3/4%3C162::AID-STAR162%3E3.0.CO;2-6.
I.H.P. Andrade, et al. Colloids and Surfaces A 586 (2020) 124277
8
https://doi.org/10.1016/j.foodchem.2014.04.068
https://doi.org/10.1016/j.polymdegradstab.2012.03.032
https://doi.org/10.1016/j.polymdegradstab.2012.03.032
https://10.1016/j.ijbiomac.2018.01.063
https://10.1016/j.ijbiomac.2018.01.063
https://doi.org/10.1016/j.foodhyd.2016.01.008
https://doi.org/10.1016/j.foodhyd.2016.01.008
https://doi.org/10.1021/jf5018725
https://doi.org/10.1016/j.carbpol.2012.10.035
https://doi.org/10.1016/j.radphyschem.2017.09.001
https://doi.org/10.1016/j.radphyschem.2017.09.001
https://doi.org/10.1016/j.carbpol.2014.02.067
https://doi.org/10.1016/j.mineng.2003.06.011
https://doi.org/10.1016/j.foodhyd.2012.02.004
https://doi.org/10.1590/S0104-14282005000400011
https://doi.org/10.1590/S0104-14282005000400011
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0295
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0295
https://doi.org/10.1016/j.carbpol.2011.01.050
https://doi.org/10.1016/j.jcis.2009.07.035
https://doi.org/10.1002/star.201700185
https://doi.org/10.1002/star.201700185
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0315
https://doi.org/10.1016/j.lwt.2013.12.001
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0325
http://refhub.elsevier.com/S0927-7757(19)31272-5/sbref0325
https://doi.org/10.1016/B978-0-12-746275-2.00020-3
https://doi.org/10.1002/1521-379X(200104)53:3/4%3C162::AID-STAR162%3E3.0.CO;2-6
https://doi.org/10.1002/1521-379X(200104)53:3/4%3C162::AID-STAR162%3E3.0.CO;2-6
	Ultrasound-assisted extraction of starch nanoparticles from breadfruit (Artocarpus altilis (Parkinson) Fosberg)
	Introduction
	Materials and methods
	Source of biological macromolecule
	Preparation of breadfruit starch nanoparticles
	Particle size and ζ potential
	Light transmission
	Morphology
	Rheological measurements
	Thermogravimetry
	Attenuated total reflectance Fourier-transform infrared spectroscopy (ATR-FTIR)
	X-ray diffraction (XRD)
	Statistical analysis
	Results and discussion
	Size and stability of SNP
	Morphology of SNP
	Light transmission through SNP suspensions
	Viscosity of SNP suspensions
	Thermal properties of SNP
	Spectroscopic characteristics of SNP
	Crystalline structure of SNP
	Conclusions
	CRediT authorship contribution statement
	mk:H1_23
	Acknowledgments
	References

Teste o Premium para desbloquear

Aproveite todos os benefícios por 3 dias sem pagar! 😉
Já tem cadastro?

Continue navegando