You are on page 1of 11

International Journal of Biological Macromolecules 168 (2021) 722–732

Contents lists available at ScienceDirect

International Journal of Biological Macromolecules

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

Xyloglucan-based hybrid nanocomposite with potential for


biomedical applications
Aiêrta Cristina Carrá da Silva a, Raimundo Rafael de Almeida b, Alexandre Carreira da Cruz Sousa a,c,
Fabián Nicolás Araneda Martínez d, Juliano Casagrande Denardin d,
Selene Maia de Morais e, Nágila Maria Pontes Silva Ricardo a,⁎
a
Laboratory of Polymers and Materials Innovation, Department of Organic and Inorganic Chemistry, Sciences Center, Federal University of Ceará, Campus of Pici, Zip Code 60440-760 Fortaleza,
CE, Brazil
b
Federal Institute of Education, Science and Technology of Ceará, Campus Camocim, Camocim, CE Zip Code 62400-000, Brazil
c
Federal Institute of Education, Science and Technology of Ceará, Campus Quixadá, Quixadá, CE, Zip Code 63902-580, Brazil
d
Department of Physics, University of Santiago and Cedenna, USACH-CEDENNA, Santiago Zip Code 9170124, Chile
e
Laboratory of Natural Products, Science and Technology Center, Ceará State University, Campus of Itaperi, Zip Code 60714-903 Fortaleza, CE, Brazil

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

Article history: Natural polymer-based hybrid nanocomposites have been proposed as one of the most promising tools for bio-
Received 24 July 2020 medical applications, including disease treatment and diagnosis procedures. Xyloglucan nanocapsules can simul-
Received in revised form 19 October 2020 taneously load magnetic iron oxide nanoparticles and bioactive for a specific tissue, reducing the processes of
Accepted 18 November 2020
degradation and metabolic inactivation of molecules with biological activity. In this work, magnetic nanocapsules
Available online 21 November 2020
of xyloglucan loaded with hydrophilic sulfated quercetin (MNXQ_SO3) were successfully synthesized by inverse
Keywords:
miniemulsion process through interfacial polymerization. The polymeric shell formation of nanocapsules was ev-
Magnetic nanocapsules of xyloglucan idenced by Fourier Transform Infrared spectroscopy and Transmission Electron Microscopy. The ferrofluid
Targeted drug delivery (Fe3O4@PAAS) incorporated into the xyloglucan nanocapsules was synthesized by hydrothermal method,
Biomedical applications using polyacrylic acid sodium salt as coating. Dynamic Light Scattering technique confirmed the nanomeric di-
mensions (202.3 nm) and the good colloidal stability (−40.2 mV) of MNXQ_SO3. The saturation magnetization
analyses pointed out the superparamagnetic behavior of Fe3O4@PAAS (48 emu/g) and MNXQ_SO3 (4.2 emu/g).
MNXQ_SO3 was able to modify the release profile of sulfated quercetin (67%) when compared to the free bioac-
tive (100%), exhibiting a release profile compatible with the zero-order kinetic model. The results showed that
the development of MNXQ_SO3 presents a new perspective for biomedical applications, including studies of
targeted drug delivery.
© 2020 Elsevier B.V. All rights reserved.

1. Introduction ferrofluid with superparamagnetic behavior, which can be employed


in biomedical applications [5]. The coating improves the dispersion
Nowadays, many researchers use nanotechnology to try to develop and chemical stability of magnetic nanoparticles, avoids its aggregation
novel nanocarriers that are selective drug delivery systems [1]. The and air oxidation, preventing the loss of magnetic properties [6,7].
drug encapsulation process promotes the protection of the bioactive Currently, the FDA (Food and Drug Administration) already ap-
from degradation in vivo, which allows its specific targeting to the dis- proved the use of magnetite (Fe3O4) as iron deficiency therapeutics
eased tissue [2]. Additionally, the incorporation of superparamagnetic (Feraheme®) and as magnetic resonance imaging (MRI) contrast
iron oxide nanoparticles (SPIONs) in these drug delivery systems, espe- agents (Feridex® and Gastromark®) [6]. Thus, the SPIONs and its asso-
cially magnetite (Fe3O4), has been an alternative to obtain nanosystems ciation between drug delivery systems allows their use in many
able to be guided by an external magnetic field to specific tissue [3,4]. theranostic applications, including trigger-controlled drug release and
The literature reports that polyacrylic acid sodium salt (PAAS) cov- magnetic hyperthermia for cancer treatment [8].
ered of magnetite nanoparticles (Fe3O4@PAAS) produces an aqueous In recent years, the scientific community has dispensed attention in
publications about the magnetic loading in polymeric nanocontainers.
In this context, there are nanocapsules, which can be based on biocom-
⁎ Corresponding author. patible polymers and emerge as an excellent nanocarrier for the admin-
E-mail address: naricard@ufc.br (N.M.P.S. Ricardo). istration of chemotherapeutic drugs [4]. The nanocapsules have a

https://doi.org/10.1016/j.ijbiomac.2020.11.128
0141-8130/© 2020 Elsevier B.V. All rights reserved.
A.C.C. da Silva, R.R. de Almeida, A.C. da Cruz Sousa et al. International Journal of Biological Macromolecules 168 (2021) 722–732

polymeric shell, a hydrophilic or hydrophobic core and can be obtained buffered saline (PBS, pH 7.4), deuterium oxide (99.9 atom % D) and poly-
for many procedures. However, the interfacial polymerization is usually acrylic acid sodium salt (PAAS) were obtained from Aldrich® and were
employed for the preparation of aqueous core nanocapsules through in- used as received. Polyglycerol polyricinoleate (Grindsted® PGPR super)
verse miniemulsion technique [9]. was donated from DuPont Danisco® (Brazil) and quercetin was acquired
The use of natural polymers (polysaccharides) on drug delivery sys- from PVP-Anonymous Society (Brazil). Distilled and ultrapure water
tems often occurs because they are biodegradable, biocompatible, (produced by Milli-Q Advantage A-10 system, Millipore®) were used
widely available, nontoxic and flexible to chemical modification [10]. in this research. Seeds of tamarind (Tamarindus indica L.) were obtained
In this research field, some authors published relevant works about from fruit pulp acquired at the local market. All the solvents and reagents
drug delivery vehicles based on polysaccharides, such as gum arabic, were analytical grade and used directly without further purification.
chitosan, and maltodextrin for curcumin delivery [11]; salecan for doxo-
rubicin delivery [12]; the self-assembly of anionic (salecan) and cationic 2.2. Experimental
(chitosan) polysaccharides into polyelectrolyte complex for release of
vitamim C [13] and green tea polyphenols [14]. 2.2.1. Extraction and molar mass determination of Tamarindus indica Linn.
In this sense, xyloglucan is also a natural polymer obtained from xyloglucan
tamarind (Tamarindus indica Linn.) seeds, a typical plant of tropical The seeds of tamarind fruit were manually separated from pulp,
region, which has been cultivated in Brazilian Northeast since the last washed with running water and stirred (400 rpm) with distilled
century. This polysaccharide is a neutral hemicellulose and contains a water at 100 °C during 30 min for enzyme inactivation and softening
β-(1 → 4)-linked D-glucan backbone chain, which is partly substituted of the seeds coats. After the separation of the cotyledons from seeds cov-
at O-6 position of D-glucopyranosyl residues with α-D-xylopyranose ering, tamarind kernels were frozen at −20 °C and dried using a freeze
or with 2-O-β-D-galactopyranosyl-α-D-xylopyranose [15,16]. Due to dryer (Model Martin Christ Alpha 1–2 LDplus) under the follow condi-
its mucoadhesive and in situ gelling properties, xyloglucan has been re- tions: pressure of 0.12 mbar and condenser temperature at −55 °C for
ported as an attractive and functional natural polymer in drug delivery 48 h. Xyloglucan was obtained through exhaustive aqueous extraction
assays [10]. of tamarind kernels, which was performed to the previously reported
Xyloglucan has been widely employed in the pharmaceutical field, method [22].
including the investigation of drug delivery via nasal, ocular, pulmonary, The aqueous xyloglucan solution was exposed to the same afore-
rectal and oral administration. Recently, a collection of works was pub- mentioned freeze drying process and the sample yield was calculated
lished about drug delivery systems based on xyloglucan, including and characterized by Nuclear Magnetic Resonance spectroscopy. Ther-
thermosensitive and thermoreversible in situ gel, nanoaggregates, nano- mal analysis was evaluated. The molar mass was estimated by Gel
spheres, films, tablets, patches, gel and hydrogel for biomedical applica- Permeation Chromatography (GPC), using a Shimadzu LC-10AD chro-
tions [17]. Therefore, this polysaccharide presents a great versatility and matograph with a refractive index detector (Model RID-10A) at 40 °C.
should be studied by the academy community. However, until now, no The chromatograph was equipped with an Ultrahydrogel linear column
polymeric nanocapsules based on xyloglucan have been synthesized. (7.8 mm × 300 mm) and used a flow rate of 1 mL/min. The analysis was
Among many bioactives, there is quercetin (3,3′,4′,5′-7-penta-hy- carried out with 20 μL of polysaccharide solution 0.1% (w/v), dissolved
droxy-flavone), a flavonoid obtained from fruits and vegetables, which in ultrapure water, and 0.1 mol/L of sodium nitrate was used as eluent
presents many biological properties, including antiviral, antioxidant, an- [23]. All the solutions were filtered in cellulose acetate membranes
tibacterial, gastroprotective, anti-inflammatory and anti-carcinogenic (Sigma-Aldrich, St. Louis, USA). A calibration curve of Pullulan was
activities. This polyphenol (C15H10O7, 302.2 g/ mol) is a yellow solid, used as standard, with different molecular weights (range: 5.9 × 103 a
which has a hydrophobic character with aqueous solubility of 0.5 g L−1 7.88 × 105 g/mol). The equation obtained from this calibration plot
at 25 °C [18–20]. Thus, quercetin presents a low biodisponibility and was: log Mw = 13.677–0.954 Ve, where Ve was the elution volume in
after its oral consumption, quercetin undergoes significant presystemic mL. The linear correlation coefficient was 0.99.
elimination, which blocks its use as chemotherapeutic agent. In addition,
quercetin showed in vitro toxic effects on normal human cell lines. 2.2.2. Synthesis and characterization of hydrophilic quercetin sulfate
Therefore, many synthetic compounds have been investigated in order (Q_SO3)
to develop a new quercetin derivative that also has similar biological Hydrophilic quercetin sulfate was obtained from adapted method by
properties to the original molecule and be able to replace this chemo- Nair & Bernstein, 1983 [24]. Briefly, 1.22 g of quercetin was stirred
therapeutic bioactive. The literature reports that hydrophilic quercetin (350 rpm) in 20 mL of N,N-dimethylformamide and 3.34 g of sulfur tri-
sulfate is one of the candidates to replace quercetin [21]. oxide pyridine complex was added at 65–70 °C for 5 h. The solution was
In the present study, we describe the synthesis by inverse cooled at room temperature and isopropyl alcohol was placed to form a
miniemulsion of novel magnetic hybrid nanocomposites based on brown precipitate which was separated by centrifugation. The brown
xyloglucan loaded with Fe3O4@PAAS and hydrophilic quercetin, which solid was washed with diethyl ether and dried using a desiccator.
was used as drug-model for the release study of this bioactive. In this Then the precipitate was dissolved in deionized water with 10 mL of so-
work we developed nanocarriers from natural and renewable sources, dium acetate 30% (w/v) using a magnetic stirring for 20 min. Finally,
which represents new perspectives for biomedical applications. 100 mL of isopropyl alcohol was added to separate the solid by decanta-
tion. The mixture was filtered, and the precipitate was washed with
2. Materials and methods diethyl ether again and dried. Then, Q_SO3 was analyzed by Fourier
Transform Infrared spectroscopy and Thermogravimetry.
2.1. Materials
2.2.3. Synthesis of magnetic nanocapsules based on xyloglucan loaded with
Ethanol (97%), acetone (99.5%), diethyl ether (98%), cyclohexane hydrophilic quercetin sulfate (MNXQ_SO3)
(99%), N,N-dimethylformamide (99%) and sodium acetate (99%) were The sample (MNXQ_SO3) was prepared in a water-in-oil
purchased from Synth® and used as received. Ammonium hydroxide miniemulsion system similarly to the method previously reported by
(28–30% of ammonia) was acquired from Vetec®. Iron (II) chloride Kang et al. [1]. The aqueous phase was formed by 4 mL of an aqueous so-
tetra-hydrate (FeCl2·4H2O, ≥99.0%), iron (III) chloride hexa-hydrate lution of xyloglucan 1% (w/v) containing 20 mg of Q_SO3 and 12 mg of
(FeCl3·6H2O, ≥99.0%) and isopropyl alcohol (99.5%) were purchased superparamagnetic nanoparticles (Fe3O4@PAAS). The organic phase
from Sigma-Aldrich®. Sulfur trioxide pyridine complex (98%), tolylene- contained 14.5 g of cyclohexane with a varying amounts of polyglycerol
2,4-diisocyanate (TDI), (95%), sodium dodecyl sulfate (99%), phosphate polyricinoleate (100, 120 or 160 mg) as surfactant and TDI (42.5 or

723
A.C.C. da Silva, R.R. de Almeida, A.C. da Cruz Sousa et al. International Journal of Biological Macromolecules 168 (2021) 722–732

Table 1 with 13 k transients, spectral window of 251.3 ppm, an acquisition


Experimental formulations for the synthesis of MNXQ_SO3 sample. time of 0.865 s and a time between each acquisition of 1 s [26].
Sample Organic phase

PGPR (mg) TDI (mg)


2.3.2. Infrared spectroscopy
The Fourier Transform Infrared spectra (FT-IR) were obtained with
S1 100 80
Shimadzu IRTracer-100 Spectrophotometer (Model 8300) using KBr
S2 120 80
S3 160 80 pellets in the wave number range of 400 and 4000 cm−1 [27]. The sam-
S4 100 42.5 ples spectra were recorded from 100 scans.
S5 120 42.5
S6 160 42.5 2.3.3. Dynamic light scattering (DLS) technique
The polydispersity index (PdI), particle size and zeta potential of
the sample were investigated using a Zetasizer Nano Series Malvern
W-ZS90. The samples were diluted 50 times with ultrapure water
80 mg) as crosslinker agent. Table 1 shows the composition of the or- and analyzed in an optical grade polystyrene cuvette at 25 °C [28].
ganic phase, which forms the samples S1–S6. The intensity-weighted hydrodynamic diameters and z-average
Initially, 8.0 g of organic phase was homogenized with the aqueous were based on three measurements, each one originated from 11
phase using a Benchmark™ vortex mixer (Model BV-1000). Then, this scans.
pre-emulsion was sonicated in an ultrasound (Branson Sonifier W-
450-Digital and a 1/2″ tip) at 70% amplitude for 3 min with a pulse of 2.3.4. Transmission electron microscopy (TEM)
20 s on and 10 s off (cooling the mixture with an ice bath). Afterwards, Transmission electron microscopy (TEM) images of MNXQ_SO3
4.5 g of organic phase was mixed with the miniemulsion again and sample (cyclohexane) were obtained using a Hitachi®HT7700 TEM sys-
placed at the ultrasound under the same conditions as for the last tem operating at an accelerating voltage of 120 kV. In order to perform
step. Then, tolylene-2,4-diisocyanate was added (42.5 or 80 mg) in 2 g the TEM investigation, the sample was prepared according to the meth-
of the oil phase and dripped slowly into the sample (Table 1). The inter- odology previously reported [29].
facial polymerization was carried out at 25 °C for 24 h, using a mechan-
ical stirring (Model RW 20 digital, IKA®) at 450 rpm. Finally, a 2.3.5. Thermal analysis
purification process was carried out: the MNXQ_SO3 nanocapsules dis- The thermal properties of the samples were evaluated by
persion in cyclohexane were centrifuged at 4000 rpm for 20 min, and thermogravimetry analysis (TGA). TGA analysis were conducted on a Si-
the precipitate was redispersed in the same amount of fresh cyclohex- multaneous Thermal Analyzer (PerkinElmer, Model STA 6000) instru-
ane. Afterwards, it was homogenized by vortex. This procedure was re- ment, using nitrogen atmosphere with flow of 50 mL/min, heating
peated three times. After drying, 2 mg of these samples (S1–S6, Table 1) rate of 10 °C/min and temperature range of 25–800 °C [26].
were investigated by Fourier Transform Infrared spectroscopy to con-
firm the crosslinking reaction. 2.3.6. X-ray powder diffraction (XRD)
To obtain the MNXQ_SO3 aqueous samples from S1 to S6, the puri- X-ray patterns diffraction of the synthesized sample was collected
fied nanocapsules dispersion in cyclohexane was added dropwise into using a PANalytical (X'pert Pro MPD) X-ray powder diffractometer op-
sodium dodecyl sulfate solution 0.1% (w/v) using a sonication bath erating in Bragg-Brentano reflection geometry with Co-Kα radiation
(Quimis®, Q335D) during 4 min. This dispersion was mechanical stirred (40 Kv, 40 mA) over the range 20° < 2θ < 100° and scan speed of
(450 rpm) overnight, in an open vial, to allow evaporation of cyclohex- 0.013°/s [3].
ane. The aqueous samples were analyzed by Dynamic light scattering
(DLS) technique. 2.3.7. Magnetic properties
After analyzing these data, the most appropriate formulation was The magnetic behavior of the samples was evaluated using a home-
used to continue this research. The selection of the best sample was made vibrating sample magnetometer (VSM) at 300 K in the magnetic
based on three parameters: the colloidal stability and the average parti- field range from −10 to 10 kOe [3]. In order to assure the magnetic mo-
cle size, which should be compatible with nanocarriers for biomedical ments values acquired, the equipment was previously calibrated using a
applications, as well as the Fourier Transform Infrared spectroscopy standard reference material (yttrium Iron Garnet Sphere) from the Na-
analysis, that was performed to assess the interfacial polymerization re- tional Institute of Standards and Technology (NIST). For all measure-
action. Then the purified nanocapsules dispersion in cyclohexane was ments, the magnetic moment obtained for each applied field was
analyzed by Transmission electron microscopy. Thermal Analysis and normalized by the mass of the samples.
Magnetic properties were performed using the powder obtained from
drying the nanocapsules dispersion in cyclohexane. 2.4. Release study and encapsulation efficiency of hydrophilic quercetin
The magnetic nanoparticles (Fe3O4@PAAS) used in this procedure sulfate
were synthesized using a hydrothermal method [25] and were investi-
gated by Transmission electron microscopy, Thermogravimetry, X-ray The release profile of quercetin sulfate into magnetic xyloglucan
powder diffraction and Magnetic properties. nanocapsules was investigated using a previously reported methodol-
ogy [3]. The experiment was conducted at 37 °C and the donor compart-
2.3. General characterization procedures ment of the system contained 1.5 mL of MNXQ_SO3 sample
(0.333 mg/mL of sulfate quercetin), furthermore the acceptor compart-
2.3.1. Nuclear magnetic resonance (NMR) spectroscopy ment was prepared with 50 mL of PBS (pH 7.4). The release assay was
The Nuclear magnetic ressonance (NMR) spectrum of the also carried out with an aqueous solution of active ([Q_SO3] =
xyloglucan sample was recorded in a 600 MHz Agilent DD2 spectrome- 0.333 mg/mL), which was placed in the donor compartment (1.5 mL)
ter equipped with an inverse detection 5-mm One-Probe (H-F/15N-31P) for data comparison. Aliquots (2 mL) were collected after 30 min of ex-
and Z-gradient coils. Xyloglucan was prepared by dissolving 10 mg of periment and at intervals of 1 h for the first 8 h of assay. Then, more al-
the material in 600 μL of D2O (deuterium oxide). For 1H NMR analysis iquots were collected each 24 h until completing 48 h of experiment.
was applied 32 scans, 32 k of points in the time dominium with a spec- The resulting aliquots were analyzed using a UV–Vis Spectrophotome-
tral window of 16.0 ppm, an acquisition time of 3.328 s and the time be- ter (ThermoScientific®, model: Genesis 10S) at 265 nm. The cumulative
tween each acquisition was 10 s. The 13C NMR spectrum was recorded percentage of dissolved Q_SO3 was calculated using a regression

724
A.C.C. da Silva, R.R. de Almeida, A.C. da Cruz Sousa et al. International Journal of Biological Macromolecules 168 (2021) 722–732

equation generated from the standard data. Usually, the most important determined by the difference between the initial amount of bioactive
mechanisms for the transport of drugs or bioactives from polymeric ma- added to the MNXQ_SO3 formulation and the free amount observed.
trices are diffusion, erosion and degradation. However, the in vitro data In this way, the encapsulation efficiency (%EE) was determined through
dissolution was fitted to the more important mathematical models the following equation:
(zero and first order, Higuchi and Korsmeyer-Peppas, which were de-
fined for Eqs. (1)–(4), respectively) to evaluate the kinetic mechanism
%EE ¼ ½ðB−AÞ=B  100
of quercetin sulfate release in the MNXQ_SO3.

Zero Order Model : Q t ¼ Kt ð1Þ where:

First Order Model : ln Q t ¼ Kt ð2Þ A = free amount of bioactive in MNXQ_SO3


B = initial amount of bioactive in MNXQ_SO3
Higguchi Model : Q t ¼ Kt1=2 ð3Þ
3. Results and discussion
Korsmeyer−Peppas : Q t ¼ Ktn ð4Þ
3.1. Extraction and characterization of Tamarindus indica L. xyloglucan
where Qt is the amount of drug/bioactive released at time t; K is the
release constant for the corresponding kinetic model and n is the re- Xyloglucan was efficiently obtained with extracted yield of 43.11 wt
lease exponent that specifies the drug/bioactive release. The litera- % (based on the dry cotyledon mass of T. indica seeds). The yield values
ture reports that the mechanism of release in Higguchi Model reported in the literature for xyloglucan of T. indica range from 10.1 to
involves a diffusion process (Fickian Model), while the mechanism 50% [15,33–35].
of release for Korsmeyer-Peppas Model can occur by diffusion, when The GPC chromatogram for xyloglucan of T. indica is shown in Fig. S1
n = 0.5; by anomalous transport (Non Fickian Model, which presents (Appendix A – Supplementary data). The number average molecular
both process: diffusion and polymeric chain relaxation), when weight (Mn) was 233 kDa, average molecular weight (Mw) was
0.5 < n < 1.0; and by case-II transport, when n = 1.0, indicating that 1227 kDa, higher average molecular weight (Mz) was 4362 kDa and
the mechanism of release is purely controlled by the polymeric chain polydispersity index was 5.26. The value of Mw for the xyloglucan sam-
relaxation [3,30–32]. ple was similar to other researches published in the literature.
The MNXQ_SO3 formulation was subjected to the magnetic separa- Periasamy et al. (2018) obtained xyloglucan of kernel powder from
tion for determination of the free bioactive compound amount [3]. The T. indica seeds with high molecular weight (Mw = 1331 kDa). This
aqueous fraction without magnetic properties was analyzed by UV– value depends on the source and on the extraction method of polysac-
Vis spectrometry (265 nm). The Q_SO3 entrapped in the sample was charide [31].

Fig. 1. 1H NMR spectrum of T. indica xyloglucan.

725
A.C.C. da Silva, R.R. de Almeida, A.C. da Cruz Sousa et al. International Journal of Biological Macromolecules 168 (2021) 722–732

Fig. 2. 13C NMR spectrum of T. indica xyloglucan.

The 1H NMR (Fig. 1) and 13C NMR (Fig. 2) analyses were performed The nanocapsule size of the samples measured by DLS is in the range
to investigate the xyloglucan structure of T. indica. The 1H spectrum of 202.3 to 552.0 nm. The literature reports that some parameters must
shows signals of β-D-glucopyranosyl (Glu) residues at 3.42 ppm and be within a range to be compatible with nanocarriers for biomedical
three signals in the anomeric region: at 5.13 ppm, which is compatible purposes. For example, zeta potential of dispersions above −30 mV
with 2-linked α-D-xylopyranosyl residues (Xyl 1); at 4.95 ppm corre- was considered to have good colloidal stability, while the PdI can
sponds to the terminal α-D-xylopyranosyl units (Xyl 2) and at range from 0 (monodisperse) to 1 (polydisperse) [41]. Moreover, the
4.57 ppm refers to an overlapped of β-glucopyranosyl and β- average particle size of nanocapsules must be in the range of 50 to
galactopyranosyl residues [15,31,33–36]. The ratio of the carbohydrate 500 nm [29].
residues for xyloglucan sample is calculated by the relative areas of It can be seen through Table 2 that the samples S1, S3 and S6 showed
the signal 1H (Glu), 1H (Gal) and 1H(Xyl), resulting in a molar ratio of an average particle size and a good colloidal stability compatible with
3:1:2 for the xyloglucan monomer Glu:Gal:Xyl [32]. As illustrated by the literature data. Therefore, the samples mentioned above were sub-
Fig. 1, the area under glucose (Glu), galactose ([Glu + Gal] – Glu) and jected to FT-IR analysis.
xylose (Xyl 1 + Xyl 2) provides a ratio of 2.83:0.95:1.99. These results The FT-IR analysis was performed to assess the crosslinking reaction.
are in accordance with other works previously published [33,37,38]. The absence of the band in 2275 cm−1 indicates that all the crosslinking
As verified in Fig. 2, the signals at 104.95, 102.93, 99.66 and agent was consumed during the reaction of nanocapsules surface for-
99.41 ppm of xyloglucan 13C NMR spectrum correspond to the anomeric mation [42]. Furthermore, a better biocompatibility of the final
carbons of β-D-galactopyranose (Gal), β-D-glucopyranose (Glu), and α- nanocapsules depends on reducing the amount of excess TDI [29].
D-xylopyranose (Xyl 2 and Xyl 1), respectively. The signal at 60.57 is Fig. 3 shows the FT-IR spectra from the samples S1, S3 and S6.
assigned to C-6 units of free β-D-galactopyranoses and the signal at Through Fig. 3 it was possible to verify that only the sample S6 did
61.71 ppm corresponds to C-5 of α-D-xylopyranose [39,40]. not present the band in 2275 cm−1. This means that all crosslinker

Table 2
3.2. General characterization procedures Obtained results of average size, zeta potential and polydispersity index.

3.2.1. Influence of surfactant (PGPR) and crosslinker agent (TDI) amounts Samplea Average particle size (nm) Zeta potential (mV) PdI

in the formulation obtained for MNXQ_SO3 samples (S1–S6) S1 422.7 −32.5 0.254
At this stage, the amounts of PGPR and TDI were varied (100, 120 or S2 552.0 −14.1 0.374
S3 515.5 −37.5 0.474
160 mg and 42.5 or 80 mg, respectively) to investigate and obtain the
S4 359.7 −23.3 0.348
most appropriated formulation of MNXQ_SO3. The synthesized S5 497.2 −12.4 0.596
nanocapsules in aqueous dispersion (SDS 0.1%) were characterized by S6 202.3 −40.2 0.417
DLS to determine the average nanocapsule size, the surface charge a
Samples S1, S2 and S3 were synthesized using: 80 mg of TDI; 100, 120 and 160 mg of
(zeta potential) and the size distribution values (polydispersity index PGPR, respectively. Samples S4, S5 and S6 were synthesized using: 42.5 mg of TDI; 100,
– PdI). These results are presented in Table 2. 120 and 160 mg of PGPR, respectively.

726
A.C.C. da Silva, R.R. de Almeida, A.C. da Cruz Sousa et al. International Journal of Biological Macromolecules 168 (2021) 722–732

Fig. 4b also shows infrared bands of Fe3O4@PAAS sample, which can


be assigned to the presence of polyacrylate sodium salt on the surface of
Fe3O4 (1400 and 1550 cm−1 represent symmetric and asymmetric
stretch of COO−, respectively; 1450 cm−1 is designated for bending vi-
bration of C\\H and 1650 cm−1 refers to C_O stretching vibration of
carboxylate) [5].
The FT-IR spectrum of xyloglucan (Fig. 4c) is in good agreement with
the one reported in the literature. The absorption at 3408 and
2922 cm−1 is assigned for stretching vibrations of O\\H and C\\H, re-
spectively [34]. Bands at region spectral between 1600 and 800 cm−1
exhibit highly coupled C–C–O, C–O–C and C–OH stretching modes of
polymer backbone [32]. The band at 1039 cm−1 is reported for C\\H
bond of the anomeric carbon (C1) and at 896 cm−1 indicates the α-
bond between D-xylopyranose units [43].
As shown in Fig. 4d, the FT-IR spectrum of Q_SO3 illustrates three
bands, which refer to the asymmetric stretching vibration of the S_O
bond (1050, 1144 and 1248 cm−1), indicating that the sulfated deriva-
tive of quercetin was successfully synthesized. These results are similar
to those published in the literature [44].

3.2.3. Dynamic light scattering (DLS) and transmission electron


microscopy (TEM)
The average particle size (202.3 ± 2.3 nm) and polydispersity index
(0.417 ± 0.045) of MNXQ_SO3 were obtained by DLS analysis. Recently,
gelatin capsules loaded with an aqueous dispersion of magnetic nano-
particles were produced by inverse miniemulsion and the DLS analysis
revealed an average size in the range of 500–700 nm [42]. On the
other hand, starch nanocapsules with encapsulated dsDNA produced
by miniemulsion technique had a size measured in the range of 320 to
920 nm and the polydispersity index varied from 0.26 to 0.36, indicating
a broad size distribution of the capsules [29].
Furthermore, the MNXQ_SO3 sample shows a good stability (zeta
potential value of −40.2 mV) and no precipitation or coagulation was
observed during three months of storage. The negative zeta-potential
value is due to the presence of sodium dodecyl sulfate (negatively
charged stabilizing agent) [45].
Fig. 3. FT-IR spectra obtained of the samples from optimization study. Fig. 5a and b shows the TEM images of the sample. Fig. 5c shows the
distribution curve of particle size, in which it is possible to see that the
average diameter of MNXQ_SO3 was approximately 126 ± 31 nm. In
agent was consumed during the polymerization process. However, only this case, the particle size obtained by TEM analysis for MNXQ_SO3 is
the sample S6 was selected to proceed with the experiments. Thus, the smaller than that obtained by DLS. Probably, the difference between
sample S6 was called of MNXQ_SO3 and was subjected to more the average sizes obtained from DLS and TEM analyses occurs because
characterizations. the TEM analysis was performed under vacuum atmosphere [46].
Baier et al. [29] obtained starch nanocapsules with encapsuled dsDNA
in which the average size determined by TEM micrograph was in the
3.2.2. Infrared spectroscopy range of 286 to 381 nm.
FT-IR analysis of MNXQ_SO3 (Fig. 4a) was recorded to investigate the Furthermore, the TEM images of MNXQ_SO3 show the core-shell
shell composition of nanocapsules and to confirm the chemical reaction structure, confirming the nanocapsule morphology. It also could be
between NCO groups of TDI with OH groups from xyloglucan and surfac- seen the presence of dark contrasts around the nanocapsules, indicating
tant molecules. The effectiveness interfacial polymerization was evalu- the successful synthesis of magnetic nanocapsules via inverse
ated by crosslinked of xyloglucan nanocapsules, resulting on the miniemulsion.
presence of urea and urethane groups. Bands at 1734 cm−1 (C_O vibra-
tion) and 1544 cm−1 (N\\H vibration) evidence the formation of ure- 3.2.4. Thermal analysis: TGA/DTG of the samples
thane groups. Furthermore, the band at 1645 cm−1 represents C_O The thermogram curve of the xyloglucan sample (Fig. S2 in Appen-
vibration of urea formed through the reaction between isocyanate with dix A – Supplementary data) shows two steps of weight loss, with
water. The absence at 2275 cm−1 indicates that all crosslinker isocyanate total weight loss of around 84%. The first region ranges from 27 °C to
(NCO) was consumed [9,42]. A strong band at 3400 cm−1 is characteris- 106 °C and represents up to about 9% of the weight loss, which is
tic of an oxygen-bonded O\\H stretching vibration. Other bands are also assigned to the residual and structural water evaporation previously
expected in this spectral region (valence vibrations of N\\H and C\\H at present in the sample. Some publications attribute the occurrence of
3300 and 3000 cm−1, respectively), but probably are overlapped by the this thermal degradation to the hydrophilic nature of the polysaccharide
O\\H vibration band. The literature reports the presence of a band at (presence of hydroxyl functional groups in the chemical structure) [40].
1600 cm−1 which refers to C_C vibration valence from aromatic sys- The second step appears in the temperature range of 239–468 °C and is
tems [29]. Furthermore, it is possible to see a band around 584 cm−1, related to the thermal decomposition of xyloglucan chain [47], which
also present in Fe3O4@PAAS spectrum (Fig. 4b), which refers to Fe\\O vi- corresponds to 75% of the weight loss. Furthermore, the Derivative
bration of the magnetite (Fe3O4) formation, probably indicating that Thermogravimetric analysis (DTG) shows a maximum degradation at
magnetite was incorporated into nanocapsule sample [5]. a temperature of 300 °C, which occurs in the second main step thermal

727
A.C.C. da Silva, R.R. de Almeida, A.C. da Cruz Sousa et al. International Journal of Biological Macromolecules 168 (2021) 722–732

Fig. S2 also illustrates the thermal behavior of Q_SO3. Three main


steps of weight loss are observed in the TGA/DTG curves. The first region
starts at 26 °C and ends at 113 °C, which corresponds to the water loss
(11%). The second weight loss (around 16%) occurs in the temperature
range of 193–411 °C and is attributed to overlapping of several pro-
cesses according to the DTG curve profile. Woznicka et al. [44] report
this degradation process generates H2O, SO2, CO2 and CO. The third
step of decomposition (weight loss up to about 36%) begins at 603 °C
and ends at 794 °C, in which it is related to production of CO and CO2
[44].
The thermogravimetric analysis (Fig. S2) of MNXQ_SO3 sample was
performed to investigate the thermal stability. The TGA profile of
MNXQ_SO3 exhibits a stepwise weight-loss curve and the total weight
loss is around of 78%. The first step appears in the range of 26–97 °C
and probably corresponds to endothermic desorption of physically
absorbed water. The second thermal event occurs in the temperature
range of 138–276 °C, which can be associated to degradation processes
of xyloglucan. The thermal stability of xyloglucan in the sample
MNXQ_SO3 is lower than the polysaccharide alone due to the reduction
of hydrogen bonds along the polymer chain, which occurs by the reac-
tion with the crosslinking agent [49]. This reduction of the thermal sta-
bility can be a relevant result to drug delivery process, which occurs in
the presence of magnetic field because the temperature increase caused
by magnetic hyperthermia should accelerate the drug release process.
Therefore, the maximum degradation for this sample is also lower
than polysaccharide and occurs at 214 °C. The last degradation step hap-
pens between 288 °C and 483 °C and perhaps can be attributed to the
thermal decomposition of the hydrophilic quercetin sulfate. The in-
creased thermal stability of bioactive in the sample, when compared
to bioactive alone, can indicate the successfully loaded Q_SO3 into the
magnetic nanocapsules.

3.2.5. X-ray powder diffraction (XRD)


The Fe3O4@PAAS sample was subject to XRD technique, which is
displayed in Fig. 6. All the peaks of the sample spectrum exhibited
very similar pattern to magnetite (JCPDS card # 19-0629) phase,
which possesses an inverse cubic spinel structure [50]. No other hema-
tite (JCPDS card # 01-086-0550) or maghemite (JCPDS card # 39-1346)
peaks were found in the sample, revealing a pure phase of magnetite
(Fe3O4) [51,52]. This result is in accordance with the data in the litera-
ture [25].
Fig. 7a and b exhibits the TEM micrographs of magnetic nanoparti-
cles coated with PAAS (Fe3O4@PAAS) and shows the morphology, size
and structural defects of the synthesized sample. It is possible to observe
the faceted shape of magnetic nanoparticles and the absence of nano-
particles clusters. Moreover, the sample corresponds to nanosized mag-
netite as described in the literature [25]. Through Fig. 7c is possible to
see the size histogram for Fe3O4@PAAS, which the mean particle size
was determined (approximately 8 nm).
Fig. 4. FT-IR spectra of the samples (a) MNXQ_SO3, (b) Fe3O4@PAAS, (c) xyloglucan and
(d) Q_SO3. 3.2.6. Magnetic properties
Fig. 8 shows the hysteresis loops as a function of the magnetic field at
room temperature (300 K). Through Fig. 8(a) it is possible to evaluate
decomposition. This result is consistent with other papers previously the influence of the coating on the magnetization in the sample
published [26,48]. Fe3O4@PAAS. The sample Fe3O4 with the coating (PAAS) reduces the
TGA/DTG curves of Fe3O4@PAAS (Fig. S2) exhibit three weight loss saturation magnetization from 56 emu/g to 48 emu/g (Fe3O4@PAAS).
steps. The first one (around 4%) is observed in the temperature range This decrease does not change the superparamagnetic behavior and
of 31–103 °C and corresponds to the sample dehydratation. The second the coercivity and remanence values remain practically zero. Different
thermal step (approximately 10%) occurs in the range of 182–415 °C values of saturation magnetization can be attributed to the different
and represents the exothermic decomposition of organic capping amounts of coating and preparation techniques [53,54]. In this paper
agent (PAAS). Finally, the third weight loss (about 16%) happens in we obtained the sample Fe3O4@PAAS with 30% of coat (determined
the temperature range of 458–730 °C and is related to further degrada- by TGA).
tion of the resultant carbonic residues, as well as the partial reduction of The magnetization of MNXQ_SO3 is shown in Fig. 8(b) and a satura-
Fe3O4 surface [25]. Thus, the total weight loss (30%) indicates the per- tion magnetization value of 4.2 emu/g is obtained. The reduction in
centage of coating formed by polyacrylic acid sodium salt in the magnetization of the sample was expected due to the presence of poly-
Fe3O4@PAAS sample. mer and organic molecules, which contribute to the increase in the total

728
A.C.C. da Silva, R.R. de Almeida, A.C. da Cruz Sousa et al. International Journal of Biological Macromolecules 168 (2021) 722–732

Fig. 5. (a) and (b) TEM micrographs of MNXQ_SO3; (c) particle size distribution curve of MNXQ_SO3.

mass of the nanocapsules, reducing the saturation magnetization [3].


This result is similar to that published by Kavousi et al. [55], in which
the authors synthesized a magnetic polymer nanocomposite for the re-
lease of metoprolol with the saturation magnetization value of 3.5 emu/
g [55]. Furthermore, the VSM analysis of MNXQ_SO3 indicates the prom-
ising use of this nanomaterial in future studies for biotechnological and
theranostic applications, such as its use in magnetic hyperthermia as-
says and diagnostic imaging tests as a contrast agent.

3.3. Release study and efficiency encapsulation of hydrophilic quercetin


sulfate

The release behavior of hydrophilic quercetin sulfate (Q_SO3) free


and encapsulated within MNXQ_SO3 sample was performed in contact
with phosphate buffered saline at pH 7.4 and can be seen in Fig. 9.
After of the first 30 min of experiment about 22% of Q_SO3 was released
from nanocapsules, while the percentage concentration of free bioactive
was 34%. The total amount of free Q_SO3 (100%) in the PBS medium was
obtained after 48 h of release assay. However, at the end of the delivery
study, the rate released of Q_SO3 from MNXQ_SO3 sample was 67%. It is
already known in the literature that the delivery of the encapsulated
material into polymeric nanocapsules shows a drug release of more
than 50% [29].
Through Fig. 9, it is possible to verify that MNXQ_SO3 shows a signif-
icant increase of the drug rate release (burst effect) in the first 8 h and
until 24 h, the percentage rate of Q_SO3 release increased slowly. The lit-
erature reports that polymer-based nanocapsules are biphasic systems
and have a release behavior of apparent zero order kinetics. Therefore,
Fig. 6. XRD pattern diffraction of magnetic nanoparticles coated with polyacrylic acid these systems have a fast initial phase of drug released and a second
sodium salt (Fe3O4@PAAS). slower phase: in the fast step the desorption of the bioactive on the

729
A.C.C. da Silva, R.R. de Almeida, A.C. da Cruz Sousa et al. International Journal of Biological Macromolecules 168 (2021) 722–732

Fig. 7. (a) and (b) TEM images of Fe3O4@PAAS; (c) particle size distribution curve of Fe3O4@PAAS.

nanocapsule surface or the degradation of the polymeric nanocapsules Recent studies demonstrate that the presence of magnetic nanopar-
wall can occur, while the slower step is attributed to diffusion of the en- ticles in drug delivery vehicles can be able to ablate cancer cells by mag-
capsulated drug [45]. netic hyperthermia. Additionally, the heat formed by this process can
Table 3 shows the results of the kinetic study of mathematical modify the release profile of bioactives, increasing the release rate.
models, such as zero and first order, Higuchi and Korsmeyer-Peppas, Therefore, the combination of both techniques can increase the effi-
which were employed to evaluate the in vitro quercetin sulfate release ciency of cancer treatment and reduce the secondary effects of non-
from MNXQ_SO3. The results indicate that the sample is in accordance targeted nanocarriers [56]. In this study, the presence of Fe3O4@PAAS
with the mathematical model of zero order and, in other words, this into MNXQ_SO3 can be enlarge the rate release of quercetin sulfate, im-
profile releases the same amount of bioactive per unit of time, being proving the response of this hybrid nanocomposite to targeted drug
one of the best ways of deliver drugs in a prolonged release [30]. delivery.

Fig. 8. Magnetic behavior of (a) Fe3O4, Fe3O4@PAAS and (b) MNXQ_SO3, at 300 K.

730
A.C.C. da Silva, R.R. de Almeida, A.C. da Cruz Sousa et al. International Journal of Biological Macromolecules 168 (2021) 722–732

Araneda Martínez: Investigation. Raimundo Rafael de Almeida: Con-


ceptualization, Supervision, Writing – review & editing. Juliano
Casagrande Denardin: Investigation. Writing – review & editing.
Selene Maia de Morais: Funding acquisition. Nágila M. P. S. Ricardo:
Conceptualization, Writing – review & editing, Supervision, Funding
acquisition.

Acknowledgments

This study was financed in part by the Coordenação de


Aperfeiçoamento de Pessoal de Nível Superior - Brasil (CAPES) - Finance
Code 001, Conselho Nacional de Desenvolvimento Científico e
Tecnológico (CNPq) and Fundação Cearense de Apoio ao
Desenvolvimento Científico e Tecnológico (FUNCAP), Edital No 03/
2019. The authors thank CNPq for their grants (N.M.P.S.R. – Project No
307837/2017-3 and A.C.C.S. Project No 140256/2017-2) and FUNCAP
(R.R.A (Project No 88887.162539/2017-00)). We also want to acknowl-
edge the analyses and data acquisition provided by Embrapa
Fig. 9. Release behaviors of Q_SO3 free and MNXQ_SO3 at pH 7.4 during 48 h. Agroindústria Tropical, X-ray Laboratory of Federal University of Ceará
(LRX-UFC), Fondecyt 1200782, CEDENNA AFB180001 (USACH) and
Central Analítica-UFC/CT-INFRA/MCTI-SISNANO/CAPES.
Table 3
Kinetic study of data obtained by Q_SO3 release profile at pH 7.4.
Data availability statement
Sample Correlation coeficiente (r2)

Zero-order First-order Higuchi Korsmeyer-Peppas The raw/processed data required to reproduce these findings cannot
be shared at this time as the data also forms part of an ongoing study.
Q_SO3 free 0.8995 0.5627 0.9826 0.9789
MNXQ_SO3 0.9986 0.5768 0.9494 0.9551
Declaration of competing interest

The encapsulation efficiency of QSO3 in the sample MNXQ_SO3 was The authors declare no conflict of interest.
determined and showed a value of 56%. Studies published previously
by other authors indicate that nanocapsules consisting of an aqueous References
core and polymer shell have encapsulation efficiency varying in the
[1] B. Kang, P. Okwieka, S. Schöttler, O. Seifert, R.E. Kontermann, K. Pfizenmaier, A.
range of 46% until close to 100%. The difference between the encapsula-
Musyanovych, R. Meyer, M. Diken, U. Sahin, V. Mailänder, F.R. Wurm, K.
tion efficiency values is assigned to the crosslinking degree [9,57]. Landfester, Tailoring the stealth properties of biocompatible polysaccharide
nanocontainers, Biomaterials 49 (2015) 125–134, https://doi.org/10.1016/j.
biomaterials.2015.01.042.
4. Conclusion
[2] L. De Matteis, D. Jary, A. Lucía, S. García-Embid, I. Serrano-Sevilla, D. Pérez, J.A. Ainsa,
F.P. Navarro, J. M. de la Fuente, New active formulations against M. tuberculosis:
In this work, magnetic nanoparticles (Fe3O4@PAAS) were obtained bedaquiline encapsulation in lipid nanoparticles and chitosan nanocapsules, Chem.
by hydrothermal synthesis. The analyses of magnetic properties and Eng. J. 340 (2018) 181–191, https://doi.org/10.1016/j.cej.2017.12.110.
[3] R.R. De Almeida, J. Gallo, A.C.C. Da Silva, A.K.O. Da Silva, O.D.L. Pessoa, T.G. Araújo,
X-ray powder diffraction reveal, respectively, a superparamagnetic be-
L.K.A.M. Leal, P.B.A. Fechine, M. Bañobre-López, N.M.P.S. Ricardo, Preliminary evalu-
havior and a pure-phase of magnetite for Fe3O4@PAAS. The obtaining ation of novel triglyceride-based nanocomposites for biomedical applications, J.
of Q_SO3 allowed its incorporation in hydrophilic-core nanocapsules, Braz. Chem. Soc. 28 (2017) 1547–1556, https://doi.org/10.21577/0103-5053.
increasing the application potential of this bioactive. Magnetic 20170007.
[4] V. Balan, G. Dodi, N. Tudorachi, O. Ponta, V. Simon, M. Butnaru, L. Verestiuc,
nanocapsules of xyloglucan loaded with hydrophilic quercetin sulfate Doxorubicin-loaded magnetic nanocapsules based on N-palmitoyl chitosan and
(MNXQ_SO3) were successfully synthesized by inverse miniemulsion magnetite: synthesis and characterization, Chem. Eng. J. 279 (2015) 188–197,
polymerization. The nanocapsule shell formation and magnetite incor- https://doi.org/10.1016/j.cej.2015.04.152.
poration in the sample were corroborated by FT-IR and TEM [5] J. Zufía-Rivas, P. Morales, S. Veintemillas-Verdaguer, Effect of the sodium
polyacrylate on the magnetite nanoparticles produced by green chemistry routes:
micrography. After the incorporation of Fe3O4@PAAS in the xyloglucan applicability in forward osmosis, Nanomaterials 8 (2018) 470, https://doi.org/10.
nanocapsules loaded with Q_SO3, the superparamagnetic behavior 3390/nano8070470.
was consistent, being an attractive nanomaterial for drug delivery to [6] R.A. Revia, M. Zhang, Magnetite nanoparticles for cancer diagnosis, treatment, and
treatment monitoring: recent advances, Mater. Today 19 (2016) 157–168, https://
specific tissues without damaging healthy cells. The sample achieved a
doi.org/10.1016/j.mattod.2015.08.022.
promising encapsulation efficiency and was able to modify the release [7] F. Morales, G. Márquez, V. Sagredo, T.E. Torres, J.C. Denardin, Structural and mag-
profile of hydrophilic quercetin sulfate (Q_SO3), when compared with netic properties of silica-coated magnetite nanoaggregates, Phys. B Condens. Matter
the free hydrophilic bioactive. This work allows for the development 572 (2019) 214–219, https://doi.org/10.1016/j.physb.2019.08.007.
[8] R. Grillo, J. Gallo, D.G. Stroppa, E. Carbó-Argibay, R. Lima, L.F. Fraceto, M. Bañobre-
of the novel multifunctional tool that has promising potential for vari-
López, Sub-micrometer magnetic nanocomposites: insights into the effect of mag-
ous biomedical applications, such as drug delivery, magnetic hyperther- netic nanoparticles interactions on the optimization of SAR and MRI performance,
mia and agent contrast in MRI. ACS Appl. Mater. Interfaces 8 (2016) 25777–25787, https://doi.org/10.1021/
acsami.6b08663.
Supplementary data to this article can be found online at https://doi.
[9] F.R. Steinmacher, G. Baier, A. Musyanovych, K. Landfester, P.H.H. Araújo, C. Sayer,
org/10.1016/j.ijbiomac.2020.11.128. Design of cross-linked starch nanocapsules for enzyme-triggered release of hydro-
philic compounds, Processes 5 (2017) 25, https://doi.org/10.3390/pr5020025.
CRediT authorship contribution statement [10] A.D. Kulkarni, A.A. Joshi, C.L. Patil, P.D. Amale, H.M. Patel, S.J. Surana, V.S. Belgamwar,
K.S. Chaudhari, C.V. Pardeshi, Xyloglucan: a functional biomacromolecule for drug
delivery applications, Int. J. Biol. Macromol. 104 (2017) 799–812, https://doi.org/
Aiêrta Cristina Carrá da Silva: Conceptualization, Investigation, For- 10.1016/j.ijbiomac.2017.06.088.
mal analysis, Writing – Original Draft. Writing – review & editing. [11] A.T. Paulino, A.G.B. Pereira, A.R. Fajardo, K. Erickson, M.J. Kipper, E.C. Muniz, L.A.
Alexandre Carreira da Cruz Sousa: Investigation. Fabián Nicolás Belfiore, E.B. Tambourgi, Natural polymer-based magnetic hydrogels: potential

731
A.C.C. da Silva, R.R. de Almeida, A.C. da Cruz Sousa et al. International Journal of Biological Macromolecules 168 (2021) 722–732

vectors for remote-controlled drug release, Carbohydr. Polym. 90 (2012) [34] S. Periasamy, C.H. Lin, B. Nagarajan, N.V. Sankaranarayanan, U.R. Desai, M.Y. Liu,
1216–1225, https://doi.org/10.1016/j.carbpol.2012.06.051. Tamarind xyloglucan attenuates dextran sodium sulfate induced ulcerative colitis:
[12] X. Hu, Y. Wang, L. Zhang, M. Xu, J. Zhang, W. Dong, Design of a pH-sensitive mag- role of antioxidation, J. Funct. Foods 42 (2018) 327–338, https://doi.org/10.1016/j.
netic composite hydrogel based on salecan graft copolymer and Fe3O4@SiO2 nano- jff.2018.01.014.
particles as drug carrier, Int. J. Biol. Macromol. 107 (2018) 1811–1820, https://doi. [35] M. Bhalekar, S. Sonawane, S. Shimpi, Synthesis and characterization of a cysteine
org/10.1016/j.ijbiomac.2017.10.043. xyloglucan conjugate as mucoadhesive polymer, Braz. J. Pharm. Sci. 49 (2013)
[13] X. Hu, Y. Wang, L. Zhang, M. Xu, Formation of self-assembled polyelectrolyte com- 285–292, https://doi.org/10.1590/S1984-82502013000200010.
plex hydrogel derived from salecan and chitosan for sustained release of vitamin [36] I.R.S. Arruda, P.B.S. Albuquerque, G.R.C. Santos, A.G. Silva, P.A.S. Mourão, M.T.S.
C, Carbohydr. Polym. 234 (2020), 115920, https://doi.org/10.1016/j.carbpol.2020. Correia, A.A. Vicente, M.G. Carneiro-da-Cunha, Structure and rheological properties
115920. of a xyloglucan extracted from Hymenaea courbaril var. courbaril seeds, Int. J. Biol.
[14] X. Hu, Y. Wang, L. Zhang, M. Xu, Construction of self-assembled polyelectrolyte com- Macromol. 73 (2015) 31–38, https://doi.org/10.1016/j.ijbiomac.2014.11.001.
plex hydrogel based on oppositely charged polysaccharides for sustained delivery of [37] A. Mishra, A.V. Malhotra, Tamarind xyloglucan: a polysaccharide with versatile ap-
green tea polyphenols, Food Chem. 306 (2020), 125632, https://doi.org/10.1016/j. plication potential, J. Mater. Chem. 19 (2009) 8528–8536, https://doi.org/10.1039/
foodchem.2019.125632. b911150f.
[15] T.P. Nguyen, C.W. Lee, S. Hassen, H.C. Le, Hybrid nanocomposites for optical applica- [38] M. Masuelli, D. Renard, Advances in Physicochemical Properties of Biopolymers
tions, Solid State Sci. 11 (2009) 1810–1814, https://doi.org/10.1016/j. (Part 2), First edit. Bentham Sciense Publishers, Argentina, 2017 198–199, https://
solidstatesciences.2009.05.011. doi.org/10.2174/97816810854491170101 n.d.
[16] E.A. Ferreira, V. Mendonça, H.A. de Souza, J.D. Ramos, Adubação Fosfatada E [39] D.K. Watt, D.J. Brasch, D.S. Larsen, L.D. Melton, Isolation, characterization, and NMR
Potássica Na Formação De Mudas De Tamarindeiro phosphate and potassic fertiliza- study of xyloglucan from enzymatically depectinized and non-depectinized apple
tion on seedling production of tamarind fruit, Sci. Agradria. 9 (2008) 475–480. pomace, Carbohydr. Polym. 39 (1999) 165–180, https://doi.org/10.1016/S0144-
[17] C.V. Pardeshi, A.D. Kulkarni, V.S. Belgamwar, S.J. Surana, Xyloglucan for Drug Deliv- 8617(99)00002-8.
ery Applications, Elsevier Ltd, 2018https://doi.org/10.1016/B978-0-08-102194-1. [40] F.R.S. Mendes, M.S.R. Bastos, L.G. Mendes, A.R.A. Silva, F.D. Sousa, A.C.O. Monteiro-
00007-4. Moreira, H.N. Cheng, A. Biswas, R.A. Moreira, Preparation and evaluation of hemicel-
[18] Y. Li, J. Yao, C. Han, J. Yang, M.T. Chaudhry, S. Wang, H. Liu, Y. Yin, Quercetin, inflam- lulose films and their blends, Food Hydrocoll. 70 (2017) 181–190, https://doi.org/
mation and immunity, Nutrients. 8 (2016) 167, https://doi.org/10.3390/nu8030167. 10.1016/j.foodhyd.2017.03.037.
[19] A. Massi, O. Bortolini, D. Ragno, T. Bernardi, G. Sacchetti, M. Tacchini, C. De Risi, Re- [41] K. Gurpreet, S.K. Singh, Review of nanoemulsion formulation and characterization
search progress in the modification of quercetin leading to anticancer agents, Mol- techniques, Indian J. Pharm. Sci. 80 (2018) 781–789.
ecules 22 (2017) 1270, https://doi.org/10.3390/molecules22081270. [42] L. Bacher, M. Maskos, A. Musyanovych, Gelatin-based capsules through interfacial
polymerization: batch and continuous flow synthesis, Chem. Eng. Technol. (2019)
[20] K.P.S. N.M.P.S. Lopes, I.M. Cavalcante, R.F. Silva, D.H.A. Brito, L.M.U. Fechine, D.R.
1–9, https://doi.org/10.1002/ceat.201900119.
Moreira, I.G.P. Vieira, F.V.C.S. Azul, L.K.A.M. Leal, M.E.N.P. Ribeiro, Ricardo, Binary mi-
[43] M. Kačuráková, A.C. Smith, M.J. Gidley, R.H. Wilson, Molecular interactions in bacte-
celles (E45 S8/F127) for quercetin and griseofulvin solubilisation, Quim Nova 43
rial cellulose composites studied by 1D FT-IR and dynamic 2D FT-IR spectroscopy,
(2020) 1011–1016 , doi:Vol. 43, No 10.21577/0100-4042.20170570.
Carbohydr. Res. 337 (2002) 1145–1153, https://doi.org/10.1016/S0008-6215(02)
[21] M. Poór, G. Boda, S. Kunsági-Máté, P.W. Needs, P.A. Kroon, B. Lemli, Fluorescence
00102-7.
spectroscopic evaluation of the interactions of quercetin, isorhamnetin, and
[44] E. Woźnicka, E. Pieniązek, L. Zapała, Byczyński, I. Trojnar, M. Kopacz, New sulfonic
quercetin-3′-sulfate with different albumins, J. Lumin. 194 (2018) 156–163,
derivatives of quercetin as complexing reagents: synthesis, spectral, and thermal
https://doi.org/10.1016/j.jlumin.2017.10.024.
characterization, J. Therm. Anal. Calorim. 120 (2015) 351–361, https://doi.org/10.
[22] F.D. de Sousa, M.L. Holanda-Araújo, J.R.R. de Souza, R. de S. Miranda, R.R. Almeida, E.
1007/s10973-014-3677-7.
Gomes-Filho, N.M. Pontes-Ricardo, A.C.O. Monteiro-Moreira, R. de A. Moreira, Phys-
[45] C.E. Mora-Huertas, H. Fessi, A. Elaissari, Polymer-based nanocapsules for drug deliv-
icochemical properties of edible seed hemicelluloses, OALib 04 (2017) 1–14, https://
ery, Int. J. Pharm. 385 (2010) 113–142, https://doi.org/10.1016/j.ijpharm.2009.10.
doi.org/10.4236/oalib.1103683.
018.
[23] A.M. de Godoi, L.C. Faccin-Galhardi, D.Z. Rechenchoski, T.B.M.G. Arruda, A.P. Cunha,
[46] S.B.F. dos Santos, S.A. Pereira, F.A.M. Rodrigues, A.C.C. da Silva, R.R. de Almeida, A.C.C.
R.R. de Almeida, F.E.A. Rodrigues, N.M.P.S. Ricardo, C. Nozawa, R.E.C. Linhares, Struc-
Sousa, L.M.U.D. Fechine, J.C. Denardin, F. Araneda, L.G.A.V. Sá, C.R. da Silva, H.V.
tural characterization and antiviral activity of pectin isolated from Inga spp, Int. J.
Nobre Júnior, N.M.P.S. Ricardo, Antibacterial activity of fluoxetine-loaded starch
Biol. Macromol. 139 (2019) 925–931, https://doi.org/10.1016/j.ijbiomac.2019.07.
nanocapsules, Int. J. Biol. Macromol. 164 (2020) 2813–2817, https://doi.org/10.
212.
1016/j.ijbiomac.2020.08.184.
[24] V.G. Nair, S. Bernstein, Rutin poly (H-) sulfate salts and related compounds, US Pat.
[47] N.L. Santos, R.C. Braga, M.S.R. Bastos, P.L.R. Cunha, F.R.S. Mendes, A.M.M.T. Galvão,
4,414,207. 12 (1983) 1–12. https://patents.google.com/patent/US4414207A/en.
G.S. Bezerra, A.A.C. Passos, Preparation and characterization of Xyloglucan films ex-
[25] Y.V. Kolen’Ko, M. Bañobre-López, C. Rodríguez-Abreu, E. Carbó-Argibay, A. Sailsman, tracted from Tamarindus indica seeds for packaging cut-up ‘Sunrise Solo’ papaya,
Y. Piñeiro-Redondo, M.F. Cerqueira, D.Y. Petrovykh, K. Kovnir, O.I. Lebedev, J. Rivas, Int. J. Biol. Macromol. 132 (2019) 1163–1175, https://doi.org/10.1016/j.ijbiomac.
Large-scale synthesis of colloidal Fe3O4 nanoparticles exhibiting high heating effi- 2019.04.044.
ciency in magnetic hyperthermia, J. Phys. Chem. C 118 (2014) 8691–8701, https:// [48] E.M. Bergstȑm, L. Salmén, J. Kochumalayil, L. Berglund, Plasticized xyloglucan for im-
doi.org/10.1021/jp500816u. proved toughness-thermal and mechanical behaviour, Carbohydr. Polym. 87 (2012)
[26] M.D.P. Farias, P.B.S. Albuquerque, P.A.G. Soares, D.M.A.T. de Sá, A.A. Vicente, M.G. 2532–2537, https://doi.org/10.1016/j.carbpol.2011.11.024.
Carneiro-da-Cunha, Xyloglucan from Hymenaea courbaril var. courbaril seeds as en- [49] L. Poon, L.D. Wilson, J.V. Headley, Chitosan-glutaraldehyde copolymers and their
capsulating agent of L-ascorbic acid, Int. J. Biol. Macromol. 107 (2018) 1559–1566, sorption properties, Carbohydr. Polym. 109 (2014) 92–101, https://doi.org/10.
https://doi.org/10.1016/j.ijbiomac.2017.10.016. 1016/j.carbpol.2014.02.086.
[27] R.B. Souza, A.F. Frota, J. Silva, C. Alves, A.Z. Neugebauer, S. Pinteus, J.A.G. Rodrigues, [50] J. Wang, M. Fan, X. Bian, M. Yu, T. Wang, S. Liu, Y. Yang, Y. Tian, R. Guan, Enhanced
E.M.S. Cordeiro, R.R. de Almeida, R. Pedrosa, N.M.B. Benevides, In vitro activities of magnetic heating efficiency and thermal conductivity of magnetic nanofluids with
kappa-carrageenan isolated from red marine alga Hypnea musciformis: antimicro- FeZrB amorphous nanoparticles, J. Magn. Magn. Mater. 465 (2018) 480–488,
bial, anticancer and neuroprotective potential, Int. J. Biol. Macromol. 112 (2018) https://doi.org/10.1016/j.jmmm.2018.06.043.
1248–1256, https://doi.org/10.1016/j.ijbiomac.2018.02.029. [51] P.P. Soares, G.S. Barcellos, C.L. Petzhold, V. Lavayen, Iron oxide nanoparticles modi-
[28] S. Cruz dos Santos, N. Osti Silva, J.B. dos Santos Espinelli, M.A. Germani Marinho, Z. fied with oleic acid: vibrational and phase determination, J. Phys. Chem. Solids 99
Vieira Borges, N. Bruzamarello Caon Branco, F.L. Faita, B. Meira Soares, A.P. Horn, (2016) 111–118, https://doi.org/10.1016/j.jpcs.2016.08.006.
A.L. Parize, V. Rodrigues de Lima, Molecular interactions and physico-chemical char- [52] I.V. Chernyshova, S. Ponnurangam, P. Somasundaran, Linking interfacial chemistry
acterization of quercetin-loaded magnetoliposomes, Chem. Phys. Lipids 218 (2019) of CO2 to surface structures of hydrated metal oxide nanoparticles: hematite,
22–33, https://doi.org/10.1016/j.chemphyslip.2018.11.010. Phys. Chem. Chem. Phys. 15 (2013) 6953–6964, https://doi.org/10.1039/
[29] G. Baier, A. Musyanovych, M. Dass, S. Theisinger, K. Landfester, Cross-linked starch c3cp44264k.
capsules containing dsDNA prepared in inverse miniemulsion as “nanoreactors” [53] I. Ban, S. Markuš, S. Gyergyek, M. Drofenik, J. Korenak, C. Helix-nielsen, I. Petrini,
for polymerase chain reaction, Biomacromolecules 11 (2010) 960–968, https://doi. Synthesis of poly-sodium-acrylate (PSA)-coated magnetic nanoparticles for use in
org/10.1021/bm901414k. forward osmosis draw solutions, 9 (2019) 1238.
[30] P.J.C. Da Costa, Avaliação in vitro da lioequivalência de formulações farmacêuticas, [54] J. Sato, M. Kobayashi, H. Kato, T. Miyazaki, M. Kakihana, Hydrothermal synthesis of
Rev. Bras. Ciencias Farm. J. Pharm. Sci. 38 (2002) 141–153, https://doi.org/10. magnetite particles with uncommon crystal facets, J. Asian Ceram. Soc. 2 (2014)
1590/s1516-93322002000200003. 258–262, https://doi.org/10.1016/j.jascer.2014.05.008.
[31] R.R. de Almeida, H.S. Magalhães, J.R.R. de Souza, M.T.S. Trevisan, Í.G.P. Vieira, J.P.A. [55] F. Kavousi, M. Goodarzi, D. Ghanbari, K. Hedayati, Synthesis and characterization of a
Feitosa, T.G. Araújo, N.M.P.S. Ricardo, Exploring the potential of Dimorphandra magnetic polymer nanocomposite for the release of metoprolol and aspirin, J. Mol.
gardneriana galactomannans as drug delivery systems, Ind. Crop. Prod. 69 (2015) Struct. 1183 (2019) 324–330, https://doi.org/10.1016/j.molstruc.2019.02.003.
284–289, https://doi.org/10.1016/j.indcrop.2015.02.041. [56] C.L. O.D.L. Moura, J. Gallo, L. Garcia-Hevia, M. Pessoa, N.M.P.S. Ricardo, Bañobre-
[32] K.A. Nobre, C.E.A. Soares, Í.G.P. Vieira, R.R. De Almeida, R.D.A. Moreira, T.G. De López, Magnetic hybrid wax nanocomposites as externally controlled theranostic
Araújo, M.E.N.P. Ribeiro, N.M.P.S. Ricardo, Adenanthera pavonina galactomannan vehicles: high MRI enhancement and synergistic magnetically-assisted thermo/
for controlled delivery of rutin-a preliminary study, Quim Nova 41 (2018) chemo therapy, Chem. Eur. J. 26 (2019) 4531–4538, https://doi.org/10.1002/chem.
607–612, https://doi.org/10.21577/0100-4042.20170226. 201904709.
[33] D.C. Rodrigues, A.P. Cunha, L.M.A. Silva, T.H.S. Rodrigues, M.I. Gallão, H.M.C. Azeredo, [57] I. Schlegel, P. Renz, J. Simon, I. Lieberwirth, S. Pektor, N. Bausbacher, M. Miederer, V.
Emulsion films from tamarind kernel xyloglucan and sesame seed oil by different Mailänder, R. Muñoz-Espí, D. Crespy, K. Landfester, Highly loaded semipermeable
emulsification techniques, Food Hydrocoll. 77 (2018) 270–276, https://doi.org/10. nanocapsules for magnetic resonance imaging, Macromol. Biosci. 18 (2018) 1–12,
1016/j.foodhyd.2017.10.003. https://doi.org/10.1002/mabi.201700387.

732

You might also like