You are on page 1of 8

International Journal of Biological Macromolecules 216 (2022) 520–527

Contents lists available at ScienceDirect

International Journal of Biological Macromolecules


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

Sulfated xyloglucan-based magnetic nanocomposite for preliminary


evaluation of theranostic potential
Aiêrta Cristina Carrá da Silva a, Raimundo Rafael de Almeida b, Cristine Soares Vidal a,
João Francisco Câmara Neto a, Alexandre Carreira da Cruz Sousa a,
Fabián Nicolás Araneda Martínez c, Daniel Pascoalino Pinheiro d, Sarah Leyenne Alves Sales d,
Cláudia Pessoa d, Juliano Casagrande Denardin c, 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, Zip Code 62400-000 Camocim, CE, Brazil
c
University of Santiago of Chile and Cedenna, USACH-CEDENNA, Department of Physics, Zip Code 9170124 Santiago, Chile
d
Laboratory of Experimental Oncology, Center for Research and Drug Development, Federal University of Ceará, Zip Code 60430-275 Fortaleza, CE, Brazil
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

Keywords: Chemically sulfated xyloglucan (S_XLG) was applied to encapsulate hydrophilic quercetin sulfate (Q_SO3) into
Sulfated xyloglucan superparamagnetic nanocomposites (MNSXQ_SO3) synthesized via interfacial polyaddition reaction. This nano­
Multifunctional magnetic nanocomposite platform was evaluated concerning in vitro antitumor activity, drug release profile, and magnetic behavior under
Targeted drug delivery
a magnetic field. MNSXQ_SO3 exhibited antitumor activity against four human tumor cell lines: leukemia HL-60
(IC50 = 1.58 ± 0.54 μg mL− 1), glioblastoma SNB-19 (IC50 = 8.77 ± 0.36 μg mL− 1), colorectal carcinoma HCT-
116 (IC50 = 8.69 ± 0.21 μg mL− 1) and prostate PC3 (IC50 = 17.42 ± 3.57 μg mL− 1). Additionally, MNSXQ_SO3
was able to delay the release of Q_SO3 (76 %) compared to the free Q_SO3 (100 %), after 48 h of drug delivery
assay, revealing a release profile compatible with the Korsmeyer-Peppas kinetic model. These results represent a
relevant progress in the production of multifunctional hybrid nanocarriers, providing a versatile nanocomposite
with potential for targeted drug delivery and theranostic applications.

1. Introduction carbohydrate backbone structure/composition of sulfated poly­


saccharide [3,4]. However, some advantages in pharmacokinetic prop­
The development of nanocomposites based on natural polymers erties have been reported to chemically sulfated polysaccharides, when
(polysaccharides) has been highlighted in different medical areas compared to natural sulfated polysaccharides, such as the differences of
because they are widely available, biocompatible, biodegradable, in vivo degradability, allowing their application in researches based on
nontoxic and flexible to chemical modification [1]. The sulfation reac­ biomaterials and drug delivery vehicles [5].
tion of polysaccharides provides steric hindrance and electrostatic Xyloglucan (XLG) is a high Mw neutral hemicellulose obtained from
repulsion between the biopolymer chains, improving the water solubi­ the Tamarindus indica Linn. seeds [6] composed by a main chain linked
lity in relation to their naturally occurring counterpart. Moreover, the to β-1,4-D-glucopyranosyl residues, partially replaced by α-1,6-D-xylo­
sulfated group causes flexion and extension of the polysaccharide pyranosyl units, which may or may not have β-1,2-galactopyranosyl
chains, resulting in the alteration of polysaccharides biological proper­ monomers as branch [7]. The large amount of hydroxyl groups in these
ties [2]. monomers units can become sites of sulfation reaction. Previously
These biological activities depend on the molecular weight (Mw), the published works show that chemically sulfated xyloglucan have anti­
position of sulfate group, the degree of sulfation (DS) and the tumor activity in vitro against human liver cancer cell line (HepG2),

* Corresponding author.
E-mail address: naricard@ufc.br (N.M.P.S. Ricardo).

https://doi.org/10.1016/j.ijbiomac.2022.06.197
Received 22 February 2022; Received in revised form 28 June 2022; Accepted 29 June 2022
Available online 6 July 2022
0141-8130/© 2022 Elsevier B.V. All rights reserved.
A.C.C. da Silva et al. International Journal of Biological Macromolecules 216 (2022) 520–527

while the unmodified polysaccharide has no anticancer effect [8]. can change these characteristics and improve the therapeutic potential
Recently, superparamagnetic iron oxide nanoparticles (SPIONs) of this biomolecule in clinical applications by the increase of its water
have been demonstrated a great potential for diagnosis and therapy in a solubility.
number of cancer diseases, due to their various advantages such as For these reasons, we performed in this work the synthesis and
biocompatibility, imaging contrast ability, improved tissue penetration, characterization of nanocomposites based on magnetic sulfated xylo­
and superparamagnetic properties [9]. Additionally, SPIONs can be used glucan (S_XLG) loaded within sulfate quercetin (Q_SO3), obtaining a
on the heating of malignant cells, which are more susceptible to heat multifunctional nanocontainer with potential for biomedical applica­
than normal cells, when SPIONs are under the action of an alternating tions, mainly the treatment of cancer by targeting the delivery of hy­
magnetic field. This technique (magnetic hyperthermia) has become an drophilic drugs.
alternative therapy to treat cancer and to reduce side effects caused by
traditional radiotherapy and chemotherapy [10]. 2. Experimental details
According to Ahmadian et al. [11], several nano-based strategies
have been testified and approved for treatment and diagnosis of 2.1. Synthesis of MNSXQ_SO3
different types of cancer. For instance, Resovist®, Cliavist®, Feridex®,
Endorem™, Feraheme®, Rienso®, Lumirem®, GastroMARK® and Magnetic nanocapsules of sulfated xyloglucan loaded with hydro­
Abdoscam® are SPIONs clinically approved, which have reached the philic quercetin sulfate (MNSXQ_SO3) were synthesized through the
market for diagnosis by magnetic resonance imaging or therapeutic water-in-oil miniemulsion technique (Fig. 1), followed by interfacial
application of iron supplementation in anemia [12]. However, SPIONs polymerization reaction using tolylene-2,4-diisocyanate (TDI) as cross­
are technological drivers of innovation and have been utilized in image- linker agent [21].
guided drug delivery, theranostic tissue engineering, targeted delivery Firstly, the aqueous phase (AP), composed by 4 mL of S_XLG 1 % (w/
and controlled release of drugs to specific sites [12,13]. w), 20 mg of Q_SO3 and 12 mg (330 μL) of SPIONs coated with poly­
Quercetin (3,3′ , 4′ , 5, 7-pentahydroxiflavone) is a flavonoid widely acrylic acid sodium salt (Fe3O4@PAAS), was dripped into the organic
found in vegetables, fruits, teas and wines [14], which can be conjugated phase (OP), which was formed by 14.5 g of cyclohexane and 160 mg of
with SPIONs [15] or entrapped in magnetic nanocomposites [16] in polyglycerol polyricinoleate (surfactant). This pre-emulsion was soni­
order to obtain targeted drug delivery systems. This compound has cated in an ultrasound (Branson Sonifier W-450-Digital and a ½” tip) at
several pharmacologic effects, such as antitumor and antioxidant 70 % amplitude during 3 min (20 s pulse; 10 s pause), while cooling the
properties, that can help in the treatment of cancer and other diseases mixture with an ice bath to obtain a nanoemulsion. Finally, 42.5 mg of
[17]. However, quercetin has poor water solubility [18], low bioavail­ TDI was dripped slowly into the obtained nanoemulsion for the inter­
ability [19], rapid elimination from body, fast metabolism and can be facial polymerization reaction to occur. After 24 h of reaction,
enzymatically degraded [20]. The sulfation reaction of this flavonoid MNSXQ_SO3 was redispersed in SDS solution 0.1 %.

Fig. 1. Schematic representation for the obtention of MNSXQ_SO3.

521
A.C.C. da Silva et al. International Journal of Biological Macromolecules 216 (2022) 520–527

All materials were described in Supplementary information, as well experimental procedure were described in Supplementary information.
as the experimental procedures of obtention and characterization of
S_XLG (% S = 4.19; Mw = 156.850 kDa) and Q_SO3 (% S = 13.41). 2.5. In vitro antioxidant activity of Q and Q_SO3 by DPPH radical
scavenging experiment
2.2. Characterization of MNSXQ_SO3
The antioxidant potential of Q and Q_SO3 was determined by the
Fourier Transform Infrared spectroscopy was performed using a ability to scavenge the 2,2-Diphenyl-1-picrylhydrazyl (DPPH•) free
Shimadzu IRTracer-100 Spectrophotometer (Model 8300) in the region radical, using previously reported method [27]. The experimental pro­
between 400 and 4000 cm− 1 [22]. The dried sample was prepared in the cedure was described in Supplementary information.
form of KBr tablets and analyzed.
The stability study of MNSXQ_SO3 was determined by Dynamic Light 3. Results and discussion
Scattering (DLS) measurements, using a Zetasizer Nano Series Malvern
W-ZS90. After preparation, MNSXQ_SO3 was stored at 25 ◦ C for 2 3.1. Characterization of MNSXQ_SO3
months. The sample was diluted 50 times with ultrapure water and
analyzed in an optical grade polystyrene cuvette at 25 ◦ C [16]. The Fourier Transform Infrared (FT-IR) spectroscopy analysis was carried
polydispersity index (PdI), hydrodynamic diameter average (DH) and out to confirm the formation of the nanocapsule polymeric shell. FT-IR
zeta potential (Zp) were determined immediately after MNSXQ_SO3 spectrum of MNSXQ_SO3 was depicted in Fig. 2a and displayed char­
preparation (Day 1) and at predetermined time intervals (7, 14, 21, 28 acteristic bands of urea and urethane groups formed by the successfully
and 60 days). interfacial polyaddition reaction between OH groups of polysaccharide
The morphology of MNSXQ_SO3 was evaluated by scanning electron and NCO groups of crosslinker agent (TDI). Absorptions bands at 1734
microscopy (SEM-FEG model Quanta 450, FEI Company) and trans­ cm− 1 and 1643 cm− 1 are signed to the carbonyl stretching vibration of
mission electron microscopy (TEM-Hitachi®HT7700), according to the urethane and urea groups, respectively. The crosslinked polymer shell is
methodology previously reported [23]. The distribution curve of parti­ also evidenced by the presence of N–H amide bending bands at 1550
cle size was generated by manual measurement of 200 particles, using cm− 1, which are related to urethane group. The flat signal at 2275 cm− 1
the software ImageJ [24]. indicates the total incorporation of NCO groups from TDI into nano­
The X-ray powder diffraction (XRD) of the synthesized sample was capsule shell by their both sides reaction with sulfated xyloglucan
collected using a PANalytical (X'pert Pro MPD) X-ray powder diffrac­ [23,28,29]. The broad band at 3304 cm− 1 is reported to the stretching
tometer operating in Bragg-Brentano reflection geometry with Co-Kα vibration of O–H groups of glucan backbone, while the bands at 2926
radiation (40 kV, 40 mA) over the range 20◦ < 2θ < 100◦ and scan speed cm− 1 and 2850 cm− 1 are attributed to C–H stretching of alkanes [30].
of 0.013◦ /s [25]. The bands at 1600 cm− 1 and 1230 cm− 1 are characteristic of the C– –C
The magnetic properties of MNSXQ_SO3 were investigated using a stretching vibrations from aromatic systems and C– – O stretching [31],
homemade Vibrating Sample Magnetometer (VSM), operating at 300 K respectively, suggesting the incorporation of sulfated quercetin into
with a magnetic field range from − 10 to 10 kOe [25]. The magnetometer nanocapsules. In addition, the band at 1420 cm− 1 corresponds to C–O
was calibrated using Yttrium Iron Garnet Sphere as standard reference stretching vibration [32], while 1138 cm− 1 and 1053 cm− 1 indicate the
material (NIST-National Institute of Standards and Technology). For the stretching vibration of pyranose [33]. The band at 813 cm− 1 is charac­
sample measurement, the magnetic moment obtained for each applied teristic of the symmetric stretching vibration of C-O-S [33], while the
field was normalized by the sample mass. band at 621 cm− 1 is related to Fe–O crystal lattice vibrations,
evidencing the presence of Fe3O4@PAAS into nanoproduct [34].
2.3. Release study of Q_SO3 MNSXQ_SO3 sample was analyzed by DLS technique to determine the
polydispersity index (PdI), hydrodynamic diameter average (DH) and
The in vitro release profile studies were carried out for quercetin zeta potential (Zp) values of nanocapsules at predetermined time in­
sulfate in its free (Q_SO3) and encapsulated forms (MNSXQ_SO3). The tervals (1, 7, 14, 21, 28, 60 days), as shown in Fig. 2b. The PdI (0.293 ±
experimental procedure was performed according to the previously 0.052), DH (274.7 ± 2.2 nm) and Zp (− 28.7 ± 0.9 mV) values obtained
published study [25]. Briefly, a dual-compartment system separated by immediately after the preparation of MNSXQ_SO3 classify this sample as
a dialysis tubing cellulose membrane (molecular weight cut off = 12 a polydispersed particle (0.1 < PdI < 0.4) and moderately stable
kDa) with donor (1.5 mL of MNSXQ_SO3 or Q_SO3) and acceptor com­ dispersed system (− 20 < Zp < − 30 mV), that can be employed in
partments (50 mL, phosphate buffered saline-PBS, pH 7.4) was main­ intravenous administration because are unable to cause embolism (DH
tained under sink conditions with constant magnetic stirring. Aliquots < 300 nm) [31,35,36].
(1 mL) of the samples were collected from the acceptor compartment at Although MNSXQ_SO3 is classified as a moderately stable dispersed
predetermined time intervals and analyzed by UV–Vis Spectrophotom­ system, the DLS analyses performed after the sample preparation reveal
eter (ThermoScientific®, model: Genesis 10S) at 265 nm. Q_SO3 was no drastic changes in the PdI, DH and Zp values (Fig. 2b) during the
quantified using a regression equation generated from the standard storage period at room temperature. Additionally, MNSXQ_SO3
calibration curve obtained from dissolving Q_SO3 in PBS solution at the remained as a transparent liquid without irreversible precipitation or
concentrations of 10 to 200 mg mL− 1. To evaluate the kinetic mecha­ coagulation during the stability study.
nism of Q_SO3, the data obtained from this release assay were fitted to The surface morphology and particle size of MNSXQ_SO3 were
the main mathematical models, as zero and first-order, Higuchi and observed by scanning electron microscopy (SEM) and transmission
Korsmeyer-Peppas. The amount of Q_SO3 encapsulated (%EE) in electron microscopy (TEM), respectively. Fig. 3 displays the SEM mi­
MNSXQ_SO3 sample was also determined by UV-VIS spectrometry using crographs of MNSXQ_SO3, where the sample appears as deflated balls,
an experimental procedure described in literature [25]. due to the drying and vacuum effects of SEM measurements, confirming
the core-shell morphology of nanocapsule [23].
2.4. In vitro antitumor activity of the samples TEM images of MNSXQ_SO3 sample are depicted at Fig. 4, showing
black dots over the gray dots, that are identified as the presence of
The colorimetric MTT assay [26] was employed to determine the in SPIONs (Fe3O4@PAAS) in the nanocapsules structure [37]. Black dots
vitro cytotoxicity activity of Q, Q_SO3 and MNSXQ_SO3 against four were not seen deposited on the surface of carbon-coated copper grid,
human tumor cell lines: SNB-19 (glioblastoma), HCT-116 (colon carci­ evidencing the total encapsulation of magnetic nanoparticles into
noma), HL-60 (leukemia) and PC3 (prostate). More details of the sulfated xyloglucan nanocapsules [25]. The distribution curve particle

522
A.C.C. da Silva et al. International Journal of Biological Macromolecules 216 (2022) 520–527

Fig. 2. (a) FTIR spectrum and (b) DLS measurements of MNSXQ_SO3 sample.

Fig. 3. SEM micrographs of MNSXQ_SO3 sample.

Fig. 4. TEM micrographs and particle size distribution histogram of (a) Fe3O4@PAAS and (b) MNSXQ_SO3.

523
A.C.C. da Silva et al. International Journal of Biological Macromolecules 216 (2022) 520–527

size of magnetic nanoparticles (Fig. 4a) and nanocapsules (Fig. 4b) were models: zero-order, first-order, Higuchi and Korsmeyer-Peppas. Table 1
obtained by Log-normal fit of the size histogram. The values of 8 ± 2 nm shows the results after applying the mathematical models of kinetic
and 126 ± 27 nm were determined as mean particle size of magnetic study for the sample, indicating a release behavior compatible with
nanoparticles and nanocapsules, respectively. These results are similar Korsmeyer-Peppas kinetic model (R2 = 0.9737). In this case, the release
to those in previously published works [38,39]. of Q_SO3 from MNSXQ_SO3 is in accordance with the Fickian diffusion
The X-ray diffractogram (XRD) of MNSXQ_SO3 was displayed in (n = 0.4233) of Q_SO3 from nanocapsules. Literature data also report
Fig. 5a, showing characteristics of amorphous structures with no sharp that Korsmeyer-Peppas is the most commonly described semi-empirical
peaks. A peak at approximately 2 θ = 20◦ can be seen in the XRD of the drug release model for polyurethane nanocarriers obtained by the cross-
nanocarrier, similar to those reported by Arruda et al. (2015) for xylo­ linking reaction [45–47].
glucan obtained from Tamarind seeds. Although the XRD of Fe3O4@­
PAAS (Fig. 5b) reveals a pattern compatible with magnetite (JCPDS card
# 19-0629) phase [40], the typical crystallinity peaks of Fe3O4@PAAS 3.3. In vitro antitumor activity of the samples
are not exhibited in the XRD of MNSXQ_SO3 due to the successful
encapsulation of the ferrofluid [41]. The MTT assay was conducted to investigate cytotoxicity of
Fig. 6a shows the magnetization versus magnetic field dependence at MNSXQ_SO3, Q_SO3 and quercetin (Q), as drug control, against four
300 K, measured with a Vibrating Sample Magnetometer (VSM). cancer cell lines (Table 2). The results obtained in this experiment are
MNSXQ_SO3 sample reveals a superparamagnetic behavior with a expressed in IC50 values (μg mL− 1), illustrated in Fig. 7, and correspond
saturation magnetization (Ms) value of 4.8 emu/g, due the presence of to the mean ± SEM of at least three independent experiments performed
Fe3O4@PAAS into nanocapsules. This result is within the saturation in triplicate.
magnetization range published for magnetic nanocomposite based on According to Table 2, Quercetin (Q) showed lower IC50 values when
polymeric capsules [42]. The magnitude of Ms. is responsible for compared to Q_SO3, meaning that (Q) has a greater antitumor activity
theranostic properties, as the heating efficiency and magnetic respon­ than its sulfated derivative (Q_SO3). Sulfated flavonoids have interesting
siveness for nanocarrier guidance [43]. Finally, a small hysteresis loop is properties for clinical preparations, as the advantage of being water-
shown in the low-field data (inset at Fig. 6a). soluble. However, these sulfated derivatives may or may not magnify
or not the biological activities in relation to the non-sulfated compound
[48].
3.2. Release study of Q_SO3 Table 2 shows that Quercetin (Q) had greater antitumor activity
against HL-60, followed by HCT-116 and finally by SNB-19, with IC50
The release profile of hydrophilic quercetin sulfate (Q_SO3) free and values of 2.04 μg mL− 1, 3.86 μg mL− 1 and 11.27 μg mL− 1, respectively.
loaded in magnetic nanocapsules based on sulfated xyloglucan The antitumor effect of Q_SO3 was similarly ranked as HL-60 > HCT-
(MNSXQ_SO3) is displayed in Fig. 4b. The release assay was conducted in 116 > SNB-19 > PC3, with IC50 values of 28.99 μg mL− 1, 64.94 μg mL− 1,
PBS medium (pH 7.4) for 48 h, because the free Q_SO3 reached the 79.78 μg mL− 1 and 91.76 μg mL− 1, respectively. According to literature
complete amount diffusion (100 %) at this time, using a dual- data, the greater antitumor effect of quercetin compared to its sulfated
compartment system separated by a cellulose membrane with a molec­ derivative (quercetin-3′ -sulfate) was also previously reported in human
ular exclusion pore of 12 kDa [25]. During the first 3 h of the experi­ breast cancer MCF-7 cells. Both compounds led to a cell-cycle arrest at
ment, about 43 % of free Q_SO3 was diffused from donor to receptor the S phase and decreasing in the number of cells at G0/G1 and G2/M.
compartment, while 35 % of this bioactive was released from Afterwards, the flavonoids induced cells apoptosis, mediated by the
MNSXQ_SO3. After 48 h of experiment, the amount of free Q_SO3 generation of intracellular reactive oxygen species (ROS) [49].
released (100 %) is greater than the amount of Q_SO3 delivered by According to Mateus et al. (2018), the production of intracellular
MNSXQ_SO3 (76 %). Probably, the addition of sulfated groups to the ROS by quercetin occurs due a cellular redox imbalance caused by the
polysaccharide chains decreases the number of hydroxyls available for capacity of this flavonoid to be highly susceptible to autoxidation. Thus,
the crosslinking reaction. Consequently, the nanocomposite of sulfated the treatment with molecules that induce oxidative stress, as quercetin
xyloglucan exhibits a lower degree of reticulation, and it does not entrap and its sulfated derivative, provides a significant dose-dependent in­
Q_SO3 strongly enough during the release study. De Almeida et al. have crease of intracellular ROS in tumor cells, pushing them to their toxic
already published similar results, where they observed the faster release threshold, which can led to cell apoptosis through the mitochondrial
profile of mangiferin from sulfated polymeric matrix than from natural pathway [49–51].
polymeric matrix [44]. Literature data report that the increasing in intracellular ROS
The kinetic mechanism of Q_SO3 release from MNSXQ_SO3 was induced by quercetin can be mainly related to 3′ -OH, 4′ -OH groups on
evaluated fitting the data obtained above to various mathematical the B ring and 7-OH on the A ring. Then, the synthesis of quercetin

Fig. 5. XRD of (a) MNSXQ_SO3 and (b) Fe3O4@PAAS.

524
A.C.C. da Silva et al. International Journal of Biological Macromolecules 216 (2022) 520–527

Fig. 6. (a) Hysteresis loops of MNSXQ_SO3 sample with a zoom in low-field region inset and the (b) mean (±s.d.) percentage of free Q_SO3 (green line) and Q_SO3
encapsulated (pink line) in MNSXQ_SO3 sample. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of
this article.)

sulfate with substitutions at these positions would decrease this effect on


Table 1
the sulfated derivative in relation to the pattern compound [49,52].
Curve fitting analyses of kinetic studies obtained for Q_SO3 released from
In the present work, the chemical structure of Q_SO3 had three hy­
MNSXQ_SO3 in PBS medium (pH 7.4).
droxyls (5-OH, 3′ -OH and 4′ -OH) replaced by sulfate groups. Probably,
Sample Correlation coefficient (R2) Release
the absence of the –OH group in those positions in Q_SO3 decreased its
exponent (n)
Zero- First Higuchi Korsmeyer- cytotoxic activity when compared to Q, as seen in Table 2 and Fig. 5. To
order order Peppas avoid the quercetin sulfation reaction at those sites, a benzylation re­
Free Q_SO3 0.8995 0.5726 0.9826 0.9789 0.0716 action could be carried out to protect the 3′ , 4′ and 7 hydroxyl groups of
MNSXQ_SO3 0.9697 0.9544 0.9708 0.9737 0.4233 this compound [53]. Thereby, the sulfation reaction proposed in this
work could only form compounds with sulfate groups at positions 3 or 5.
According to Jang and Kang [53], the protective groups must be
Table 2 removed later by dissolution of the synthesized product in ethanol/THF
Antiproliferative activity of Q, Q_SO3 and MNSXQ_SO3 in four tumor cell lines (1:1, v/v), followed by the addition of Pd/C under stirring for 12 h.
evaluated by the MTT assay after 72 h treatment. Additionally, the results obtained in this research indicate that the
MTT (IC50 μg mL− 1) – 72 h encapsulation of Q_SO3 into magnetic nanocapsules based on sulfated
xyloglucan (MNSXQ_SO3) improves the antitumor activity of the free
Samples Cell lines
bioactive (Q_SO3) against all the four tested tumor cell lines. This result
SNB-19 HCT-116 HL-60 PC3
was similar to the one published by Huang et al., where the authors
(glioblastoma) (colorectal) (leukemia) (prostate)
demonstrated an increase of the doxorubicin inhibitory effects on the
Q 11.27 ± 2.76 3.86 ± 1.69 2.04 ± 1.02 n.d.a proliferation of human breast cancer (MCF-7) cells, when this bioactive
Q_SO3 79.78 ± 11.69 64.94 ± 5.89 28.99 ± 4.92 91.76 ±
13.85
was incorporated into polyurethane nanoparticles [54].
MNSXQ_SO3 8.77 ± 0.36 8.69 ± 0.21 1.58 ± 0.54 17.76 ± Nanocarriers based on polyurethane linkage are biocompatible and
3.42 demonstrate no significant cell toxicity [28,55]. Previous works re­
a
n.d.: not determined.
ported the biocompatibility of polyurethane nanoparticles, which was
assessed by HEK293 cells from the human embryonic kidney, demon­
strating a cell viability >75 % [45].
Although in vivo experiments have revealed an inflammatory state in
Swiss albino mice when the animals were treated with polyurethane
nanoparticles [56], in our previous work, polyurethane nanocomposites
based on starch provided no deaths in the in vivo acute toxicity assay
(zebrafish model) [47]. For those reasons, in vivo toxicity assays of
MNSXQ_SO3 will also be further performed to be included in future
studies.
According to results showed in Table 2, MNSXQ_SO3 exhibited better
antitumor activity against HL-60 (leukemia, IC50 = 1.58 μg mL− 1), fol­
lowed by SNB-9 (glioblastoma, IC50 = 8.77 μg mL− 1), HCT-116 (colo­
rectal, IC50 = 8.69 μg mL− 1), and PC3 (prostate, IC50 = 17.76 μg mL− 1).
Li et al. (2021) synthesized gold nanoparticles based on hyaluronic acid,
polyethylene glycol, and adipic dihydrazide for drug delivery of doxo­
rubicin. This nanocomposite shows better antitumor activity against
HCT-116 (IC50 = 54.6 μg mL− 1) and HepG2 (IC50 = 67.3 μg mL− 1) cells
Fig. 7. Antiproliferative activity of Q, Q_SO3 and MNSXQ_SO3 in four tumor [57].
cell lines evaluated by the MTT assay after 72 h treatment. Data are presented Data literature report that among the nanocomposites applied in
in IC50 values (μg mL− 1) and correspond to the mean ± SEM (n = 3). therapeutic approaches, polymeric nanoparticles are widely used for in

525
A.C.C. da Silva et al. International Journal of Biological Macromolecules 216 (2022) 520–527

vitro and in vivo experiments, due to their acceptable efficacy as a useful Appendix A. Supplementary data
carrier for targeted delivery of anticancer drugs [11]. The presence of
magnetic nanoparticles provides a great potential for diagnosis and Supplementary data to this article can be found online at https://doi.
therapy of several types of cancer [9]. Thus, the combination of these org/10.1016/j.ijbiomac.2022.06.197.
two nanocarrier can improve the theranostic potential of the nanoplat­
form. In the present work we developed hybrid nanocarriers References
(MNSXQ_SO3) and performed preliminary experiments to determine the
in vitro antitumor activity of this nanoplatform, highlighting the po­ [1] T.G. Barclay, C.M. Day, N. Petrovsky, S. Garg, Review of polysaccharide particle-
based functional drug delivery, Carbohydr. Polym. 221 (2019) 94–112, https://
tential of MNSXQ_SO3 for cancer treatment and encouraging further doi.org/10.1016/j.carbpol.2019.05.067.
investigation about the nanocomposite. [2] L. Huang, M. Shen, G.A. Morris, J. Xie, Sulfated polysaccharides:
immunomodulation and signaling mechanisms, Trends Food Sci. Technol. 92
(2019) 1–11, https://doi.org/10.1016/j.tifs.2019.08.008.
3.4. In vitro antioxidant activity of Q and Q_SO3 by DPPH radical [3] M. Andrew, G. Jayaraman, Marine sulfated polysaccharides as potential antiviral
scavenging experiment drug candidates to treat Corona Virus disease (COVID-19), Carbohydr. Res. 505
(2021) 1–18, https://doi.org/10.1016/j.carres.2021.108326.
[4] S. Iravani, R.S. Varma, Important roles of oligo-and polysaccharides against sars-
DPPH• is a stable free radical applied to estimate the free radical cov-2: recent advances, Appl. Sci. 11 (2021) 1–15, https://doi.org/10.3390/
scavenging activities of compounds in a relatively short time [8]. Sig­ app11083512.
nificant differences were obtained for Q and Q_SO3. The antioxidant [5] Ø. Arlov, D. Rütsche, M. Asadi Korayem, E. Öztürk, M. Zenobi-Wong, Engineered
sulfated polysaccharides for biomedical applications, Adv. Funct. Mater. 31 (2021)
capacity of Q was greater than Q_SO3, showing IC50 values of 6.03 ±
1–52, https://doi.org/10.1002/adfm.202010732.
0.41 μg mL− 1 (**p < 0.01) and 628.64 ± 52.84 μg mL− 1 (values [6] C.V. Pardeshi, A.D. Kulkarni, V.S. Belgamwar, S.J. Surana, Xyloglucan for Drug
expressed as the mean ± SEM, n = 6), respectively. As expected, quer­ Delivery Applications, Elsevier Ltd., 2018, https://doi.org/10.1016/B978-0-08-
cetin has well known view scavenging capacity, while the derivatization 102194-1.00007-4.
[7] N.L. Santos, R.C. Braga, M.S.R. Bastos, P.L.R. Cunha, F.R.S. Mendes, A.M.M.
of quercetin hydroxyl groups significantly reduces the antioxidant T. Galvão, G.S. Bezerra, A.A.C. Passos, Preparation and characterization of
ability [58]. Compounds with antioxidant properties are very important xyloglucan films extracted from Tamarindus indica seeds for packaging cut-up
to health body because they can accept free radicals and prevent many ‘Sunrise Solo’ papaya, Int. J. Biol. Macromol. 132 (2019) 1163–1175, https://doi.
org/10.1016/j.ijbiomac.2019.04.044.
diseases [14]. [8] Y. Cao, I. Ikeda, Antioxidant activity and antitumor activity (in vitro) of xyloglucan
selenious ester and surfated xyloglucan, Int. J. Biol. Macromol. 45 (2009) 231–235,
4. Conclusion https://doi.org/10.1016/j.ijbiomac.2009.05.007.
[9] A. Eftekhari, A. Arjmand, A. Asheghvatan, H. Švajdlenková, O. Šauša, H. Abiyev,
E. Ahmadian, O. Smutok, R. Khalilov, T. Kavetskyy, M. Cucchiarini, The potential
Quercetin sulfate (Q_SO3) was evaluated against four tumor cell application of magnetic nanoparticles for liver fibrosis theranostics, Front. Chem. 9
lines, using quercetin (Q) as control. Q_SO3 had its antitumor activity (2021) 1–15, https://doi.org/10.3389/fchem.2021.674786.
[10] Q. Tang, H. Lei, R. Wu, H. Fan, Y. Lü, Progress of magnetic hyperthermia based on
decreased, compared to Q, due the presence of sulfate groups in the magnetic nanomaterials, Kexue Tongbao/ChineseSci. Bull. 66 (2021) 3462–3473,
chemical structure of Q_SO3. Despite the lower cytotoxicity of Q_SO3, it https://doi.org/10.1360/TB-2020-1646.
is important to highlight that the sulfation reaction increases solubility, [11] E. Ahmadian, S.M. Dizaj, S. Sharifi, S. Shahi, R. Khalilov, A. Eftekhari,
M. Hasanzadeh, The potential of nanomaterials in theranostics of oral squamous
which might be an advantage for further in vivo studies. Q_SO3 also
cell carcinoma: recent progress, TrAC -Trends Anal. Chem. 116 (2019) 167–176,
shows lower antioxidant properties than Q. Sulfated xyloglucan (S_XLG) https://doi.org/10.1016/j.trac.2019.05.009.
was employed as a polymeric matrix of magnetic nanocomposites to [12] S.M. Dadfar, K. Roemhild, N.I. Drude, S. von Stillfried, R. Knüchel, F. Kiessling,
encapsulate quercetin sulfate (MNSXQ_SO3). The nanocarriers were T. Lammers, Iron oxide nanoparticles: diagnostic, therapeutic and theranostic
applications, Adv. Drug Deliv. Rev. 138 (2019) 302–325, https://doi.org/10.1016/
produced via interfacial polyaddition reaction through inverse mini­ j.addr.2019.01.005.
emulsion technique, forming nanocapsules with a core-shell [13] W.O. Pinheiro, M.L. Fascineli, G.R. Farias, F.H. Horst, L.R. de Andrade, L.H. Corrêa,
morphology, superparamagnetic characteristics, good colloidal stabil­ K.G. Magalhães, M.H. Sousa, M.C. de Almeida, R.B. Azevedo, Z.G.M. Lacava, The
influence of female mice age on biodistribution and biocompatibility of citrate-
ity and able to entrap hydrophilic bioactives as Q_SO3. MNSXQ_SO3 coated magnetic nanoparticles, Int. J. Nanomedicine 14 (2019) 3375–3388,
delayed the Q_SO3 delivery, when compared to the free Q_SO3, https://doi.org/10.2147/IJN.S197888.
emphasizing the ability of the sulfated polysaccharide to modify the [14] X. Song, Y. Wang, L. Gao, Mechanism of antioxidant properties of quercetin and
quercetin-DNA complex, J. Mol. Model. 26 (2020) 1–8, https://doi.org/10.1007/
release profile of bioactives. Aditionally, MNSXQ_SO3 showed better s00894-020-04356-x.
antitumor activity than Q_SO3. Synthesis and study of new magnetic [15] E. Amanzadeh, A. Esmaeili, R.E.N. Abadi, N. Kazemipour, Z. Pahlevanneshan,
nanocomposites might reveal the promising candidates for future S. Beheshti, Quercetin conjugated with superparamagnetic iron oxide
nanoparticles improves learning and memory better than free quercetin via
application in theranostic area as drug delivery of antitumor drugs or interacting with proteins involved in LTP, Sci. Rep. 9 (2019) 1–19, https://doi.org/
diagnostic and treatment of diseases. 10.1038/s41598-019-43345-w.
[16] C.S.Dos Santos, N.Osti Silva, J.B. dos Santos Espinelli, M.A.Germani Marinho, Z.
Vieira Borges, N.Bruzamarello Caon Branco, F.L. Faita, B.Meira Soares, A.P. Horn,
Acknowledgments
A.L. Parize, V.Rodrigues de Lima, Molecular interactions and physico-chemical
characterization of quercetin-loaded magnetoliposomes, Chem. Phys. Lipids 218
This work received financial support in part by Coordenação de (2019) 22–33, https://doi.org/10.1016/j.chemphyslip.2018.11.010.
Aperfeiçoamento de Pessoal de Nível Superior – Brasil (CAPES) – [17] H.G. Ulusoy, N. Sanlier, A minireview of quercetin: from its metabolism to possible
mechanisms of its biological activities, Crit. Rev. Food Sci. Nutr. 60 (2020)
Finance Code 001, Conselho Nacional de Desenvolvimento Científico e 3290–3303, https://doi.org/10.1080/10408398.2019.1683810.
Tecnológico (CNPq) and Fundação Cearense de Apoio ao Desenvolvi­ [18] N. López, I. Delso, D. Matute, C. Lafuente, M. Artal, Characterization of xylitol or
mento Científico e Tecnológico (FUNCAP). The authors would like to citric acid:choline chloride:water mixtures: structure, thermophysical properties,
and quercetin solubility, Food Chem. 306 (2020) 1–9, https://doi.org/10.1016/j.
thank for the research grants of N.M.P.S.R. (CNPq Project No 309795/ foodchem.2019.125610.
2021-4), A.C.C.S. (CNPq Project No 140256/2017-2) and R.R.A. (FUN­ [19] A. Boretti, Quercetin supplementation and COVID-19, Nat. Prod. Commun. 16
CAP Project No 88887.162539/2017-00). We also acknowledge the (2021) 1–3, https://doi.org/10.1177/1934578X211042763.
[20] A. Massi, O. Bortolini, D. Ragno, T. Bernardi, G. Sacchetti, M. Tacchini, C. De Risi,
analyses and data acquisition provided by Northeastern Center for the Research progress in the modification of quercetin leading to anticancer agents,
Application and use of Nuclear Magnetic Resonance (CENAUREM-UFC), Molecules 22 (2017) 1270, https://doi.org/10.3390/molecules22081270.
Brazilian Agricultural Research Corporation (EMBRAPA), Central Ana­ [21] B. Kang, P. Okwieka, S. Schöttler, O. Seifert, R.E. Kontermann, K. Pfizenmaier,
A. Musyanovych, R. Meyer, M. Diken, U. Sahin, V. Mailänder, F.R. Wurm,
lítica-UFC/CT-INFRA/MCTI-SISNANO/CAPES, Fondecyt 1200782 and K. Landfester, Tailoring the stealth properties of biocompatible polysaccharide
CEDENNA AFB180001 (USACH). nanocontainers, Biomaterials 49 (2015) 125–134, https://doi.org/10.1016/j.
biomaterials.2015.01.042.

526
A.C.C. da Silva et al. International Journal of Biological Macromolecules 216 (2022) 520–527

[22] R.B. Souza, A.F. Frota, J. Silva, C. Alves, A.Z. Neugebauer, S. Pinteus, J.A. [40] X. Hu, Y. Wang, L. Zhang, M. Xu, J. Zhang, W. Dong, Design of a pH-sensitive
G. Rodrigues, E.M.S. Cordeiro, R.R. de Almeida, R. Pedrosa, N.M.B. Benevides, In magnetic composite hydrogel based on salecan graft copolymer and Fe3O4@SiO2
vitro activities of kappa-carrageenan isolated from red marine alga Hypnea nanoparticles as drug carrier, Int. J. Biol. Macromol. 107 (2018) 1811–1820,
musciformis: antimicrobial, anticancer and neuroprotective potential, Int. J. Biol. https://doi.org/10.1016/j.ijbiomac.2017.10.043.
Macromol. 112 (2018) 1248–1256, https://doi.org/10.1016/j. [41] P.H.P., H.S. Mahajan, Mixed micelles for bioavailability enhancement of nelfinavir
ijbiomac.2018.02.029. mesylate: in vitro characterisation and In vivo pharmacokinetic study, Mater.
[23] G. Baier, A. Musyanovych, M. Dass, S. Theisinger, K. Landfester, Cross-linked Technol. 33 (2018) 793–802, https://doi.org/10.1080/10667857.2018.15113.
starch capsules containing dsDNA prepared in inverse miniemulsion as [42] R. Grillo, J. Gallo, D.G. Stroppa, E. Carbó-Argibay, R. Lima, L.F. Fraceto,
“nanoreactors” for polymerase chain reaction, Biomacromolecules 11 (2010) M. Bañobre-López, Sub-micrometer magnetic nanocomposites: insights into the
960–968, https://doi.org/10.1021/bm901414k. effect of magnetic nanoparticles interactions on the optimization of SAR and MRI
[24] X. Jin, X. Chen, Y. Cheng, L. Wang, B. Hu, J. Tan, Effects of hydrothermal performance, ACS Appl. Mater. Interfaces 8 (2016) 25777–25787, https://doi.org/
temperature and time on hydrothermal synthesis of colloidal hydroxyapatite 10.1021/acsami.6b08663.
nanorods in the presence of sodium citrate, J. Colloid Interface Sci. 450 (2015) [43] J. Wang, M. Fan, X. Bian, M. Yu, T. Wang, S. Liu, Y. Yang, Y. Tian, R. Guan,
151–158, https://doi.org/10.1016/j.jcis.2015.03.010. Enhanced magnetic heating efficiency and thermal conductivity of magnetic
[25] R.R. De Almeida, J. Gallo, A.C.C. Da Silva, A.K.O. Da Silva, O.D.L. Pessoa, T. nanofluids with FeZrB amorphous nanoparticles, J. Magn. Magn. Mater. 465
G. Araújo, L.K.A.M. Leal, P.B.A. Fechine, M. Bañobre-López, N.M.P.S. Ricardo, (2018) 480–488, https://doi.org/10.1016/j.jmmm.2018.06.043.
Preliminary evaluation of novel triglyceride-based nanocomposites for biomedical [44] R.R. de Almeida, H.S. Magalhães, J.R.R. de Souza, M.T.S. Trevisan, Í.G.P. Vieira, J.
applications, J. Braz. Chem. Soc. 28 (2017) 1547–1556, https://doi.org/10.21577/ P.A. Feitosa, T.G. Araújo, N.M.P.S. Ricardo, Exploring the potential of
0103-5053.20170007. Dimorphandra gardneriana galactomannans as drug delivery systems, Ind. Crop.
[26] T. Mosmann, Rapid colorimetric assay for cellular growth and survival: application Prod. 69 (2015) 284–289, https://doi.org/10.1016/j.indcrop.2015.02.041.
to proliferation and cytotoxicity assays, J. Immunol. Methods 65 (1983) 55–63, [45] B.S. Eftekhari, A. Karkhaneh, A. Alizadeh, Physically targeted intravenous
https://doi.org/10.1039/c6ra17788c. polyurethane nanoparticles for controlled release of atorvastatin calcium, Iran.
[27] W. Brand-Williams, M.E. Cuvelier, C. Berset, Use of a free radical method to Biomed. J. 21 (2017) 369–379, https://doi.org/10.18869/acadpub.ibj.21.6.369.
evaluate antioxidant activity, LWT - Food Sci. Technol. 28 (1995) 25–30, https:// [46] S.A. Pereira, S.B.F. dos Santos, F.A.M. Rodrigues, L.M.U.D. Fechine, Í.G.P. Vieira, J.
doi.org/10.1016/S0023-6438(95)80008-5. C. Denardin, F. Araneda, M.E.N.P. Ribeiro, N.V. Gramosa, N.M.P.S. Ricardo,
[28] S.K. Pramanik, S. Seneca, M. Peters, L. D’Olieslaeger, G. Reekmans, Hydroxyethyl starch nanocapsules by multiple nanoemulsions for carrying and
D. Vanderzande, P. Adriaensens, A. Ethirajan, Morphology-dependent pH- controlled release of lapachol, Mater. Lett. 274 (2020) 1–3, https://doi.org/
responsive release of hydrophilic payloads using biodegradable nanocarriers, RSC 10.1016/j.matlet.2020.127983.
Adv. 8 (2018) 36869–36878, https://doi.org/10.1039/C8RA07066K. [47] A.C.C. Sousa, A.I.B. Romo, R.R. Almeida, C.C. Silva, L.M.U. Fechine, H.A. Brito, R.
[29] F.R. Steinmacher, G. Baier, A. Musyanovych, K. Landfester, P.H.H. Araújo, M. Freire, D.P. Pinheiro, M. Larissa, M.P.S. Ricardo, R. Silva, O.D.L. Pessoa, J.
C. Sayer, Design of cross-linked starch nanocapsules for enzyme-triggered release C. Denardin, C. Pessoa, Starch-based magnetic nanocomposite for targeted delivery
of hydrophilic compounds, Processes 5 (2017) 25, https://doi.org/10.3390/ of hydrophilic bioactives as anticancer strategy, Carbohydr. Polym. 264 (2021)
pr5020025. 1–9, https://doi.org/10.1016/j.carbpol.2021.118017.
[30] P.R. Babu, M.T. Mathai, A. Julius, V. Madhu, Dual property of tamarind seed [48] C. Noleto-Dias, C. Harflett, M.H. Beale, J.L. Ward, Sulfated flavanones and
polysaccharide aid wound healing, Int. J. Adv. Sci. Technol. 28 (2019) 1130–1141. dihydroflavonols from willow, Phytochem. Lett. 35 (2020) 88–93, https://doi.org/
http://sersc.org/journals/index.php/IJAST/article/view/3500. 10.1016/j.phytol.2019.11.008.
[31] S.S. Leong, W.M. Ng, J.K. Lim, S.P. Yeap, Dynamic light scattering: effective sizing [49] Q. Wu, P.W. Needs, Y. Lu, P.A. Kroon, D. Ren, X. Yang, Different antitumor effects
technique for characterization of magnetic nanoparticles, in: Handb. Mater. of quercetin, quercetin-3′ -sulfate and quercetin-3-glucuronide in human breast
Charact, Springer, 2018, pp. 77–111, https://doi.org/10.1007/978-3-319-92955- cancer MCF-7 cells, Food Funct. 9 (2018) 1736–1746, https://doi.org/10.1039/
2_3. c7fo01964e.
[32] V.P. Chakka, T. Zhou, Carboxymethylation of polysaccharides: synthesis and [50] P.G. Mateus, V.G. Wolf, M.S. Borges, V.F. Ximenes, Quercetin: Prooxidant Effect
bioactivities, Int. J. Biol. Macromol. 165 (2020) 2425–2431, https://doi.org/ and Apoptosis in Cancer, 1st ed., Elsevier B.V., 2018 https://doi.org/10.1016/
10.1016/j.ijbiomac.2020.10.178. B978-0-444-64056-7.00009-X.
[33] C. Wang, Y. He, X. Tang, N. Li, Sulfation, structural analysis, and anticoagulant [51] H. Zhang, M. Zhang, L. Yu, Y. Zhao, N. He, X. Yang, Antitumor activities of
bioactivity of ginger polysaccharides, J. Food Sci. 85 (2020) 2427–2434, https:// quercetin and quercetin-5’,8-disulfonate in human colon and breast cancer cell
doi.org/10.1111/1750-3841.15338. lines, Food Chem. Toxicol. 50 (2012) 1589–1599, https://doi.org/10.1016/j.
[34] N. Saxena, H. Agraval, K.C. Barick, D. Ray, V.K. Aswal, A. Singh, U.C.S. Yadav, C. fct.2012.01.025.
L. Dube, Thermal and microwave synthesized SPIONs: energy effects on the [52] Y. Zhang, D. Wang, L. Yang, D. Zhou, J. Zhang, Purification and characterization of
efficiency of nano drug carriers, Mater. Sci. Eng. C 111 (2020), https://doi.org/ flavonoids from the leaves of Zanthoxylum bungeanum and correlation between
10.1016/j.msec.2020.110792. their structure and antioxidant activity, PLoS One 9 (2014) 1–11, https://doi.org/
[35] C. Cano-Sarmiento, D.I. Téllez-Medina, R. Viveros-Contreras, M. Cornejo-Mazón, C. 10.1371/journal.pone.0105725.
Y. Figueroa-Hernández, E. García-Armenta, L. Alamilla-Beltrán, H.S. García, G. [53] J. Jang, D.W. Kang, Synthesis of monogallic acid conjugates of quercetin at the 3-,
F. Gutiérrez-López, Zeta potential of food matrices, Food Eng. Rev. 10 (2018) 7-, and 4′ -hydroxyl groups through selective protection of hydroxyl groups,
113–138, https://doi.org/10.1007/s12393-018-9176-z. J. Korean Chem. Soc. 63 (2019) 138–143, https://doi.org/10.5012/
[36] I. Antal, M. Kubovcikova, V. Zavisova, M. Koneracka, O. Pechanova, A. Barta, jkcs.2019.63.2.138.
M. Cebova, V. Antal, P. Diko, M. Zduriencikova, M. Pudlak, P. Kopcansky, [54] D. Huang, Y. Zhou, Y. Xiang, M. Shu, H. Chen, B. Yang, X. Liao, Polyurethane/
Magnetic poly(D, L-lactide) nanoparticles loaded with aliskiren: a promising tool doxorubicin nanoparticles based on electrostatic interactions as pH-sensitive drug
for hypertension treatment, J. Magn. Magn. Mater. 380 (2015) 280–284, https:// delivery carriers, Polym. Int. 67 (2018) 1186–1193, https://doi.org/10.1002/
doi.org/10.1016/j.jmmm.2014.10.089. pi.5618.
[37] A.M. Atta, G.A. El-Mahdy, H.A. Al-Lohedan, A.M. El-Saeed, Preparation and [55] K. Błażek, J. Datta, Renewable natural resources as green alternative substrates to
application of crosslinked poly(sodium acrylate)-coated magnetite nanoparticles as obtain bio-based non-isocyanate polyurethanes-review, Crit. Rev. Environ. Sci.
corrosion inhibitors for carbon steel alloy, Molecules 20 (2015) 1244–1261, Technol. 49 (2019) 173–211, https://doi.org/10.1080/10643389.2018.1537741.
https://doi.org/10.3390/molecules20011244. [56] A.H. Silva, C. Locatelli, F.B. Filippin-monteiro, P. Martin, N.J. Liptrott, B.
[38] L.I. Ronco, P.E. Feuser, A. da Cas Viegas, R.J. Minari, L.M. Gugliotta, C. Sayer, P.H. G. Zanetti-ramos, L.C. Benetti, E.M. Nazari, C.A.C. Albuquerque, A.A. Pasa,
H. Araújo, Incorporation of magnetic nanoparticles in poly(methyl methacrylate) A. Owen, T.B. Creczynski-pasa, Toxicity and inflammatory response in Swiss albino
nanocapsules, Macromol. Chem. Phys. 219 (2018) 1–7, https://doi.org/10.1002/ mice after intraperitoneal and oral administration of polyurethane nanoparticles,
macp.201700424. Toxicol. Lett. 246 (2016) 17–27, https://doi.org/10.1016/j.toxlet.2016.01.018.
[39] D.M.A. Neto, R.M. Freire, J. Gallo, T.M. Freire, D.C. Queiroz, N.M.P.S. Ricardo, I. [57] L. Li, B. Ren, X. Yang, Z. Cai, X. Zhao, M. Zhao, Hyaluronic acid-modified and
F. Vasconcelos, G. Mele, L. Carbone, S.E. Mazzetto, M. Bañobre-López, P.B. doxorubicin-loaded gold nanoparticles and evaluation of their bioactivity,
A. Fechine, Rapid sonochemical approach produces functionalized Fe3O4 Pharmaceuticals 14 (2021) 101–113.
nanoparticles with excellent magnetic, colloidal, and relaxivity properties for MRI [58] M. Lesjak, I. Beara, N. Simin, D. Pintać, T. Majkić, K. Bekvalac, D. Orčić, N. Mimica-
application, J. Phys. Chem. C 121 (2017) 24206–24222, https://doi.org/10.1021/ Dukić, Antioxidant and anti-inflammatory activities of quercetin and its
acs.jpcc.7b04941. derivatives, J. Funct. Foods 40 (2018) 68–75, https://doi.org/10.1016/j.
jff.2017.10.047.

527

You might also like