Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Carbohydrate Polymers 200 (2018) 93–99

Contents lists available at ScienceDirect

Carbohydrate Polymers
journal homepage: www.elsevier.com/locate/carbpol

Hyaluronic acid functionalized green reduced graphene oxide for targeted T


cancer photothermal therapy
Rita Lima-Sousaa, Duarte de Melo-Diogoa, Cátia G. Alvesa, Elisabete C. Costaa, Paula Ferreirab,

Ricardo O. Louroc, Ilídio J. Correiaa,b,
a
CICS-UBI – Centro de Investigação em Ciências da Saúde, Universidade da Beira Interior, 6200-506 Covilhã, Portugal
b
CIEPQPF – Departamento de Engenharia Química, Universidade de Coimbra, Rua Silvio Lima, 3030-790 Coimbra, Portugal
c
ITQB – Instituto de Tecnologia Química e Biológica António Xavier, Universidade Nova de Lisboa, 2780-157 Oeiras, Portugal

A R T I C LE I N FO A B S T R A C T

Keywords: Reduced graphene oxide (rGO) nanomaterials display promising properties for application in cancer photo-
2D materials thermal therapy (PTT). rGO is usually obtained by treating graphene oxide (GO) with hydrazine hydrate.
Green chemistry However, this reducing agent contributes for the low cytocompatibility exhibited by rGO. Furthermore, rGO has
Hyaluronan a low water stability and does not show selectivity towards cancer cells. Herein, rGO attained using an en-
Photothermal effect
vironmentally-friendly method was functionalized with a novel hyaluronic acid (HA)-based amphiphilic polymer
Targeted therapy
to be used in targeted cancer PTT. Initially, the green-reduction of GO with L-Ascorbic acid was optimized
considering the near infrared absorption and the size distribution of the nanomaterials. Then, rGO was func-
tionalized with the HA-based amphiphile. The functionalization of rGO improved its stability, cytocompatibility
and internalization by CD44 overexpressing cells, which indicates the targeting capacity of this nanoformulation.
Furthermore, the on-demand PTT mediated by HA-functionalized rGO induced cancer cells’ ablation, thereby
confirming its potential for targeted cancer therapy.

1. Introduction high NIR absorption are the most appealing for application in cancer
PTT, since they may produce an on-demand therapeutic effect upon
Photothermal therapy (PTT) mediated by nanomaterials has been interaction with NIR light (Akhavan, Ghaderi, Aghayee, Fereydoonia, &
showing promising results in cancer treatment (de Melo-Diogo, Pais- Talebia, 2012; Akhavan, Ghaderi, & Emamy, 2012; de Melo-Diogo,
Silva, Costa, Louro, & Correia, 2017; de Melo-Diogo, Pais-Silva, Dias, Pais-Silva, Costa et al., 2017).
Moreira, & Correia, 2017). This type of therapy benefits from the ability Among the different NIR-responsive nanomaterials, nanosized gra-
of nanostructures to accumulate in the tumor through the so called phene oxide (GO) has sparked a great interest for cancer PTT (Chung
enhanced permeability and retention effect (EPR effect) (Fang, et al., 2013; Li, Tang, Yuan, Sun, & Wang, 2013). This 2D carbon ma-
Nakamura, & Maeda, 2011; Maeda, 2012). Afterwards, the tumor zone terial has a graphitic lattice with several types of oxygen functional
is irradiated with laser light, and the nanomaterials used for this type of groups (e.g. epoxy, hydroxyl or carboxyl groups) and presents NIR ab-
therapy absorb laser light energy and convert it into heat (Frazier & sorption (Chen, Feng, & Li, 2012; Liu, Cui, & Losic, 2013). The reduc-
Ghandehari, 2015). For photothermal applications, the use of near in- tion of GO can further improve its NIR absorption, by restoring the
frared light (NIR; 750–1000 nm) is crucial due to its low interaction aromatic lattice of this material, leading to an improved photothermal
with biological components, thus achieving a high penetration depth capacity (Zhang, Yang et al., 2010). Hydrazine hydrate (HH) is the most
(Strangman, Boas, & Sutton, 2018). In this way, nanostructures with a commonly used agent to produce reduced GO (rGO) (Hu, Sun, Ding,

Abbreviations: ANOVA, analysis of variance; CLSM, confocal laser scanning microscopy; DLS, dynamic light scattering; DMEM-F12, Dulbecco’s modified Eagle’s
medium-F-12; DMSO, dimethyl sulfoxide; EDC, 1-ethyl-3-(3-dimethylaminopropyl) carbodiimide; EPR, enhanced permeability and retention; FBS, fetal bovine
serum; FTIR, fourier transform infrared spectroscopy; GO, graphene oxide; HA, hyaluronic acid; HA-g-PMAO, HA-grafted PMAO; HA-rGO, HA-g-PMAO functionalized
rGO; HH, hydrazine hydrate; IC50, half maximal inhibitory concentration; LAA, L-ascorbic acid; MCF-7, Michigan cancer foundation-7; NIR, near infrared; NHDF,
normal human dermal fibroblasts; NHS, N-hydroxysuccinimide; PBS, phosphate buffered saline; PEG, poly(ethylene glycol); PMAO, poly(maleic anhydride-alt-1-
octadecene); PTT, photothermal therapy; rGO, reduced graphene oxide; Rhod B, Rhodamine B; SD, standard deviation; TEM, transmission electron microscopy

Corresponding author at: CICS-UBI – Centro de Investigação em Ciências da Saúde, Universidade da Beira Interior, 6200-506 Covilhã, Portugal.
E-mail address: icorreia@ubi.pt (I.J. Correia).

https://doi.org/10.1016/j.carbpol.2018.07.066
Received 25 May 2018; Received in revised form 23 July 2018; Accepted 23 July 2018
Available online 24 July 2018
0144-8617/ © 2018 Elsevier Ltd. All rights reserved.
R. Lima-Sousa et al. Carbohydrate Polymers 200 (2018) 93–99

Chen, & Chen, 2016; Robinson et al., 2011). However, the rGO obtained 2.2. Methods
by using HH is highly cytotoxic, in part due to the inherent toxicity of
this hazardous reducing agent (Robinson et al., 2011). Therefore, it is 2.2.1. Optimization of the green reduction of GO
imperative to develop and optimize alternative methods for the re- The optimization of the environmentally-friendly reduction of GO
duction of GO. was performed by adapting previously described methods (Fernández-
L-Ascorbic acid (LAA), or Vitamin C, is a water soluble natural Merino et al., 2010; Zhang, Yang et al., 2010). GO was sonicated for 6 h
compound with reducing capacities (De Silva, Huang, Joshi, & before its use. In brief, 1 mL solutions containing GO (200 μg/mL) and
Yoshimura, 2017; Du, Cullen, & Buettner, 2012). This molecule has LAA (3 mM or 1.5 mM) were reacted at 80 °C during 30, 45, 60, 90 or
revealed promising results in the reduction of GO (Huang et al., 2015; 120 min. Afterwards, the reduction process was stopped by cooling the
Liu et al., 2012). However, the conditions for the LAA-mediated re- samples in an ice-water bath.
duction of GO are yet to be optimized considering the NIR absorption
and the size distribution of the obtained materials. This is of paramount 2.2.2. Functionalization of rGO with HA-g-PMAO
importance, since the potential of rGO-based nanomaterials in cancer For the functionalization of rGO, this material (200 μg/mL, 1 mL)
therapy is highly dependent on these properties. was mixed with HA-g-PMAO (500 μg; the synthesis of HA-g-PMAO is
Furthermore, the direct use of rGO in cancer therapy is also im- described in the supplementary information) and sonicated for 60 min
paired by its poor water solubility (Hashemi et al., 2017; Liu et al., (Branson 5800, Branson Ultrasonics, CT, USA). Afterward, this solution
2011; Yang et al., 2012) and lack of selectivity towards cancer cells was dialyzed against water (14 kDa cut-off dialysis membrane) for
(Robinson et al., 2011). To surpass the solubility problems, rGO can be 90 min to remove LAA and centrifuged to remove any aggregates. The
functionalized with amphiphilic polymers through non-covalent inter- supernatant was then recovered, yielding HA-g-PMAO functionalized
actions (Li et al., 2014). For instance, the functionalization of rGO with rGO (HA-rGO).
poly(ethylene glycol) (PEG)-based amphiphilic polymers can greatly
improve their water solubility (Robinson et al., 2011). Moreover, this 2.2.3. Characterization of rGO and HA-rGO-based materials
type of functionalization can also improve the biocompatibility of rGO The reduction efficiency of GO was monitored by UV–vis absorption
materials (Li et al., 2014; Miao et al., 2013; Robinson et al., 2011). spectroscopy on an Evolution 201 spectrophotometer (Thermo
However, the majority of the amphiphilic polymers that have been used Scientific Inc.), over the wavelength range from 200 to 1000 nm, and by
to functionalize rGO do not possess targeting capacity. Therefore, in- dynamic light scattering (DLS) in a Zetasizer Nano ZS (Malvern
vestigating amphiphilic polymers with cancer cell targeting properties Instruments, Worcestershire, UK), at a scattering angle of 173°. The
is of upmost interest in order to produce rGO materials capable of successful functionalization of rGO with HA-g-PMAO was confirmed by
performing a selective PTT. DLS, UV–vis absorption spectroscopy and Fourier transform infrared
In this work, we hypothesized that the concentration of LAA and the spectroscopy (FTIR), using a Nicolet iS10 spectrometer (Thermo
reaction time used in the environmentally-friendly reduction of GO Scientific Inc., MA, USA). Furthermore, HA-rGO nanosized dimensions
could be fine-tuned in order to yield rGO with a suitable size and NIR were determined by transmission electron microscopy (TEM), using a
absorption for cancer-related applications. Furthermore, it was also HT7700 microscope (Hitachi, Japan), operated at an accelerating vol-
hypothesized that the functionalization of rGO with a novel hyaluronic tage of 100 kV. For this purpose, samples were stained with phospho-
acid (HA)-based amphiphilic polymer, produced by grafting HA onto tungstic acid (2% (w/v)). The photothermal capacity of HA-rGO was
poly(maleic anhydride-alt-1-octadecene) (PMAO), could render nanos- assessed by irradiating the samples with NIR radiation (808 nm, 1.7 W/
tructures with an improved stability and cytocompatibility, capable of cm2) over a period of 5 min, and the temperature changes were mon-
performing a targeted cancer PTT. In this regard, HA was selected due itored with a thermocouple thermometer.
to its hydrophilic character and targeting capacity to the CD44 re-
ceptors, which are overexpressed on cancer cells’ membrane (Calcagno 2.2.4. Evaluation of the cytocompatibility of rGO and HA-rGO
et al., 2010; Olsson et al., 2011). Additionally, CD44 receptors are in a The cytotoxic profile of rGO and HA-rGO was determined in MCF-7
quiescent state on normal cells, i.e. they do not display binding capacity cells and NHDF as we previously reported (Gaspar et al., 2015). In brief,
to HA (Yang et al., 2015), which further highlights the therapeutic 96-well plates were seeded with 1 × 104 cells/well in DMEM-F12
potential of targeting this receptor. To produce the amphiphilic medium supplemented with 10% FBS and 1% of penicillin/strepto-
polymer, HA will be deacetylated in order to be grafted onto PMAO, mycin, and were incubated in a humified incubator at 37 °C, 5 % CO2.
due to the capacity of the later to adsorb on rGO surface (Shi, Chen, After 24 h, the medium was replaced by fresh cell culture medium
Ehlerding, & Cai, 2014; Yang et al., 2012). containing rGO or HA-rGO at different concentrations, and cells were
incubated for 24 and/or 48 h. After this period, the medium was re-
placed with fresh cell culture medium containing resazurin (10% (v/v))
2. Experimental section and cells were incubated for 4 h in the dark (37 °C, 5% CO2). Cells’
viability was then assessed by analyzing the fluorescence of resorufin
2.1. Materials (λex = 560 nm; λem = 590 nm) in a Spectramax Gemini EM spectro-
fluorometer (Molecular Devices LLC, CA, USA). Cells solely incubated
Dimethyl sulfoxide (DMSO) and LAA were purchased from Fisher with culture medium (without rGO derivatives) were used as negative
Scientific (Oeiras, Portugal). Dulbecco’s modified Eagle’s medium-F-12 (K-) control, whereas those treated with ethanol (70% (v/v)) represent
(DMEM-F12), GO nanocolloids, N-hydroxysuccinimide (NHS), paraf- the positive (K+) control.
ormaldehyde, PMAO (30000–50000 Da), penicillin/streptomycin and
resazurin were acquired from Sigma–Aldrich (Sintra, Portugal). HA 2.2.5. Investigation of the targeting capacity of HA-rGO
sodium salt (9.27 kDa) was obtained from Carbosynth (Berkshire, UK). To evaluate HA-rGO cellular uptake, this nanomaterial was labelled
1-Ethyl-3-(3-dimethylaminopropyl) carbodiimide (EDC) was purchased with rhodamine B (Rhod B) as previously described elsewhere (Zhang,
from Merck (Darmstadt, Germany). Michigan cancer foundation-7 Xia, Zhao, Liu, and Zhang, 2010). In brief, HA-rGO (200 μg/mL, 1 mL)
(MCF-7) cell line was acquired from ATCC (Middlesex, UK). Normal was sonicated with Rhod B for 30 min. Afterwards, the solution was
human dermal fibroblasts (NHDF) were obtained from Promo-Cell dialyzed against water for 90 min (1000 Da cut-off dialysis membrane)
(Heidelberg, Germany). Fetal bovine serum (FBS) was purchased from to remove the non-bounded Rhod B and centrifuged to remove any
Biochrom AG (Berlin, Germany). Water used in all experiments was bigger aggregates.
double deionized (0.22 μm filtered; 18.2 MΩ cm). To confirm the targeting capacity of HA-rGO to CD44

94
R. Lima-Sousa et al. Carbohydrate Polymers 200 (2018) 93–99

overexpressing cancer cells, the uptake of Rhod B labelled HA-rGO by 3.2. Functionalization of rGO with HA-g-PMAO
MCF-7 cells (overexpressing CD44) (Calcagno et al., 2010; Olsson et al.,
2011) and NHDF (that do not overexpress CD44) (Miyatake et al., A simple sonication method was used to functionalize rGO with HA-
2013) was analyzed by confocal laser scanning microscopy (CLSM). g-PMAO (the successful synthesis of HA-g-PMAO was confirmed by FTIR
Briefly, μ-slide 8-well imaging plates (Ibidi GmbH, Munich, Germany) (Fig. S2 and S3), 1H NMR (Fig. S4 and S5) and TGA analysis (Fig. S6) -
were seeded with 1.5 × 104 cells/well. After 48 h, cells were incubated reported in the supplementary information). In this process, the hy-
with cell culture medium containing Rhod B labelled HA-rGO (68.7 μg/ drophobic alkyl tail of HA-g-PMAO binds to the aromatic lattice of rGO
mL of Rhod B equivalents) or free Rhod B (68.7 μg/mL) during 4 h. through non-covalent interactions (hydrophobic-hydrophobic interac-
Subsequently, cells were fixed with paraformaldehyde 4% for 15 min, tions). Moreover, the HA segments form a hydrophilic corona, which
and then rinsed with a phosphate buffered saline (PBS) solution. Zeiss may confer a cancer cell targeting capacity to this nanoformulation
LSM 710 confocal microscope (Carl Zeiss AG, Oberkochen, Germany) (Fig. 1A).
was used for imaging experiments (λex = 514 nm; λem = 513–703 nm). The successful functionalization of rGO with HA-g-PMAO (HA-rGO)
Non-treated cells and cells incubated with free-Rhod B were used as was confirmed through FTIR analysis (Fig. 1B). Before functionaliza-
controls. tion, rGO displays a peak at 1614 cm−1 (C]C stretch) and one with a
low intensity at 1737 cm−1 (C]O stretch), which also confirm the
restoration of the graphitic matrix of this material upon reduction.
2.2.6. In vitro evaluation of the phototherapeutic effect mediated by HA- Furthermore, the peak at 3258 cm−1 (OeH stretch) suggests that some
rGO LAA traces are still present in rGO. The FTIR spectrum of HA-rGO
The photothermal effect mediated by HA-rGO was evaluated by presents peaks at 2919 cm−1 (CeH stretch) and 1730 cm−1 (C]O
using a previously described protocol (de Melo-Diogo, Pais-Silva, Costa stretch), which belong to the alkyl and carbonyl groups of HA-g-PMAO,
et al., 2017). In brief, 96-well plates were seeded with MCF-7 cells at a respectively, thus confirming the successful preparation of HA-rGO.
density of 1 × 104 cells/well. After 24 h, the medium was replaced by The DLS analysis demonstrated that the functionalization of rGO
fresh medium containing different concentrations of HA-rGO (25, 50 with HA-g-PMAO did not impact on its size distribution (Fig. 1C). In
and 75 μg/mL of rGO equivalents). After 4 h, cells were irradiated with fact, when compared to rGO, HA-rGO presents a monomodal size dis-
NIR light (808 nm, 1.7 W/cm2, 5 min). After 24 h of incubation, cells’ tribution, which suggests that the HA-based coating can improve the
viability was determined through the resazurin assay as described stability of the nanostructures. The nanosized dimensions of HA-rGO
above. were then confirmed by TEM (Fig. S7), revealing that these materials
have an average lateral size of 108 ± 51 nm. These values are within
the optimal range for passive tumor accumulation through the EPR
2.2.7. Statistical analysis effect (Hobbs et al., 1998). The zeta potential of HA-rGO was de-
One-way analysis of variance (ANOVA) with the termined to be slightly more negative than that of rGO (LAA stabilized)
Student–Newman–Keuls test was used for multiple groups comparison. (HA-rGO: −28.6 ± 1.0 mV; rGO: −26.9 ± 0.3 mV). Nevertheless, the
A p-value lower than 0.05 (p < 0.05) was considered statistically surface charge of HA-rGO is in line with that of other HA-based mate-
significant. For data analysis, GraphPad Prism v6.0 (Trial version, rials reported in the literature (He, Zhao, Yin, Tang, & Yin, 2009; Huang
GraphPad Software, CA, USA) was used. et al., 2014; Li et al., 2012). Furthermore, rGO and HA-rGO exhibited a
similar NIR absorption, thus confirming that the functionalized rGO
preserved its photothermal potential (Fig. 2A).
3. Results and discussion The direct functionalization of as-prepared rGO with HA-g-PMAO
led to the production of nanomaterials with suitable physicochemical
3.1. Optimization of the environmentally-friendly reduction of GO properties for application in cancer therapy. In contrast, the functio-
nalization of GO reduced by HH usually requires an additional pur-
The application of reduced forms of GO in cancer therapy is ap- ification step to remove this toxic reductant (Mungse & Khatri, 2014;
pealing since these materials have a higher NIR absorption than GO, Yang et al., 2012). This fact highlights the convenience of our approach.
resulting in an improved photothermal capacity (Robinson et al., 2011).
However, rGO is usually obtained by treating GO with HH, which is a 3.3. Photothermal capacity of HA-rGO
highly toxic hazardous reducing agent (Hussain & Frazier, 2002).
In order to avoid the problems associated with the use of HH, an The temperature variations induced, under NIR laser irradiation, by
environmentally-friendly method was performed to accomplish the HA-rGO were then studied to confirm the photothermal capacity of this
reduction of GO (Fernández-Merino et al., 2010; Huang et al., 2015; Liu nanomaterial (Fig. 2B). In general, HA-rGO produced a time and con-
et al., 2012; Zhang, Yang et al., 2010). For such, LAA was investigated centration-dependent temperature increase when it was irradiated.
in the reduction of GO at two different concentrations (1.5 and 3 mM) After 5 min of irradiation, HA-rGO could induce a temperature increase
and various incubation times at 80 °C (Fig. 1A). Overall, by increasing of about 22 °C at 25 μg/mL (of rGO equivalents) (Fig. 2B). In turn, a
the LAA concentration and the reaction time, a darkening of the GO photoinduced heat of 33 °C was achieved for the highest concentration
solution and an increase on materials’ NIR absorption were attained tested (75 μg/mL of rGO equivalents) (Fig. 2B). Attaining such tem-
(Fig. S1 A and B). Nevertheless, the size distribution of the materials perature increase is of paramount importance since it can induce the
remained un-affected for the conditions tested (Fig. S1 C and D). This is death of cancer cells (Chu & Dupuy, 2014; de Melo-Diogo, Pais-Silva,
of uttermost importance, since the reduction of GO greatly reduces its Costa et al., 2017). Furthermore, the response of water (control) to NIR
solubility and could have increased the nanomaterials’ size through light was meaningless (< 2 °C) (Fig. 2B). Together, these results attest
agglomeration. Therefore, rGO obtained by using 3 mM of LAA and the photothermal efficiency of HA-rGO.
60 min of reaction time was selected for the subsequent assays since it In a previous work, our laboratory developed a hydrothermal
presented a good balance between NIR absorption (mass extinction treatment (80 °C for 24 h) to reduce GO and functionalize it with a PEG
coefficient of 12.67 L/(g.cm), at 808 nm) and preparation time. derivative (de Melo-Diogo, Pais-Silva, Costa et al., 2017). When com-
In comparison to a previous study where LAA was used to reduce pared to the previous protocol, the HA-rGO preparation method is more
GO (Zhang, Yang et al., 2010), we used a lower concentration of LAA (3 convenient since it requires a lower reduction time (1 vs. 24 h) (de
vs. 5.7 mM) and a shorter reaction time (1 vs. 12–48 h). Such results Melo-Diogo, Pais-Silva, Costa et al., 2017). Moreover, HA-rGO also
attest the convenience of the optimized green reduction method. produced a higher photoinduced heat (33 vs. 27 °C, at 75 μg/mL of rGO

95
R. Lima-Sousa et al. Carbohydrate Polymers 200 (2018) 93–99

Fig. 1. Preparation and characterization of HA-rGO. Schematic representation of the reduction and functionalization of rGO with HA-g-PMAO and its application in
cancer photothermal therapy (A). FTIR spectra of rGO, HA-g-PMAO and HA-rGO (B). DLS size distribution of GO, rGO (LAA stabilized) and HA-rGO (C).

equivalents) (de Melo-Diogo, Pais-Silva, Costa et al., 2017), thus dis- (Robinson et al., 2011), while HA-rGO did not affect meaningfully the
playing an enhanced photothermal capacity. viability of these cells up to 75 μg/mL (of rGO equivalents). In another
work, HH-reduced GO functionalized with HA decreased the viability of
3.4. Cytocompatibility of HA-rGO epidermal carcinoma cells (KB cells) to about 80% at 40 μg/mL (of rGO
equivalents) (Miao et al., 2013). Therefore, the environmentally-
Before assessing the phototherapeutic capacity of HA-rGO, first the friendly approach used to reduce GO produces materials with improved
cytocompatibility of this nanomaterial when not irradiated was in- cytocompatibility upon functionalization with HA-g-PMAO, that may be
vestigated in MCF-7 cells (breast cancer cell model) and NHDF (healthy able to perform a spatial-temporal controlled therapy.
cell model). HA-rGO did not induce meaningful effects on the viability
of both cell lines, in the tested concentration range (1–75 μg/mL of rGO 3.5. Targeting capacity of HA-rGO
equivalents) and incubation times (24 or 48 h) (Fig. 3). In contrast, rGO
(non-functionalized) was showed to induce some cytotoxicity to both To confirm the targeting capacity of HA-rGO, the internalization of
cells (Fig. S8). In this way, the functionalization of rGO with HA-g- this nanostructure by MCF-7 cells (CD44 overexpressing cell line)
PMAO improved its cytocompatibility, which is fundamental for the (Calcagno et al., 2010; Olsson et al., 2011) and NHDF (low CD44 ex-
application of HA-rGO in cancer on-demand therapy. pression) (Miyatake et al., 2013) was analyzed by CLSM. Prior to this
HA-rGO displayed an enhanced cytocompatibility in comparison to assay, HA-rGO was labelled with Rhod B for its cellular uptake be vi-
other functionalized rGO materials attained using HH reported in the sualized (Zhang, Xia et al., 2010). The Rhod B labelling was confirmed
literature (Miao et al., 2013; Robinson et al., 2011). For instance, PE- by analyzing the UV–vis spectrum of Rhod B labelled HA-rGO (Fig. 4A
Gylated rGO displayed an IC50 towards MCF-7 cells of about 80 μg/mL and S9A), which displays an absorption peak at 575 nm that is

Fig. 2. Characterization of the photothermal


capacity of HA-rGO. UV-Vis absorption spectra
of GO (25 μg/mL), rGO (LAA stabilized; 25 μg/
mL of rGO equivalents), and HA-rGO (25 μg/
mL of rGO equivalents) (A). Temperature var-
iation curves of HA-rGO at different con-
centrations (of rGO equivalents) during 5 min
of NIR irradiation (808 nm, 1.7 W/cm2) (B).

96
R. Lima-Sousa et al. Carbohydrate Polymers 200 (2018) 93–99

Fig. 3. Evaluation of the cytocompatibility profile of HA-rGO. Cytocompatibility of HA-rGO towards MCF-7 cells (A) and NHDF (B) at different concentrations (of
rGO equivalents) for 24 and 48 h. Data represent mean ± SD, n = 5. K- and K + represent negative and positive controls, respectively.

attributed to the fluorescent probe (Zhang, Xia et al., 2010). Moreover, effect, the fluorescence signals emitted by Rhod B labelled HA-rGO
DLS analysis confirmed that the Rhod B labelling did not affect the size should still enable its use in CLSM studies (Zhong et al., 2016).
distribution of the nanomaterials (Fig. 4B). Taking into account that The uptake of Rhod B labelled HA-rGO by MCF-7 cells and NHDF
rGO-based materials have a high fluorescence quenching capacity, we was then studied by CLSM (Fig. 4C). The results revealed that the in-
then studied the fluorescence emitted by Rhod B labelled HA-rGO (Fig. ternalization profile of the nanostructures by MCF-7 cells and NHDF
S9B). Compared to free Rhod B, the Rhod B labelled HA-rGO emitted was different (Fig. 4C). Fluorescence signals with a higher intensity
fluorescence with a lower intensity (Fig. S9B). Despite of the quenching were observed in the cytoplasm of MCF-7 cells incubated with Rhod B

Fig. 4. Targeting capacity of HA-rGO. UV-Vis absorption spectra of Rhod B labelled HA-rGO (A). DLS size distribution of HA-rGO and Rhod B labelled HA-rGO (B).
Representative CLSM images of Rhod B labelled HA-rGO and free Rhod B uptake by MCF-7 and NHDF cells (λex = 514 nm; λem = 513–703 nm) (C).

97
R. Lima-Sousa et al. Carbohydrate Polymers 200 (2018) 93–99

Fig. 5. Evaluation of the photothermal capa-


city of HA-rGO. Schematic representation of
the PTT mediated by HA-rGO (A). Effect of HA-
rGO in MCF-7 cells without NIR (w/o NIR) and
with NIR (w/ NIR) laser irradiation (808 nm,
1.7 W/cm2, 5 min) (B). K- w/o NIR represents
the negative control. K- w/ NIR represents cells
solely treated with NIR light. Data represents
mean ± SD, n = 5 (*p < 0.01; **p <
0.0001), n.s.= non significant.

labelled HA-rGO when compared to those found on NHDF (Fig. 4C). As PTT.
expected, the internalization of free Rhod B appeared to be similar on
both types of cells (Fig. 4C). In other reports, nanomaterials functio- Declarations of interest
nalized with HA also demonstrated a higher uptake by MCF-7 cells (Gao
et al., 2017; Wang et al., 2016). Taken together these results suggest None.
that HA-rGO has a preferential internalization by MCF-7 cells, thus
confirming the nanostructures’ CD44 targeting ability. Acknowledgements

This work was supported by FEDER funds through the POCI –


3.6. Phototherapeutic effect mediated by HA-rGO
COMPETE 2020 – Operational Programme Competitiveness and
Internationalisation in Axis I – Strengthening research, technological
After assessing the HA-rGO cytocompatibility and preferential up-
development and innovation (Project POCI-01-0145-FEDER-007491)
take by MCF-7 cells, the phototherapeutic effect mediated by this na-
and National Funds by FCT – Foundation for Science and Technology
nomaterial was investigated (Fig. 5A). HA-rGO produced, under NIR
(Project UID/Multi/00709/2013). The funding from CENTRO-01-0145-
laser irradiation, a dose-dependent reduction on MCF-7 cells’ viability
FEDER-028989 is also acknowledged. Duarte de Melo-Diogo and
(Fig. 5B). At the lowest concentration tested (25 μg/mL of rGO
Elisabete C. Costa acknowledge individual PhD fellowships from FCT
equivalents), HA-rGO did not induce meaningful cytotoxicity towards
(SFRH/BD/103506/2014 and SFRH/BD/103507/2014).
cancer cells (viability of ≈ 96%). In stark contrast, the photothermal
effect mediated by HA-rGO at 75 μg/mL (of rGO equivalents) induced a
Appendix A. Supplementary data
reduction of cancer cells’ viability to ≈ 6%. Furthermore, the use of
NIR light alone did not induce any toxicity on cancer cells (Fig. 5B).
Supplementary material related to this article can be found, in the
Additionally, non-irradiated HA-rGO did not induce any deleterious
online version, at doi:https://doi.org/10.1016/j.carbpol.2018.07.066.
effect on cancer cells (Fig. 5B). Taken together, these results confirm
that HA-rGO can produce an on-demand therapeutic effect upon NIR
References
laser irradiation.
In another report, rGO produced using dopamine as a reducing
Akhavan, O., Ghaderi, E., Aghayee, S., Fereydoonia, Y., & Talebia, A. (2012). The use of a
agent and functionalized with the antiarrhythmic peptide 10 (AAP10) glucose-reduced graphene oxide suspension for photothermal cancer therapy. Journal
produced a phototherapeutic effect towards MCF-7 cells similar to that of Materials Chemistry, 22(27), 13773–13781. https://doi.org/10.1039/
of HA-rGO, although it required a higher power density (2–2.2 vs. C2JM31396K.
Akhavan, O., Ghaderi, E., & Emamy, H. (2012). Nontoxic concentrations of PEGylated
1.7 W/cm2) and a higher dose (120 vs. 75 μg/mL) (Yu et al., 2017). In graphene nanoribbons for selective cancer cell imaging and photothermal therapy.
other work, the PTT mediated by arginine-glycine-aspartic acid (RGD)- Journal of Materials Chemistry, 22(38), 20626–20633. https://doi.org/10.1039/
functionalized PEGylated rGO nanoribbons induced cancer cells abla- C2JM34330D.
Calcagno, A. M., Salcido, C. D., Gillet, J. P., Wu, C. P., Fostel, J. M., Mumau, M. D., et al.
tion at a low dose (10 μg/mL) but demanded a higher intensity (7.5 W/
(2010). Prolonged drug selection of breast cancer cells and enrichment of cancer stem
cm2) and a longer period of irradiation (8 min) (Akhavan, Ghaderi, & cell characteristics. Journal of the National Cancer Institute, 102(21), 1637–1652.
Emamy, 2012). In this way, HA-rGO is a promising nanomaterial for https://doi.org/10.1093/jnci/djq361.
Chen, D., Feng, H., & Li, J. (2012). Graphene oxide: Preparation, functionalization, and
CD44 targeted cancer PTT.
electrochemical applications. Chemical Reviews, 112(11), 6027–6053. https://doi.
org/10.1021/cr300115g.
Chu, K. F., & Dupuy, D. E. (2014). Thermal ablation of tumours: Biological mechanisms
4. Conclusion and advances in therapy. Nature Reviews Cancer, 14(3), 199–208. https://doi.org/10.
1038/nrc3672.
The two-major hypotheses of this research work were confirmed. Chung, C., Kim, Y. K., Shin, D., Ryoo, S.-R., Hong, B. H., & Min, D.-H. (2013). Biomedical
applications of graphene and graphene oxide. Accounts of Chemical Research, 46(10),
The environmentally-friendly reduction of GO was optimized con- 2211–2224. https://doi.org/10.1021/ar300159f.
sidering nanomaterials’ NIR absorption and size distribution. On the de Melo-Diogo, D., Pais-Silva, C., Costa, E. C., Louro, R. O., & Correia, I. J. (2017). D-
other hand, the HA-g-PMAO was successfully synthesized (confirmed by alpha-tocopheryl polyethylene glycol 1000 succinate functionalized nanographene
oxide for cancer therapy. Nanomedicine (London), 12(5), 443–456. https://doi.org/
FTIR, 1H NMR and TGA analysis) and explored in the coating of rGO. 10.2217/nnm-2016-0384.
The functionalization of rGO with HA-g-PMAO improved its stability de Melo-Diogo, D., Pais-Silva, C., Dias, D. R., Moreira, A. F., & Correia, I. J. (2017).
and cytocompatibility. Moreover, HA-rGO appeared to have a pre- Strategies to improve cancer Photothermal therapy mediated by nanomaterials.
Advanced Healthcare Materials, 6(10), https://doi.org/10.1002/adhm.201700073.
ferential internalization by CD44 overexpressing cells, which indicates
De Silva, K. K. H., Huang, H.-H., Joshi, R. K., & Yoshimura, M. (2017). Chemical reduction
the targeting capacity of this nanoformulation. Furthermore, the on- of graphene oxide using green reductants. Carbon, 119, 190–199. https://doi.org/10.
demand PTT mediated by HA-rGO induced cancer cells’ ablation. 1016/j.carbon.2017.04.025.
Overall, HA-rGO is a promising nanomaterial for CD44 targeted cancer Du, J., Cullen, J. J., & Buettner, G. R. (2012). Ascorbic acid: Chemistry, biology and the

98
R. Lima-Sousa et al. Carbohydrate Polymers 200 (2018) 93–99

treatment of cancer. Biochimica et Biophysica Acta, 1826(2), 443–457. https://doi. phospholipid monolayer membrane functionalized graphene for drug delivery.
org/10.1016/j.bbcan.2012.06.003. Journal of Materials Chemistry, 22(38), 20634–20640. https://doi.org/10.1039/
Fang, J., Nakamura, H., & Maeda, H. (2011). The EPR effect: Unique features of tumor C2JM34494G.
blood vessels for drug delivery, factors involved, and limitations and augmentation of Liu, S., Zeng, T. H., Hofmann, M., Burcombe, E., Wei, J., Jiang, R., et al. (2011).
the effect. Advanced Drug Delivery Reviews, 63(3), 136–151. https://doi.org/10.1016/ Antibacterial activity of graphite, graphite oxide, graphene oxide, and reduced gra-
j.addr.2010.04.009. phene oxide: Membrane and oxidative stress. ACS Nano, 5(9), 6971–6980. https://
Fernández-Merino, M. J., Guardia, L., Paredes, J. I., Villar-Rodil, S., Solís-Fernández, P., doi.org/10.1021/nn202451x.
Martínez-Alonso, A., et al. (2010). Vitamin C is an ideal substitute for hydrazine in Maeda, H. (2012). Macromolecular therapeutics in cancer treatment: The EPR effect and
the reduction of graphene oxide suspensions. The Journal of Physical Chemistry C, beyond. Journal of Controlled Release, 164(2), 138–144. https://doi.org/10.1016/j.
114(14), 6426–6432. https://doi.org/10.1021/jp100603h. jconrel.2012.04.038.
Frazier, N., & Ghandehari, H. (2015). Hyperthermia approaches for enhanced delivery of Miao, W., Shim, G., Kang, C. M., Lee, S., Choe, Y. S., Choi, H. G., et al. (2013). Cholesteryl
nanomedicines to solid tumors. Biotechnology and Bioengineering, 112(10), hyaluronic acid-coated, reduced graphene oxide nanosheets for anti-cancer drug
1967–1983. https://doi.org/10.1002/bit.25653. delivery. Biomaterials, 34(37), 9638–9647. https://doi.org/10.1016/j.biomaterials.
Gao, X., Zhang, J., Xu, Q., Huang, Z., Wang, Y., & Shen, Q. (2017). Hyaluronic acid-coated 2013.08.058.
cationic nanostructured lipid carriers for oral vincristine sulfate delivery. Drug Miyatake, Y., Oliveira, A. L., Jarboui, M. A., Ota, S., Tomaru, U., Teshima, T., et al.
Development and Industrial Pharmacy, 43(4), 661–667. https://doi.org/10.1080/ (2013). Protective roles of epithelial cells in the survival of adult T-Cell Leukemia/
03639045.2016.1275671. Lymphoma cells. The American Journal of Pathology, 182(5), 1832–1842. https://doi.
Gaspar, V. M., Baril, P., Costa, E. C., de Melo-Diogo, D., Foucher, F., Queiroz, J. A., et al. org/10.1016/j.ajpath.2013.01.015.
(2015). Bioreducible poly(2-ethyl-2-oxazoline)-PLA-PEI-SS triblock copolymer mi- Mungse, H. P., & Khatri, O. P. (2014). Chemically functionalized reduced graphene oxide
celles for co-delivery of DNA minicircles and Doxorubicin. Journal of Controlled as a novel material for reduction of friction and wear. The Journal of Physical
Release: Official Journal of the Controlled Release Society, 213, 175–191. https://doi. Chemistry C, 118(26), 14394–14402. https://doi.org/10.1021/jp5033614.
org/10.1016/j.jconrel.2015.07.011. Olsson, E., Honeth, G., Bendahl, P. O., Saal, L. H., Gruvberger-Saal, S., Ringnér, M., et al.
Hashemi, M., Omidi, M., Muralidharan, B., Smyth, H., Mohagheghi, M. A., Mohammadi, (2011). CD44 isoforms are heterogeneously expressed in breast cancer and correlate
J., et al. (2017). Evaluation of the photothermal properties of a reduced graphene with tumor subtypes and cancer stem cell markers. BMC Cancer, 11, 418. https://doi.
Oxide/Arginine nanostructure for near-infrared absorption. ACS Applied Materials & org/10.1186/1471-2407-11-418.
Interfaces, 9(38), 32607–32620. https://doi.org/10.1021/acsami.7b11291. Robinson, J. T., Tabakman, S. M., Liang, Y., Wang, H., Casalongue, J. T., Vinh, D., et al.
He, M., Zhao, Z., Yin, L., Tang, C., & Yin, C. (2009). Hyaluronic acid coated poly(butyl (2011). Ultrasmall reduced graphene oxide with high near-infrared absorbance for
cyanoacrylate) nanoparticles as anticancer drug carriers. International Journal of photothermal therapy. Journal of the American Chemical Society, 133(17), 6825–6831.
Pharmaceutics, 373(1), 165–173. https://doi.org/10.1016/j.ijpharm.2009.02.012. https://doi.org/10.1021/ja2010175.
Hobbs, S. K., Monsky, W. L., Yuan, F., Roberts, W. G., Griffith, L., Torchilin, V. P., et al. Shi, S., Chen, F., Ehlerding, E. B., & Cai, W. (2014). Surface engineering of graphene-
(1998). Regulation of transport pathways in tumor vessels: Role of tumor type and based nanomaterials for biomedical applications. Bioconjugate Chemistry, 25(9),
microenvironment. Proceedings of the National Academy of Sciences, 95(8), 1609–1619. https://doi.org/10.1021/bc500332c.
4607–4612. https://doi.org/10.1073/pnas.95.8.4607. Strangman, G., Boas, D. A., & Sutton, J. P. (2018). Non-invasive neuroimaging using near-
Hu, Y., Sun, D., Ding, J., Chen, L., & Chen, X. (2016). Decorated reduced graphene oxide infrared light. Biological Psychiatry, 52(7), 679–693. https://doi.org/10.1016/S0006-
for photo-chemotherapy. Journal of Materials Chemistry B, 4(5), 929–937. https://doi. 3223(02)01550-0.
org/10.1039/C5TB02359A. Wang, J., Ma, W., Guo, Q., Li, Y., Hu, Z., Zhu, Z., et al. (2016). The effect of dual-func-
Huang, J., Zhang, H., Yu, Y., Chen, Y., Wang, D., Zhang, G., et al. (2014). Biodegradable tional hyaluronic acid-vitamin E succinate micelles on targeting delivery of doxor-
self-assembled nanoparticles of poly (d,l-lactide-co-glycolide)/hyaluronic acid block ubicin. International Journal of Nanomedicine, 11, 5851–5870. https://doi.org/10.
copolymers for target delivery of docetaxel to breast cancer. Biomaterials, 35(1), 2147/ijn.s113882.
550–566. https://doi.org/10.1016/j.biomaterials.2013.09.089. Yang, C., He, Y., Zhang, H., Liu, Y., Wang, W., Du, Y., et al. (2015). Selective killing of
Huang, P., Wang, S., Wang, X., Shen, G., Lin, J., Wang, Z., et al. (2015). Surface func- breast cancer cells expressing activated CD44 using CD44 ligand-coated nanoparticles
tionalization of chemically reduced graphene oxide for targeted photodynamic in vitro and in vivo. Oncotarget, 6(17), 15283–15296. https://doi.org/10.18632/
therapy. Journal of Biomedical Nanotechnology, 11(1), 117–125. https://doi.org/10. oncotarget.3681.
1166/jbn.2015.2055. Yang, K., Wan, J., Zhang, S., Tian, B., Zhang, Y., & Liu, Z. (2012). The influence of surface
Hussain, S. M., & Frazier, J. M. (2002). Cellular toxicity of hydrazine in primary rat chemistry and size of nanoscale graphene oxide on photothermal therapy of cancer
hepatocytes. Toxicological Sciences, 69(2), 424–432. https://doi.org/10.1093/toxsci/ using ultra-low laser power. Biomaterials, 33(7), 2206–2214. https://doi.org/10.
69.2.424. 1016/j.biomaterials.2011.11.064.
Li, J., Huo, M., Wang, J., Zhou, J., Mohammad, J. M., Zhang, Y., et al. (2012). Redox- Yu, J., Lin, Y. H., Yang, L., Huang, C. C., Chen, L., Wangu, W. C., et al. (2017). Improved
sensitive micelles self-assembled from amphiphilic hyaluronic acid-deoxycholic acid anticancer photothermal therapy using the bystander effect enhanced by antiar-
conjugates for targeted intracellular delivery of paclitaxel. Biomaterials, 33(7), rhythmic peptide conjugated dopamine‐modified reduced graphene oxide nano-
2310–2320. https://doi.org/10.1016/j.biomaterials.2011.11.022. composite. Advanced Healthcare Materials, 6(2), 1600804. https://doi.org/10.1002/
Li, J. L., Tang, B., Yuan, B., Sun, L., & Wang, X.-G. (2013). A review of optical imaging and adhm.201600804.
therapy using nanosized graphene and graphene oxide. Biomaterials, 34(37), Zhang, L., Xia, J., Zhao, Q., Liu, L., & Zhang, Z. (2010). Functional graphene oxide as a
9519–9534. https://doi.org/10.1016/j.biomaterials.2013.08.066. nanocarrier for controlled loading and targeted delivery of mixed anticancer drugs.
Li, Y., Feng, L., Shi, X., Wang, X., Yang, Y., Yang, K., et al. (2014). Surface coat- Small, 6(4), 537–544. https://doi.org/10.1002/smll.200901680.
ing‐dependent cytotoxicity and degradation of graphene derivatives: Towards the Zhang, J., Yang, H., Shen, G., Cheng, P., Zhang, J., & Guo, S. (2010). Reduction of gra-
design of non‐toxic, degradable nano‐graphene. Small, 10(8), 1544–1554. https:// phene oxide vial-ascorbic acid. Chemical Communications, 46(7), 1112–1114. https://
doi.org/10.1002/smll.201303234. doi.org/10.1039/B917705A.
Liu, J., Cui, L., & Losic, D. (2013). Graphene and graphene oxide as new nanocarriers for Zhong, Y., Goltsche, K., Cheng, L., Xie, F., Meng, F., Deng, C., et al. (2016). Hyaluronic
drug delivery applications. Acta Biomaterialia, 9(12), 9243–9257. https://doi.org/10. acid-shelled acid-activatable paclitaxel prodrug micelles effectively target and treat
1016/j.actbio.2013.08.016. CD44-overexpressing human breast tumor xenografts in vivo. Biomaterials, 84,
Liu, J., Guo, S., Han, L., Wang, T., Hong, W., Liu, Y., et al. (2012). Synthesis of 250–261. https://doi.org/10.1016/j.biomaterials.2016.01.049.

99

You might also like