Вы находитесь на странице: 1из 7

Materials Science and Engineering C 68 (2016) 798–804

Contents lists available at ScienceDirect

Materials Science and Engineering C

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

One-step synthesis of soy protein/graphene nanocomposites and their


application in photothermal therapy
Xuejiao Jiang, Zhao Li, Jinrong Yao, Zhengzhong Shao, Xin Chen ⁎
State Key Laboratory of Molecular Engineering of Polymers, Collaborative Innovation Center of Polymers and Polymer Composite Materials, Department of Macromolecular Science, Laboratory of
Advanced Materials, Fudan University, Shanghai 200433, People's Republic of China

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

Article history: Photothermal therapy, due to its security and effectiveness, has recently become a promising cancer treatment
Received 12 April 2016 after surgery, radiotherapy, chemotherapy, and biological therapy. In this article, soy protein isolate/reduced
Received in revised form 4 July 2016 graphene oxide (SPI/rGO) nanocomposites are prepared via a simple and green process. That is, GO is reduced
Accepted 16 July 2016
in situ and stabilized by SPI, an abundant, low-cost, sustainable natural material, and simultaneously forms
Available online 18 July 2016
SPI/rGO nanocomposites. The resulting SPI/rGO nanocomposites disperse in water very well and exhibit good
Keywords:
biocompatibility due to the attachment of SPI to the surface of rGO. Such SPI/rGO nanocomposites demonstrate
Plant proteins an excellent photothermal capacity and are able to kill HeLa cells efficiently with near-infrared irradiation
Carbon materials (808 nm). The results in this work suggest that such a SPI/rGO hybrid material could be a promising candidate
Hybrid materials for future applications of photothermal therapy.
Photothermal transformation © 2016 Elsevier B.V. All rights reserved.
Biocompatibility

1. Introduction polymer composites. They found that graphene had the potential to
convert light to heat efficiently and that the conversion was even better
In the past quarter century, although there were many explorations than carbon nanotubes when graphene was used as a nanoscale heater
in cancer biology, such studies have not been translated into even and energy-transfer unit for infrared (IR)-triggered actuators. The
remotely comparable advances in clinical practice [1,2]. Current cancer perfect sp2 carbon-network structure of graphene ensures that it
treatments still primarily focus on traditional surgical intervention, efficiently absorbs and transforms IR light into thermal energy [22].
radiation and chemotherapeutic drugs, which often damage healthy However, the potential of graphene in bionanotechnology is somehow
cells and tissues, or even cause toxicity in patients. However, a new underestimated due to its disadvantages, such as highly hydrophobic
and non-invasive modality called photothermal therapy (PTT) has and featureless surfaces, low aqueous solubility, and poor biocompatibil-
been increasingly recognized as a promising alternative to traditional ity [23]. To improve these drawbacks, several natural biomacromolecules
cancer therapies, such as surgery, radiotherapy, and chemotherapy, (DNAs and proteins) have been applied to modify the surface of graphene
due to its spatiotemporal selectivity and minimal invasiveness [3,4]. nanosheets [24–26]. In our previous research, we demonstrated that silk
PTT usually utilizes photo-absorbing agents, including graphene oxide fibroin was able to grow on the surface of graphene nanosheets in the
(GO) [5,6], carbon nanotubes [7,8], gold nanoparticles [9–12], and or- form of nanofibrils by self-assembly. The graphene/silk nanofibril hybrid
ganic near-infrared (NIR) dyes [6,13] (such as cyanine) to generate a preserved the high electrical conductivity of pure graphene. In addition,
high localized temperature for ablating tumors. This advantage allows the graphene/silk nanofibril hybrid film prepared via a vacuum-assisted
clinicians to save most of the healthy tissues close to the targeted lesions filtering process showed good flexibility and biocompatibility [23,27].
by using near infrared (NIR) light irradiation [13,14]. Soy protein isolate (SPI), which is one of the most abundant natural
Graphene, which is one atom thick and has a two-dimensional (2D) polymers extracted from soybeans, has attracted comprehensive inter-
single carbon layer, is emerging as a rising star in material science and ests in the field of material science [28] due to its sustainability, low
condensed matter physics [15] due to its high specific surface area and cost, functionality, biocompatibility, and tunable degradability. There
excellent mechanical, electrical, and thermal properties [16–18]. These have been several soy protein-based materials, such as plastics [29],
properties enable graphene to be considered an ideal material for a gels [30], films [31,32], and biomedical materials [33] reported in the
variety of applications. Recently, Yang [19], Kuilla [20], Meng [21] and literature.
their coworkers have explored the photothermal effect of graphene in The synthesis of SPI/inorganic composites in previous research has
mainly focused on improving the mechanical performance of SPI mate-
⁎ Corresponding author. rials by using inorganic materials as strengthening agents. For example,
E-mail address: chenx@fudan.edu.cn (X. Chen). Lee and his coworkers prepared SPI/montmorillonite (SPI/MMT)

http://dx.doi.org/10.1016/j.msec.2016.07.034
0928-4931/© 2016 Elsevier B.V. All rights reserved.
X. Jiang et al. / Materials Science and Engineering C 68 (2016) 798–804 799

composite film with a melt extrusion method for drug release [34]. How- with a GO concentration of 0.1 wt% and solid mass ratios of 10:1, 50:1,
ever, the preparation conditions were relatively harsh, and the structure 100:1 for SPI/GO. These samples were named SG10, SG50, and SG100,
of the obtained composite was not homogeneous at the nanoscale. respectively, for both the SPI/GO mixtures and final SPI/rGO nanocom-
Zhuang and his coworkers have reported the preparation of SPI/ posites. Subsequently, the SPI/GO mixtures were incubated at 70 °C
graphene aerogel with a hydrothermal method, and the obtained com- for 24 h to obtain SPI/rGO nanocomposites.
posite was used as an adsorbent to remove antibiotics (ciprofloxacin TEM images of the synthesized nanocomposites were observed with
and tetracycline) from aqueous solutions [35,36]. They found the SPI/ an FEI Tecnai G2 20 TWIN transmission electron microscope at 200 kV.
graphene composites showed better hydrophilicity, improved biologi- The samples were prepared by drying the diluted nanocomposite
cal compatibility, and lower cytotoxicity compared with pure graphene. solutions on a formvar/carbon-coated copper grid. AFM images were ac-
During their preparation, they need add ascorbic acid as a reductant, quired using a Multi-mode 8 atomic force microscope in tapping mode.
and the application of such a SPI/graphene composite was for water X-ray diffraction (XRD) data were recorded on an X'pert Pro X-ray
treatment. diffractometer with Cu Kα radiation. Raman spectra were obtained
SPI is rich with sulfhydryl groups, and it also contains reducing with a Renishaw inVia Reflex spectrometer coupled to a Lieca micro-
amino acid residues such as tyrosine. Therefore, SPI may be able to scope at a wavelength of 785 nm for a He-Ne laser. The energy of the
reduce GO to prepare graphene, which we call reduced GO, i.e., rGO. laser was 0.6 mW, which is only 1/10 that used for a normal laser,
In addition, SPI is an amphiphilic polymer (a nature for all proteins), because the graphene would be carbonized under high energy.
which makes it well known for its adhesiveness to solid surfaces [37].
There are already several reports indicating that other proteins, such 2.3. Measurement of free sulfhydryl with the Ellman method [40]
as bovine serum albumin (BSA) and silk fibroin, are able to attach to
the surface of graphene by hydrophobic interaction and π-π stacking Ellman's reagent (10 mM DTNB) was prepared by suspending
[24]. Thus, SPI may have the potential to reduce GO in situ and then 198.2 mg DTNB in 50 mL of 50 mmol/L Na2HPO4 solution, and all
attach to the surface of the resulting rGO nanosheets to ultimately DTNB was dissolved. Aliquots of all reagents were immediately frozen
form a hybrid nanocomposite. at −25 °C, and thawed portions were used within 1 day. Next, 100 μL
In this article, we show evidence that we were able to obtain a SPI/ 1 wt% of SG10 and SPI samples was diluted with Tris-HCl buffer
rGO nanocomposite via a facile, mild, and one-step protocol. SPI was (pH = 8) to a final volume of 1 mL. After adding 100 μL of 10 mmol/L
used as both the reducing and stabilizing agent, and it improved the Ellman reagent, the samples were immediately vortexed, and the A412
biocompatibility of the resulting SPI/rGO nanocomposite. Then we values were read against the samples using 100 μL Tris-HCl buffer
demonstrated that a SPI/rGO nanocomposite could be a promising instead of Ellman's regent as a blank after a 10-min incubation time
photothermal agent for cancer therapy. (1 cm light path).
The amount of free sulfhydryl in the SPI sample was calculated as
2. Materials and methods follows:

2.1. Materials ½SH ðμmol=g soy proteinÞ ¼ 1; 000; 000  ½ASPI −½ASG10 =ðε  b  cÞ

SPI powder (protein content N90%) was obtained from Shenyuan where the standard value ε412 = 14,150 L/mol·cm, b = 1 cm, and c
Food Co. Ltd., Shanghai, China. Dithiothreitol (99%) and 5,5′-Dithiobis- refers to the concentration of SPI (g/L)
(2-nitrobenzoic acid) (DTNB) were purchased from Aladdin, Shanghai,
China. Graphite powder, concentrated H2SO4 (98%), KMnO4, NaOH 2.4. Cytotoxicity tests
and guanidine hydrochloride (99%) were purchased from Sinopharm
Chemical Reagent Co., Ltd., Shanghai, China. All chemicals used in this The cytotoxicity of GO and SPI/rGO nanocomposites was determined
work were analytical grade and used without further purification. by a CCK-8 assay on L929 cells. L929 cells were plated in a 96-well plate
The preparation of an SPI aqueous solution followed a well- at a seeding density of 10,000 cells/well with 100 μL of DMEM media at
established procedure, as reported in our previous work [38]. Briefly, various concentrations of GO and SPI/rGO. After incubating for 24 h at
SPI powder was dissolved in aqueous solution containing 6 mol/L guani- 37 °C, the solutions were removed from the well plates immediately
dine hydrochloride and 1 mmol/L dithiothreitol. After dialysis against prior to the addition of a colorimetric indicator. PBS was used to wash
NaOH aqueous solution (pH = 11.5) for two days and deionized the residual nanocomposites to prevent any interference in the absor-
water for another day at room temperature, the SPI solution was obtain- bance readings.
ed. The resulting SPI solution was transferred into the freezer compart-
ment of a refrigerator and then freeze-dried to obtain pure SPI powder. 2.5. In vitro photothermal performance and photothermal therapy (PTT)
GO was synthesized using a modified Hummers method [39]. Brief- effects on HeLa cells
ly, 3.7 g of graphite power were mixed into 120 mL concentrated
sulphuric acid (98.5%) that was pre-cooled by an ice bath. Next, 2.5 g The SPI/rGO nanocomposite solution samples were diluted to a
NaNO3 and 11.6 g KMnO4 were slowly added under continuous vigor- desired concentration of 100 μg/mL with distilled water. Then, 1.2 mL
ous stirring such that the temperature was constantly maintained SPI/rGO nanocomposite solution samples (i.e., SG10, SG50 and
below 20 °C. The mixture was then stirred continuously at 35 °C for SG100), pure GO solution (100 μg/mL), pure SPI solution (0.1 wt%),
7 h. After 9.5 g KMnO4 was added, the reaction was left running uninter- and deionized water were continuously irradiated by an NIR laser
rupted at 35 °C for 14 h. The reaction was quenched by slowly pouring (808 nm, 2.5 W) at a distance of 5 cm for 5 min in 1.5 mL vials. The
the mixture into 600 mL H2O under stirring. Finally, 20 mL of 30% photostability of the SPI/rGO nanocomposite samples was assessed by
H2O2 was added, and the mixture turned bright yellow and bubbled. switching the laser on/off five times. Temperature was measured by
The mixture was filtered and washed with 1 L 5% HCl aqueous solution an infrared thermal camera. All experiments were conducted at room
and 200 mL ethanol. The solid powder was dried at 70 °C. temperature (~20 °C).
The PTT effect of the SPI/rGO nanocomposite was determined by a
2.2. Preparation and characterization of SPI/rGO nanocomposites CCK-8 assay on HeLa cells. Digested HeLa cells were mixed with a
desired amount of pure GO, SG10, SG50 and SG100 solution in 1.5 mL
Desired quantities of SPI powder were added to the GO solutions vials. The concentrations of graphene in each vial were set as 40, 60,
under vigorous stirring to prepare a final SPI/GO mixture (10 mL) 80, and 100 μg/mL, respectively. After 1 h of incubation, these HeLa
800 X. Jiang et al. / Materials Science and Engineering C 68 (2016) 798–804

However, after adding SPI to the GO solution and incubating it for a


period of time, the surface of the GO nanosheet became rough. The
more SPI that was added, the rougher the GO surface was (Fig. 2b–d).
This phenomenon clearly indicates that globular soy proteins were
assembled on the surface of the GO nanosheets.
In contrast, AFM could be used not only to observe the surface mor-
phology of GO nanosheet but also to estimate the average thickness of
the GO nanosheets before and after the addition of SPI. Fig. 3a shows
that the thickness of pure GO was b1 nm, which was consistent with
the height of mono-layer GO. However, the thickness of GO after adding
Fig. 1. Digital photos of solutions containing (a) pure GO, (b) SG10, (c) SG50, (d) SG100. SPI (for example, SG50 in Fig. 3b) significantly increased to 10–15 nm,
which further confirmed the attachment of soy protein on the surface
of the GO nanosheet.
cell solutions were then irradiated for 3 min with a 2.5 W 808 nm diode Although we confirmed that soy proteins were successfully attached
laser. Immediately after heating, HeLa cells were plated in a 96-well to the surface of the GO nanosheet, we still did not know whether the
plate with a cell density of 10,000 cells/well. After 24 h of incubation soy protein changed the structure of GO. Therefore, X-ray diffraction
at 37 °C, the cell viability was determined by a CCK-8 assay using the (XRD) and Raman spectroscopy were used to characterize the structural
same protocol as the cytotoxicity test shown above. details of the nanocomposite. The XRD pattern of pure soy protein
showed a main crystalline peak at approximately 20° and a small peak
3. Results and discussion at approximately 10° (Fig. 4a, curve a), which correlates with those
reported in the literature [41,42]. In the meantime, pure GO showed a
3.1. Preparation and characterization of SPI/rGO nanocomposites strong peak at 10.97° (curve b), which corresponded to the typical
diffraction peak of the (001) plane in the GO nanosheet [43]. However,
To prepare the SPI/rGO nanocomposites we designed, we first added the typical diffraction peak of GO at 10.97° disappeared after the
a certain amount of SPI powder to 0.1 wt% GO solution and then incu- addition of SPI, and the nanocomposite only showed a broad peak at
bated the solution at 70 °C under continuous stirring for 24 h. We approximately 20° (curve c–e) that was similar to pure SPI. Such a result
found the resulting solution was very stable, which was almost the demonstrated that the GO nanosheets were exfoliated and provided
same as pure GO solution (Fig. 1). In addition, these solutions were additional evidence for the successful separation of graphene in the
able to be stored under 4 °C for several months without any aggregation. form of a mono-layer [20,44] after the addition of SPI.
Although these solutions have the same appearance, the TEM im- Meanwhile, the Raman spectra shown in Fig. 4b suggest that GO
ages of the surface morphology of the GO nanosheets were different was reduced by SPI. The pure GO and nanocomposites showed strong
(Fig. 2). Fig. 2a shows that the surface of the GO nanosheet was smooth. twin peaks at 1308 and 1583 cm−1. It is well known that the peak

Fig. 2. TEM images of (a) pure GO, (b) SG10, (c) SG50, and (d) SG100.
X. Jiang et al. / Materials Science and Engineering C 68 (2016) 798–804 801

group in tyrosine has an electron donating ability that is sufficiently


strong to reduce the noble metal to nanoparticles [49] and GO to rGO
[37]. Therefore, there is no doubt that tyrosine in SPI was also able to
reduce GO in the current system. On the other hand, when we prepared
an SPI aqueous solution, we used dithiothreitol to destroy the disulfide
bond generate abundant free sulfhydryl groups. Monahan and his
coworkers [50] reported that the oxidization of sulfhydryl occurred
when the temperature increased to 70 °C or above in neutral solution,
which demonstrated that the sulfhydryl groups had reductive proper-
ties. Therefore, we have reason to believe that the sulfhydryl groups in
SPI also contributed to the reduction process in our study.
To prove whether the sulfhydryl groups in SPI also served as a reduc-
ing agent to reduce GO, the Ellman method was used to measure the
amount of free sulfhydryl in SPI and SG10. Fig. 5 shows that the SPI
solution had a sulfhydryl characteristic peak at 412 nm, which almost
disappeared in the SG10 sample and implies that there was a significant
decrease in free sulfhydryl after forming the nanocomposite of SPI and
GO. We were able to calculate that the reaction amount of free sulfhy-
dryl in our pure SPI sample was 36.7 μmol/g soy protein, which is
reasonable because the theoretical sulfhydryl amount was 90 μmol/g
soy protein, according to the cysteine percentage in the 7S and 11S
components of SPI [34]. These results confirmed that there were free
sulfhydryl groups in our pure SPI sample and that the reducing sulfhy-
Fig. 3. AFM images and height profiles for (a) pure GO and (b) SG50. dryl groups were consumed after they reacted with GO and reduced
GO to rGO. These characterizations show that we successfully prepared
at 1308 cm− 1 is assigned to the D band (associated with the order/ a SPI/rGO nanocomposite.
disorder degree of the system), which is attributed to the breathing
mode of the k-point phonons of A1g symmetry. The peak at 1583 cm−1
is assigned to the G band (an indicator of the stacking structure), 3.2. Cytotoxicity tests
which results from the E2g phonon of sp2 carbon atoms [45]. Thus, the
D/G intensity ratio is defined as a parameter for measuring the sp3 do- Biocompatibility is the one of the most important aspects when de-
main percentage of a carbon structure containing both sp3 and sp2 veloping a biomaterial. We know that SPI has good biocompatibility
bonds [46] and the degree of exfoliation/disorder in graphene material [49], but GO remains somewhat cytotoxicity [51] and rGO even has
[45]. Previous studies indicate that when GO is chemically reduced to more cytotoxicity [52]. Therefore, we used SG10 as an example to inves-
rGO, the surface defects and the carbon bond fractures are increased, tigate the cytotoxicity because it contains less SPI than SG50 and SG100.
and the topological disorder is introduced in graphene structure, Fig. 6 shows the cytotoxicity of GO and SG10 on L929 cells by measuring
resulting in the increase of D/G intensity ratio. The higher the D/G inten- their relative cell viabilities after 24 h of incubation through a CCK-8
sity ratio is observed, the more the GO is reduced to rGO [47,48]. From assay. We can see that at low graphene concentrations, neither pure
Fig. 4b, we were able to calculate the D/G intensity ratios of pure GO, GO nor SG10 show obvious cytotoxicity, and there was no significant
SG10, SG50, SG100 as 1.29, 1.83, 2.03, and 2.30, respectively. It is clear difference between them. However, with an increase in the graphene
that the D/G intensity ratio of the nanocomposite was larger than that concentration to 50 μg/mL, significant differences appeared in the cell
of pure GO and increased as the amount of added SPI increased. There- viability between pure GO and SG10. When the graphene concentration
fore, we are confident that GO has been partially reduced by SPI, and increased to 250 μg/mL, the cell viability for pure GO dropped to b 80%
the more SPI was added, the more GO was reduced. Additionally, the and exhibited some cytotoxicity. In contrast, the cell viabilities of SG10
Raman intensity of SPI in our experimental conditions was negligible remained very high for all graphene concentrations that were tested,
compared to that of the graphene materials, so it had little effect on the even higher than 100%, because the soy protein itself may be considered
D/G intensity ratio. an additional nutrilite for L929 cells. Therefore, the cytotoxicity result
SPI contains reducing amino acid residues, such as tyrosine and cys- implies that the SPI/rGO nanocomposite is more biocompatible than
teine [34]. Previous studies have already confirmed that the phenolic pure GO.

Fig. 4. XRD patterns (a) and Raman spectra (b) of pure SPI, pure GO, and the
corresponding nanocomposites. (I) pure SPI; (II) pure GO; (III) SG10; (IV) SG 50; (V)
SG100. Fig. 5. UV–vis spectra of (a) pure SPI and (b) SG10.
802 X. Jiang et al. / Materials Science and Engineering C 68 (2016) 798–804

Fig. 8. Measurement of sample temperature changes versus time during 808 nm NIR
Fig. 6. Cytotoxicity of pure GO and SG10 in L929 cells after incubation for 24 h. Data were irradiation (2.5 W, 8 min). The concentration of graphene was 100 μg/mL.
expressed as mean ± standard deviations (SD). Statistical analysis was performed by one-
way analysis of variance (ANOVA), * denotes p b 0.05.

could be attributed to the successful reduction of GO by SPI. It is clear


3.3. In vitro photothermal heating characterization and photothermal that as more SPI was added, more GO was reduced. On the other
therapy effects on HeLa cells hand, more SPI aggregates in the nanocomposite solution may also
have scattered more NIR light, which probably also contributed to the
To verify the potential of the SPI/rGO nanocomposite as a photothermal effect [54].
photothermal agent, the photothermal effect of the composites was There are several reports regarding the photothermal transforma-
evaluated by measuring the temperature change of the solutions after tion properties of graphene-based materials in the literature. However,
they were exposed to an 808 nm laser (2.5 W, 5 min). Fig. 7 shows it is not easy to compare their properties directly with ours because the
the visual photothermal images of the vials containing pure GO and experimental conditions that carried out in different research groups
different SPI/rGO nanocomposites. The image of the SPI/rGO nanocom- were varied. The volume of the solution to be irradiated, the concentra-
posite was brighter than that of pure GO, and sample SG100 was the tion of photothermal agent in the solution, the irradiation time, the
brightest. Such a phenomenon is easy to be understood because it is power and even the wavelength of the laser may be different [55–60].
well known that the photothermal performance of GO is much worse Generally, if we compare the increase of the temperature after 5 min
than rGO [53]. The temperature increase curves are also shown in of irradiation, such a value was from 3 to 42 °C for these graphene-
Fig. 8. With the increase in irradiation time, the temperature increased based materials reported. Among them, the most similar system to
faster with an increasing amount of SPI added when the nanocompos- ours is BSA/rGO, but its temperature increase was 30 °C [55] or 40 °C
ites formed. The temperature for the SG10, SG50, and SG100 solutions [56], which is lower than our SPI/rGO system. In addition, we shall
reached approximately 40, 70, and 80 °C, respectively, from room emphasis that we did not introduce any reducing agent, like hydrazine
temperature at approximately 20 °C. In contrast, under the same laser hydrate in our preparation process, which may greatly improve the
irradiation conditions, the temperature curve for 100 μg/mL of pure biocompatibility and environmental friendliness in our SPI/rGO
SPI solution was similar to that of pure water without any obvious nanocomposites.
temperature changes. In the meantime, pure GO showed a weak We also tested the photothermal stability of SPI/rGO nanocompos-
photothermal effect with a temperature increase b 10 °C. Therefore, ites by switching the laser on/off for five cycles. The temperature
the significant photothermal effect of the SPI/rGO nanocomposite increased when the laser was on and decreased to room temperature
when the laser was turned off. The increase/decrease curve of the
temperature showed a very good repeatability at least for five cycles
(Figures not shown). This outcome indicates that the SPI/rGO nanocom-
posite presented an excellent photothermal capability and stability
under repeated laser irradiation, which makes it a promising candidate
for a PTT agent.
Finally, we evaluated the feasibility of using the SPI/rGO nanocom-
posite as a photothermal agent for cancer therapy. The in vitro PTT effect
of the SPI/rGO nanocomposite was studied in HeLa cells under 808 nm
laser irradiation. Fig. 9 shows all SPI/rGO nanocomposites killed the
HeLa cells more efficiently than pure GO under NIR irradiation. For ex-
ample, at the lowest graphene concentration (40 μg/mL), pure GO
only killed b20% of the HeLa cells, but the worst SPI/rGO composite
(sample SG10) killed N50% of the cells, and the best composite (sample
SG100) killed almost 80% of the HeLa cells. With an increase in graphene
concentration, the difference between pure GO and SPI/rGO becomes
more significant. When the graphene concentration increased to
80 μg/mL, all SPI/rGO nanocomposites killed N95% of the HeLa cells.
This outcome coincides with the photothermal heating characterization
results. Interestingly, the SPI/rGO nanocomposites also exhibited an SPI
dependent cytotoxicity to the HeLa cells that correlated with the
photothermal test results. At relatively low graphene concentrations
Fig. 7. NIR imaging of pure GO (a), SG10 (b), SG50 (c) and SG100 (d) solution
temperatures after exposure to an 808 nm laser. Each image shows the surface
(40 and 60 μg/mL), the cell viability showed a significant difference
temperature of the vial. The temperature scale bar is between 20 and 80 °C. The among SG10, SG50, and SG100, that is SG100 can kill most HeLa cells,
concentration of graphene was 100 μg/mL. whereas SG10 showed the worst cell toxicity. It is possible that with
X. Jiang et al. / Materials Science and Engineering C 68 (2016) 798–804 803

[6] K. Yang, J. Wan, S. Zhang, B. Tian, Y. Zhang, Z. Liu, The influence of surface chemistry
and size of nanoscale graphene oxide on photothermal therapy of cancer using
ultra-low laser power, Biomaterials 33 (2012) 2206–2214.
[7] A. Burlaka, S. Lukin, S. Prylutska, O. Remeniak, Y. Prylutskyy, M. Shuba, S.
Maksimenko, U. Ritter, P. Scharff, Hyperthermic effect of multi-walled carbon
nanotubes stimulated with near infrared irradiation for anticancer therapy: in
vitro studies, Exp. Oncol. 32 (2010) 48–50.
[8] Y. Hashida, H. Tanaka, S. Zhou, S. Kawakami, F. Yamashita, T. Murakami, T.
Umeyama, H. Imahori, M. Hashida, Photothermal ablation of tumor cells using a
single-walled carbon nanotube–peptide composite, J. Control. Release 173 (2014)
59–66.
[9] J. Kim, S. Park, J.E. Lee, S.M. Jin, J.H. Lee, I.S. Lee, I. Yang, J.-S. Kim, S.K. Kim, M.-H. Cho,
T. Hyeon, Designed fabrication of multifunctional magnetic gold nanoshells and
their application to magnetic resonance imaging and photothermal therapy,
Angew. Chem. Int. Ed. 45 (2006) 7754–7758.
[10] A.Y. Lin, J.K. Young, A.V. Nixon, R.A. Drezek, Encapsulated Fe3O4/Ag complexed cores
in hollow gold nanoshells for enhanced theranostic magnetic resonance imaging
and photothermal therapy, Small 10 (2014) 3246–3251.
Fig. 9. Photothermal therapy effects on HeLa cells with pure GO, SG10, SG50, and SG100. [11] Y. Ma, X. Liang, S. Tong, G. Bao, Q. Ren, Z. Dai, Gold nanoshell nanomicelles for poten-
Data are expressed as mean ± standard deviation (SD). Statistical analysis was tial magnetic resonance imaging, light-triggered drug release, and photothermal
performed with one-way analysis of variance (ANOVA), * denotes p b 0.05, ** denotes therapy, Adv. Funct. Mater. 23 (2013) 815–822.
p b 0.01, *** denotes p b 0.001. [12] I.H. El-Sayed, X. Huang, M.A. El-Sayed, Selective laser photo-thermal therapy of
epithelial carcinoma using anti-EGFR antibody conjugated gold nanoparticles,
Cancer Lett. 239 (2006) 129–135.
the increase in graphene concentrations, even SG10 can generate [13] M. Guo, J. Huang, Y. Deng, H. Shen, Y. Ma, M. Zhang, A. Zhu, Y. Li, H. Hui, Y. Wang, X.
enough heat to kill HeLa cells. As such, there was no obvious difference Yang, Z. Zhang, H. Chen, pH-responsive cyanine-grafted graphene oxide for fluores-
cence resonance energy transfer-enhanced photothermal therapy, Adv. Funct.
among SG10, SG50, and SG100. In summary, SG 100 showed the best Mater. 25 (2015) 59–67.
PPT effect in HeLa cells with an effective concentration of 60 μg/mL, [14] S. Wang, K. Li, Y. Chen, H. Chen, M. Ma, J. Feng, Q. Zhao, J. Shi, Biocompatible
but SG10 and SG50 needed a higher effective concentration of 80 μg/mL. PEGylated MoS2 nanosheets: controllable bottom-up synthesis and highly efficient
photothermal regression of tumor, Biomaterials 39 (2015) 206–217.
It should be noted that the PTT still has some drawbacks. The main [15] A.K. Geim, K.S. Novoselov, The rise of graphene, Nat. Mater. 6 (2007) 183–191.
possible side effect of photothermal agents in PTT is that, while the [16] G. Tang, Z.G. Jiang, X. Li, H.B. Zhang, Z.Z. Yu, Simultaneous functionalization and
high temperature kills cancer cells, it increases the risk of damaging reduction of graphene oxide with polyetheramine and its electrically conductive
epoxy nanocomposites, Chin. J. Polym. Sci. 32 (2014) 975–985.
the normal tissues near the lesions through heat conduction, which
[17] H.Y. Mao, S. Laurent, W. Chen, O. Akhavan, M. Imani, A.A. Ashkarran, M. Mahmoudi,
may affect the therapeutic accuracy [61]. Graphene: promises, facts, opportunities, and challenges in nanomedicine, Chem.
Rev. 113 (2013) 3407–3424.
[18] R. Raccichini, A. Varzi, S. Passerini, B. Scrosati, The role of graphene for electrochem-
4. Conclusions ical energy storage, Nat. Mater. 14 (2015) 271–279.
[19] K. Yang, S. Zhang, G. Zhang, X. Sun, S.-T. Lee, Z. Liu, Graphene in mice: ultrahigh in
In this study, we successfully prepared SPI/rGO nanocomposites vivo tumor uptake and efficient photothermal therapy, Nano Lett. 10 (2010)
3318–3323.
through an easy, green, and one-step approach. We demonstrated that
[20] T. Kuilla, S. Bhadra, D. Yao, N.H. Kim, S. Bose, J.H. Lee, Recent advances in graphene
SPI molecules attached to the surface of the GO nanosheets and reduced based polymer composites, Prog. Polym. Sci. 35 (2010) 1350–1375.
GO to rGO in situ with various characterizations, such as SEM, AFM, [21] D. Meng, S. Yang, L. Guo, G. Li, J. Ge, Y. Huang, C.W. Bielawski, J. Geng, The enhanced
photothermal effect of graphene/conjugated polymer composites: photoinduced
XRD, and Raman spectroscopy. We found that the SPI/rGO nanocom-
energy transfer and applications in photocontrolled switches, Chem. Commun. 50
posites showed good biocompatibility when graphene content changed (2014) 14345–14348.
from 2.5 to 500 μg/mL. The photothermal tests indicated that the [22] J. Liang, Y. Xu, Y. Huang, L. Zhang, Y. Wang, Y. Ma, F. Li, T. Guo, Y. Chen, Infrared-
temperature of the SPI/rGO nanocomposite solutions was able to triggered actuators from graphene-based nanocomposites, J. Phys. Chem. C 113
(2009) 9921–9927.
increase as high as 80 °C after 5 min of irradiation with 808 nm NIR [23] S.J. Ling, C. Li, J. Adamcik, S.H. Wang, Z.Z. Shao, X. Chen, R. Mezzenga, Directed
laser. The more SPI added when forming the nanocomposite, the higher growth of silk nanofibrils on graphene and their hybrid nanocomposites, ACS
the temperature it can be reached. Finally, the PTT tests on HeLa cells Macro Lett. 3 (2014) 146–152.
[24] C. Li, J. Adamcik, R. Mezzenga, Biodegradable nanocomposites of amyloid fibrils and
indicated that these SPI/rGO nanocomposites exhibited much better graphene with shape-memory and enzyme-sensing properties, Nat. Nanotechnol. 7
photothermal effects than that of pure GO. We believe that the biocom- (2012) 421–427.
patible SPI/rGO nanocomposite we generated for this study may have [25] T.H. Han, W.J. Lee, D.H. Lee, J.E. Kim, E.Y. Choi, S.O. Kim, Peptide/graphene hybrid
assembly into core/shell nanowires, Adv. Mater. 22 (2010) 2060–2064.
great potential as a photothermal agent for cancer treatment. [26] S. Hatamie, O. Akhavan, S.K. Sadrnezhaad, M.M. Ahadian, M.M. Shirolkar, H.Q. Wang,
Curcumin-reduced graphene oxide sheets and their effects on human breast cancer
Acknowledgments cells, Mater. Sci. Eng. C 55 (2015) 482–489.
[27] Y. Wang, Y. Song, Y. Wang, X. Chen, Y. Xia, Z. Shao, Graphene/silk fibroin based car-
bon nanocomposites for high performance supercapacitors, J. Mater. Chem. A 3
This work is supported by the National Natural Science Foundation (2015) 773–781.
of China (No. 21274028, 21574023 and 21574024). The authors [28] F. Song, D.L. Tang, X.L. Wang, Y.Z. Wang, Biodegradable soy protein isolate-based
materials: a review, Biomacromolecules 12 (2011) 3369–3380.
sincerely thank Dr. Yuhong Yang, Dr. Shengjie Ling, Dr. Suhang Wang, [29] H. Tian, W. Wu, G. Guo, B. Gaolun, Q. Jia, A. Xiang, Microstructure and properties of
Mr. Yingxin Liu, and Mr. Cheng Zhang for their valuable suggestions glycerol plasticized soy protein plastics containing castor oil, J. Food Eng. 109 (2012)
and discussions. 496–500.
[30] J. Jiang, Y.L. Xiong, Extreme pH treatments enhance the structure-reinforcement role
of soy protein isolate and its emulsions in pork myofibrillar protein gels in the
References presence of microbial transglutaminase, Meat Sci. 93 (2013) 469–476.
[31] H. Bai, J. Xu, P. Liao, X. Liu, Mechanical and water barrier properties of soy protein
[1] M. Ferrari, Cancer nanotechnology: opportunities and challenges, Nat. Rev. Cancer 5 isolate film incorporated with gelatin, J. Plast. Film Sheeting 29 (2013) 174–188.
(2005) 161–171. [32] L. Ma, Y.H. Yang, J.R. Yao, Z.Z. Shao, X. Chen, Robust soy protein films obtained by
[2] D. Peer, J.M. Karp, S. Hong, O.C. Farokhzad, R. Margalit, R. Langer, Nanocarriers as an slight chemical modification of polypeptide chains, Polym. Chem. 4 (2013)
emerging platform for cancer therapy, Nat. Nanotechnol. 2 (2007) 751–760. 5425–5431.
[3] D.K. Chatterjee, P. Diagaradjane, S. Krishnan, Nanoparticle-mediated hyperthermia [33] K. Chien, E. Chung, R. Shah, Investigation of soy protein hydrogels for biomedical ap-
in cancer therapy, Ther. Deliv. 2 (2011) 1001–1014. plications: materials characterization, drug release, and biocompatibility, J.
[4] L. Cheng, K. Yang, Q. Chen, Z. Liu, Organic stealth nanoparticles for highly effective in Biomater. Appl. 28 (2014) 1085–1096.
vivo near-infrared photothermal therapy of cancer, ACS Nano 6 (2012) 5605–5613. [34] J. Lee, K. Kim, Characteristics of soy protein isolate-montmorillonite composite films,
[5] K. Yang, L. Hu, X. Ma, S. Ye, L. Cheng, X. Shi, C. Li, Y. Li, Z. Liu, Multimodal imaging J. Appl. Polym. Sci. 118 (2010) 2257–2263.
guided photothermal therapy using functionalized graphene nanosheets anchored [35] Y. Zhuang, F. Yu, J. Ma, J. Chen, Adsorption of ciprofloxacin onto graphene–soy
with magnetic nanoparticles, Adv. Mater. 24 (2012) 1868–1872. protein biocomposites, New J. Chem. 39 (2015) 3333–3336.
804 X. Jiang et al. / Materials Science and Engineering C 68 (2016) 798–804

[36] Y. Zhuang, F. Yu, J. Ma, J. Chen, Facile synthesis of three-dimensional graphene–soy [50] F.J. Monahan, J.B. German, J.E. Kinsella, Effect of pH and temperature on protein
protein aerogel composites for tetracycline adsorption, Desalin. Water Treat. 57 unfolding and thiol/disulfide interchange reactions during heat-induced gelation
(2015) 1–10. of whey proteins, J. Agric. Food Chem. 43 (1995) 46–52.
[37] J. Liu, S. Fu, B. Yuan, Y. Li, Z. Deng, Toward a universal “adhesive nanosheet” [51] S. Gurunathan, J. Han, E. Kim, J. Park, J. Kim, Reduction of graphene oxide by
for the assembly of multiple nanoparticles based on a protein-induced reduction/ resveratrol: a novel and simple biological method for the synthesis of an effective
decoration of graphene oxide, J. Am. Chem. Soc. 132 (2010) 7279–7281. anticancer nanotherapeutic molecule, Int. J. Nanomedicine 10 (2015) 2951–2969.
[38] K. Tian, Z.Z. Shao, X. Chen, Natural electroactive hydrogel from soy protein isolation, [52] B. Zhang, W. Yang, G. Zhai, Biomedical applications of the graphene-based materials,
Biomacromolecules 11 (2010) 3638–3643. Mater. Sci. Eng. C 61 (2016) 953–964.
[39] D.C. Marcano, D.V. Kosynkin, J.M. Berlin, A. Sinitskii, Z. Sun, A. Slesarev, L.B. Alemany, [53] J.T. Robinson, S.M. Tabakman, Y.Y. Liang, H.L. Wang, H.S. Casalongue, D. Vinh, H.J.
W. Lu, J.M. Tour, Improved synthesis of graphene oxide, ACS Nano 4 (2010) Dai, Ultrasmall reduced graphene oxide with high near-infrared absorbance for
4806–4814. photothermal therapy, J. Am. Chem. Soc. 133 (2011) 6825–6831.
[40] C.K. Riener, G. Kada, H.J. Gruber, Quick measurement of protein sulfhydryls with [54] V.M.P. Ruiz-Henestrosa, M.J. Martinez, J.M.R. Patino, A.M.R. Pilosof, A dynamic light
Ellman's reagent and with 4, 4′-dithiodipyridine, Anal. Bioanal. Chem. 373 (2002) scattering study on the complex assembly of glycinin soy globulin in aqueous
266–276. solutions, J. Am. Oil Chem. Soc. 89 (2012) 1183–1191.
[41] V.D. Gupta, A quantitative X-ray study of alpha-beta transformation in keratin, [55] Z. Sheng, L. Song, J. Zheng, D. Hu, M. He, M. Zheng, G. Gao, P. Gong, P. Zhang, Y. Ma,
Nature 181 (1958) 113. Protein-assisted fabrication of nano-reduced graphene oxide for combined in vivo
[42] X.Y. Zhao, H.T. Zhu, B.W. Zhang, J. Chen, Q. Ao, X.Y. Wang, XRD, SEM, and XPS photoacoustic imaging and photothermal therapy, Biomaterials 34 (2013)
analysis of soybean protein powders obtained through extraction involving reverse 5236–5243.
micelles, J. Am. Oil Chem. Soc. 92 (2015) 975–983. [56] Y.A. Cheon, J.H. Bae, B.G. Chung, Reduced graphene oxide nanosheet for chemo-
[43] T. Blanton, D. Majumdar, Characterization of X-ray irradiated graphene oxide photothermal therapy, Langmuir 32 (2016) 2731–2736.
coatings using X-ray diffraction, X-ray photoelectron spectroscopy, and atomic [57] X. Shi, H. Gong, Y. Li, C. Wang, L. Cheng, Z. Liu, Graphene-based magnetic plasmonic
force microscopy, Powder Diffract. 28 (2013) 68–71. nanocomposite for dual bioimaging and photothermal therapy, Biomaterials 34
[44] H. Feng, Y. Li, J. Li, Strong reduced graphene oxide–polymer composites: hydrogels (2013) 4786–4793.
and wires, RSC Adv. 17 (2012) 6988–6993. [58] M.-C. Wu, A.R. Deokar, J.-H. Liao, P.-Y. Shih, Y.-C. Ling, Graphene-based
[45] D.R. Dreyer, S. Park, C.W. Bielawski, R.S. Ruoff, The chemistry of graphene oxide, photothermal agent for rapid and effective killing of bacteria, ACS Nano 7 (2013)
Chem. Soc. Rev. 39 (2010) 228–240. 1281–1290.
[46] O. Akhavan, E. Ghaderi, S. Aghayee, Y. Fereydooni, A. Talebi, The use of a glucose-re- [59] S.M. Sharker, J.E. Lee, S.H. Kim, J.H. Jeong, I. In, H. Lee, S.Y. Park, pH triggered in vivo
duced graphene oxide suspension for photothermal cancer therapy, J. Mater. Chem. photothermal therapy and fluorescence nanoplatform of cancer based on respon-
22 (2012) 13773–13781. sive polymer-indocyanine green integrated reduced graphene oxide, Biomaterials
[47] M.M. Storm, M. Overgaard, R. Younesi, N.E.A. Reeler, T. Vosch, U.G. Nielson, K. 61 (2015) 229–238.
Edstrom, P. Norby, Reduced graphene oxide for Li–air batteries: The effect of [60] X. Li, Y. Zhang, Y. Wu, Y. Duan, X. Luan, Q. Zhang, Q. An, Combined photothermal and
oxidation time and reduction conditions for graphene oxide, Carbon 85 (2015) surface-enhanced Raman spectroscopy effect from spiky noble metal nanoparticles
233–244. wrapped within graphene-polymer layers: using layer-by-layer modified reduced
[48] P.G. Ren, D.X. Yan, X. Ji, T. Chen, Z.M. Li, Temperature dependence of graphene oxide graphene oxide as reactive precursors, ACS Appl. Mater. Interfaces 7 (2015)
reduced by hydrazine hydrate, Nanotechnology 22 (2011) 55705–55712. 19353–19361.
[49] S. Zhao, J.R. Yao, X. Fei, Z.Z. Shao, X. Chen, An antimicrobial film by embedding in situ [61] X. Zhu, W. Feng, J. Chang, Y.-W. Tan, J. Li, M. Chen, Y. Sun, F. Li, Temperature-
synthesized silver nanoparticles in soy protein isolate, Mater. Lett. 95 (2013) feedback upconversion nanocomposite for accurate photothermal therapy at facile
142–144. temperature, Nat. Commun. 7 (2016) 10437.

Вам также может понравиться