You are on page 1of 7

Colloids and Surfaces B: Biointerfaces 128 (2015) 515–521

Contents lists available at ScienceDirect

Colloids and Surfaces B: Biointerfaces


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

Systemic modulation of the stability of pluronic hydrogel by a small


amount of graphene oxide
Da-Ae Won, Manse Kim, Giyoong Tae ∗
School of Materials Science and Engineering, Gwangju Institute of Science and Technology, 123 Cheomdan-gwagiro, Oryong-dong, Buk-gu, Gwangju
500-712, Republic of Korea

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

Article history: Thermo-sensitive and injectable hydrogels have been widely investigated for drug delivery, tissue engi-
Received 22 August 2014 neering, and other biomedical applications. Pluronic copolymers can form thermo-sensitive physical gel
Received in revised form 26 February 2015 state, thus applicable for injectable hydrogels. However, they are not stable in vivo, showing a very fast
Accepted 2 March 2015
dissolution, which limits their applications. We propose a novel Pluronic-based physical hydrogel with
Available online 7 March 2015
enhanced stability by simply adding a small quantity of graphene oxide (GO) which has a large surface
area and can make strong interactions with Pluronic. Further carboxylated GO could act as a more efficient
Keywords:
additive. The addition of GO increased the moduli of hydrogels, but more importantly, it enhanced the
Pluronic
Nanogel
stability of Pluronic gel dramatically. The in vitro dissolution rate of Pluronic hydrogel could be system-
Graphene oxide atically modulated by increasing GO content. Upon subcutaneous injection at a sol state, GO-containing
Degradation hydrogel induced a stable gel state, and was maintained over several weeks whereas very fast degra-
Release dation was observed without the addition of GO. Furthermore, histological analyses demonstrated that
Injectable the GO-containing Pluronic hydrogel was biocompatible and showed no severe inflammatory response.
Similarly, GO-containing hydrogel resulting from the packing of Pluronic-based nanogel also showed
the more enhanced stability by the addition of GO both in vitro and in vivo. In both systems, hydrogels
with remarkably enhanced stability by the addition of GO were also effective for the sustained release
of loaded protein, and the release rates were mainly determined by the degradation rates of hydrogels.
Thus, these GO-containing Pluronic systems can be used as a thermo-sensitive injectable system with a
sufficient stability in vivo.
© 2015 Elsevier B.V. All rights reserved.

1. Introduction Pluronic, biocompatible triblock copolymers composed of


hydrophobic poly(propylene oxide) and hydrophilic poly(ethylene
Hydrogels have many advantages such as biocompatibility, oxide), are a good example of thermo-sensitive polymers as
ability for absorbing water and low inflammatory responses an injectable hydrogel systems [10]. Pluronic can self-assemble
[1], applicable as a reservoir for slow elution of drugs or pro- into micelles in aqueous solutions above critical micelle con-
teins [2]. Particularly, injectable hydrogels are more preferred for centration (CMC) and critical micelle temperature (CMT), and
drug/protein delivery and tissue engineering. Injectable hydrogels exhibit temperature-responsive sol–gel transition behaviors [5].
can be made by using responses to environmental stimuli such as Above CMT and at high concentration of Pluronic, gelation
temperature [3–6] or pH [7]. For example, these materials are in a occurs due to micelle packing, resulting in a dramatic change in
flow state before administration, but once injected, they rapidly their rheological properties [11,12]. Using this thermo-reversible
become a gel under physiological conditions. Such systems can gelation, Pluronic F127 (PF127) has been studied widely, espe-
provide several advantages including delivery of large amount in cially as an injectable system for local drug delivery [13].
a minimally invasive manner, gelation at the desired tissue sites, However, the application of PF127 hydrogel has been limited
and minimized scar formation, thus reducing the risk of infection because of a very fast dissolution, thus fast drug release as
[6,8]. In addition, bioactive molecules or cells can be incorporated well as weak mechanical properties in vivo. Therefore, various
by simple mixing before injection [9]. reports have attempted to enhance the stability of PF127 gel
in vivo. Chen et al. showed a thermo-responsive nanocompos-
ite hydrogel composed of hexamethylene diisocyanate, PF127
∗ Corresponding author. Tel.: +82 62 715 2305; fax: +82 62 715 2304. and hyaluronic acid [6]. Our group also reported an injectable
E-mail address: gytae@gist.ac.kr (G. Tae). and photo-polymerizable Pluronic based hydrogel by using a

http://dx.doi.org/10.1016/j.colsurfb.2015.03.002
0927-7765/© 2015 Elsevier B.V. All rights reserved.
516 D.-A. Won et al. / Colloids and Surfaces B: Biointerfaces 128 (2015) 515–521

sufficiently long induction period during photo-polymerization sodium phosphate dibasic, potassium chloride, sodium chloride,
[14]. and sodium azide were purchased from Sigma-Aldrich (St. Louis,
Recently, there has been a remarkably increased attention MO, USA). Anhydrous diethyl ether was purchased from Fisher
to biological applications of graphene and graphene oxide (GO). Scientific Inc. (Pittsburgh, PA, USA). 4-(2-Hydroxyethoxy) phenyl-
Graphene is a single layer of sp2 -hybridized carbon atoms in a 2-d (2-hydroxyl-2-propyl) ketone (Irgaure 2959) was obtained from
crystalline lattice [15]. GO is an oxidized derivative of graphene to Ciba Specialty Chemicals Inc. (Basel, Switzerland). Single layer
make it water dispersible. The planar structure and ultra-high sur- graphene oxide (X, Y dimension < 500 nm) was obtained from
face area of graphene and GO make them suitable for loading a large Angstrom Materials Inc. (Dayton, OH, USA) as a 0.5% (w/v) solu-
number of substances including metal, drugs, biomolecules, and tion. Chloroacetic acid and sodium hydroxide were purchased from
even cells [16–18]. Graphene is very hydrophobic, thus graphene Aldrich (Milwaukee, WI, USA). Dialysis membrane [MWCO 50,000
can interact with hydrophobic molecules based on ␲–␲ stacking & 3500] was the product of Spectrum Laboratories, Inc. (Houston,
and van der waals interaction, whereas GO is negatively charged TX, USA). 0.2 ␮m cellulose sterilization syringe filters and 0.8 ␮m
with carboxyl and hydroxyl groups, so hydrogen bonding and cellulose sterilization syringe filters were purchased from Toyo
charge interaction as well as hydrophobic association can con- Roshi Kaisha. Ltd. (Tokyo, Japan). 0.2 ␮m nylon syringe filter was
tribute to the interaction with other molecules. Therefore, GO can purchased from Whatman (Florham Park, NJ, USA). Human VEGF
associate with variety of molecules including amphiphilic poly- ELISA Development Kit (900-K10) was purchased from Peprotech
mers. Based on the non-covalent interaction of GO with variety of (Rocky Hill, NJ, USA).
molecules, GO composite was shown to significantly increase the
mechanical and thermal properties of host materials [18–20]. For
e.g., Zhang et al. added GO as a nano-filler into a polyvinyl alcohol 2.2. Preparation of Pluronic-based nanogel
(PVA) matrix, and showed the increase in tensile strength, break-
ing elongation, and compressive strength [18]. There have also Chemically crosslinked, Pluronic-based nanogel was prepared
been several studies reporting the formation of hydrogel by mix- by simple photo-crosslinking of diacylated Pluronic [3]. First,
ing GO with polymers that can provide sufficient interaction with diacrylated Pluronic F127 (DA-PF 127) was synthesized as previ-
GO [21–25], including PVA [21,22], DNA [23], chitosan [24], and ously reported [3,29,30]. Shortly, dried Pluronic F127 were reacted
Pluronic [25] in certain concentration ranges. Also, GO modified with 10 molar excess of triethyamine and acryloyl chloride in
with cyclodextrin (CD) was used to form complex hydrogel with anhydrous toluene with stirring under argon over 17 h, then it
azo-functionalized polymer [26], and Pluronic-coated GO was used was precipitated in anhydrous diethyl ether in an ice-water bath,
to make a complex hydrogel with CD [27]. In all of these reports, filtered, and dried under vacuum for a few days. The degree of acry-
both GO or modified GO and polymer were necessary for hydrogel lation of Pluronic was over 98%, as determined by comparing the
formation. Without GO, no gelation occurs, thus GO functions as a acryl protons ( CH2 , 5.7–6.4 ppm) and methyl protons in propyl-
main gelator. In contrast, there has been no study reporting that ene oxide units ( CH3 , 1.1 ppm) in 40 MHz 1 H NMR spectroscopy
the addition of a small amount of GO can increase and modulate (D2 O, JNM-ECX-400P, JEOL, Japan). After that, 0.77 wt% of DA-PF
the thermodynamic stability of pre-existing physical gel systems. 127 in de-ionized water (DIW) was photo-crosslinked by adding
GO interacts strongly with Pluronic, mainly by the hydrophobic a photoinitiator Irgacure 2959 in 70% (v/v) ethanol and DIW and
association between PPO part of Pluronic and GO, so the Pluronic UV-irradiation using an unfiltered UV lamp (VL-4.LC, 8 W, Vilber
coating of GO nanosheet (∼40 nm) could increase the colloidal sta- Lourmat, France) for 15 min with 1.3 mW cm−2 intensity. Finally,
bility of GO in physiological media and neutralize the surface charge the unreacted precursors were removed by dialysis (MWCO 50,000)
of carboxylated GO [28]. In addition, the presence of carboxyl group for 1 day.
and hydroxyl group of carboxylated GO can induce the hydrogen
bonding with Pluronic, evidenced by the effect of pH in the inter-
action between carboxylated GO and Pluronic [25], similar to the 2.3. Preparation of carboxylated graphene oxide
interaction between PVA and GO [22].
In this paper, we propose GO as an effective additive to The suspension of carboxylated GO sheet was prepared by
increase and modulate the stability of physical hydrogel systems, chemical treatment and ultrasonication, according to the reference
here hydrogels formed by Pluronic or Pluronic-based nanogel, in [31]. To 10 ml of GO (0.1 w/v%) dispersion in DIW, chloroacetic
addition to its contribution to the mechanical properties of the acid (150 mg) and sodium hydroxide (200 mg) were added. The
hydrogels. We investigate the physical gel system at high concen- solution was stirred at 45 ◦ C for overnight. Then, the solution was
tration of Pluronic (17 wt%) or Pluronic-based nanogel (11 wt%), sonicated by an ultrasonic probe with 750 W, 30% intensity (Vibra-
where they can form hydrogel by themselves by the packing of cell VCX 500, Sonics & Materials Inc., Newtown, CT, USA) for 2 h.
micelle or nanogel. However, since these gels are achieved by The solution was neutralized by dialysis (dialysis bag MWCO 3500)
lyotropic transition, the hydrogels formed by themselves are not against distilled water for a few days. After dialysis, the suspen-
stable but quickly disappear when excess water is added. We report sion was sonicated for 30 min and then filtered through a 0.8 ␮m
here that by adding a very small amount GO to Pluronic gel systems, cellulose acetated filter. The size of carboxylated GO was ∼400 nm,
the stability of physical gels already formed without GO can be analyzed by using an electrophoretic light-scattering spectropho-
greatly enhanced and modulated in an excess water environment tometer (632.8 nm, ELS-8000, Otsuka Electronics Co., Toyko, Japan).
like physiological situation.

2.4. Preparation of the sol state of injectable hydrogels


2. Materials and methods
Two kinds of injectable state hydrogels were prepared by dis-
2.1. Materials solving PF 127 itself, or Pluronic-based nanogel in GO solution at
4 ◦ C for 48 h by using a rotary shaker after vortexing. The final con-
Pluronic F127 (PEO100 PPO65 PEO100, MW 12.6 kDa) was centration of PF 127 was 17 wt%, and that of nanogel was 11 wt%.
donated from BASF Corp. (Seoul, Korea). Acryloyl chloride, tri- Three different concentration of GO was used: 0.04, 0.08, and
ethylamine, anhydrous toluene, potassium phosphate monobasic, 0.1 wt%.
D.-A. Won et al. / Colloids and Surfaces B: Biointerfaces 128 (2015) 515–521 517

2.5. Sol–gel transition phase diagram These skins were fixed in 10% formalin, dehydrated in alcohols, and
embedded in paraffin. The paraffin was sectioned by using a micro-
The sol–gel transition phase diagrams of both bare hydrogels tome at 5 ␮m thickness, and hematoxylin and eosin (H&E) staining
and GO-containing hydrogels were determined by a vial tilting was applied. The photographs of stained sections were taken by
method. Each sample in PBS solution was prepared in 1.5 ml tube, optical microscopy.
and samples were placed in a water-bath. Temperature was started
at 4 ◦ C and gradually increased by 3 ◦ C. The temperature of gel point
3. Results and discussions
was decided when the fluidity was not observed by inverting the
test tube for 1 min. The concentration of GO was set at 0.1 wt%
To increase and modulate the stability of physical gel systems,
(n = 3).
GO was used as an effective additive in this study. Our group pre-
viously reported that the addition of a small amount of Pluronic
2.6. In vitro degradation rates of hydrogels
into GO suspension could induce an insoluble, but rather fragile,
gel-like state in a limited concentration range [25]. In this case, GO
To analyze the stability of various hydrogels, the degradation
acts as a main gelator. No gel-like state was formed without GO
rates of hydrogels in vitro were measured. Hydrogels made of
or below 0.3 wt% of GO. And, the concentration range of Pluronic
Pluronic itself, Pluronic with GO or carboxylated GO, nanogel itself,
was limited to below 1 wt%. Above 1 wt% of Pluronic, the gel-like
or nanogel with carboxylated GO were placed at the bottom of 15 ml
state disappeared. So, both GO and Pluronic were necessary for gel
tubes. The volume of each hydrogel was 200 ␮l. After gelation by
formation, and the excess amount of Pluronic prevented the forma-
increasing temperature to 37 ◦ C, 5 ml of release buffer (PBS with
tion of a gel state. In contrast, in this study, the lyotropic physical
2 mM% sodium azide) was added to each tube. Samples were kept
hydrogels formed by the packing of Pluronic micelle or Pluronic-
in a shaking incubator (SI-600R, Lab. Companion, JEIO TECH, city,
based nanogel at high concentration of Pluronic were the systems
Korea) at 37 ◦ C and 100 rpm. The whole release buffer was replaced
of interest. The addition of a very small amount of GO (up to 0.1 wt%)
with a fresh one daily, and the remaining volumes of hydrogels
could greatly improve and systematically modulate the stability of
were measured at each time points (n = 3).
hydrogels formed by Pluronic micelle (17 wt%) or Pluronic-based
nanogel (11 wt%) in an open environment both in vitro and in vivo.
2.7. In vitro release rates of a growth factor from hydrogels
Thus, the role of GO is an additive, not a gelator, in this study.
The release rates of vascular endothelial growth factor (VEGF)
from hydrogels containing carboxylated GO were characterized. 3.1. In vitro stability of Pluronic hydrogel
100 ng of VEGF was loaded into 200 ␮l of Pluronic (17 wt%) or
Pluronic-based nanogel (11 wt%) with 0.1 wt% of carboxylated GO By the micelle packing, Pluronic solution shows a thermo-
and placed at the bottom of a 15 ml tubes. After gelation by increas- reversible, lyotropic sol–gel phase transition. To characterize the
ing temperature to 37 ◦ C, 5 ml of release buffer (PBS with 2 mM% effect of the addition of GO on the stability of Pluronic hydrogels,
sodium azide and 0.01 wt% BSA) was added to each tube and incu- the degradation rate of Pluronic hydrogel at 17 wt% solution in PBS
bated at 37 ◦ C under shaking at 100 rpm. The whole release buffer in vitro was measured in a sink condition under shaking at 37 ◦ C.
was replaced with a fresh one daily, and the released amounts of The remaining volume of hydrogel in the presence of excess vol-
VEGF from hydrogels were calculated by using an ELISA kit (n = 3). ume of buffer was monitored and the buffer was exchanged with
a fresh one daily. As expected, without GO, Pluronic gel itself was
2.8. Measurement of rheological properties degraded very fast within 1 day in an excess water environment.
In contrast, by adding 0.1 wt% GO, the stability of Pluronic hydrogel
The rheological properties of various hydrogels were measured was greatly improved (Fig. 1a).
by using a rheometer (Gemini, Malvern Instruments, Malvern, Since hydrogen bonding, in addition to hydrophobic associa-
Worcestershire, UK) with parallel plate geometry of 15 mm diam- tion, can contribute to the interaction between Pluronic and GO,
eter and a roughened surface. The gap size was set to 0.45 ␮m and the difference between normal GO and further carboxylated GO
the sample volume was 100 ␮l. The temperature of plate was low- for improving the stability of Pluronic hydrogel (Fig. 1a) was com-
ered to 4 ◦ C, and then the gelation kinetics was measured at 37 ◦ C. pared. Further carboxlyated GO was prepared by reacting GO with
A single frequency at 1 rad/s was used to get gelation kinetics with chloroacetic acid (ClCH2 COOH) in an alkaline condition, which con-
different concentrations of GO, and then a frequency sweep with verts hydroxyl groups (–OH) of GO into carboxyl group (–COOH)
0.1% strain was performed at 37 ◦ C. [31]. After chemical FTIR data (Fig. 1b) shows the increased C O
(1670 cm−1 ) and C O (1100 cm−1 ) bond compared to –OH group
2.9. In vivo animal experiments (3400 cm−1 ) after treatment, which supports the carboxylation of
GO [32]. The difference between carboxylated GO and normal GO
In vivo stability and biocompatibility of various hydrogels were for improving the stability of Pluronic hydrogel was also large; most
studied using male BALB/c mice (8 weeks old). All animals were of gel was degraded in day 5 with normal GO, whereas more than
obtained from Orient Bio Inc. (Seoul, Korea) and were handled 40% of hydrogel was remained with carboxylated GO at the same
in accordance with the guidelines of the Animal Care and Use time. This large difference might result from the more increased
Committee of Gwangju Institute of Science and Technology (GIST). hydrogen bonding that carboxylated GO could provide as well as
Hydrogels were prepared as a sol state by lowering temperature the better dispersion of carboxylated GO in physiological buffer
at 4 ◦ C and injected subcutaneously at the back of mice using a due to an increased negative charge, thus providing more uniform
syringe with a 22G needle. To characterize the in vivo stability distribution in the hydrogel compared to normal GO. Therefore,
of injected hydrogels, the sizes of injected hydrogels were mon- carboxylated GO was used instead of normal GO for all other exper-
itored at determined time points: 1day, 1 week, 2 weeks, and iments.
8 weeks. The stability of GO-containing Pluronic hydrogel (17 wt%) was
For analyzing biocompatibility of the GO-containing hydrogel, analyzed by varying the amount of added GO. Compared to Pluronic
histochemical staining of injected sites was done. Animals were hydrogel itself, the degradation rate was greatly reduced even at
sacrificed and the skin parts containing injected gels were cut out. 0.04 wt% GO. And, by increasing GO concentration, the degradation
518 D.-A. Won et al. / Colloids and Surfaces B: Biointerfaces 128 (2015) 515–521

Fig. 2. Sol–gel transition phase diagrams of (a) Plruonic and GO-containing Pluronic
in PBS, and (b) Pluronic-based nanogel and GO-containing, Pluronic-based nanogel
in PBS (GO: 0.1 wt%, n = 3).

physical Pluronic hydrogel in an open environment, suggesting the


possibility for applying as an injectable hydrogel system in an open
environment like in vivo with a practically useful residence time
at the injection site. Furthermore, the degradation rate of Pluronic
hydrogel could be systemically modulated by the concentration of
GO.
Chemically crosslinked, Pluronic-based nanogel was reported to
present a temperature-sensitive large size variation (∼450 nm at
4 ◦ C and ∼60 nm at 37 ◦ C) while maintaining its structural integrity
[3]. This nanogel solution also becomes a gel state by the packing
of nanogels instead of micelles. The effect of GO on the stability
of hydrogel formed by Pluronic-based nanogel was also investi-
gated. Pluronic-based nanogel (11 wt%) itself was degraded within
7 days at 37 ◦ C, so the gel state formed by the nanogel packing was
more stable than the gel state formed by the micelle packing. By
adding GO, the degradation rate of gel formed by the nanogel pack-
ing was also significantly reduced; time for complete degradation
Fig. 1. (a) Degradation of hydrogel composed of 17 wt% Pluronic with 0.1 wt% car-
was prolonged from 10 days to 15 days, and more than 20 days by
boxylated or unmodified GO. (b) FT-IR spectra of carboxylated GO and unmodified increasing the GO concentration from 0.04, 0.08 to 0.1 wt% (Fig. 2d).
GO. Degradation of (c) hydrogel composed of 17 wt% Pluronic and carboxylated GO The choice of different concentration of Pluronic nanogel (11 wt%)
with various concentrations of GO and (d) hydrogel composed of 11 wt% Pluronic- from Pluronic micelle solution (17 wt%) to measure the dissolution
based nanogel and carboxylated GO with various concentrations of GO. PBS with
rate of the hydrogels was based on the different sol–gel transition
2 mM sodium azide in a shaking incubator at 37 ◦ C and 100 rpm was used (n = 3).
boundary between two systems; in both cases, the concentrations
of making gel state were slight higher than the sol–gel transition
rate was significantly reduced further systematically; the degrada- boundary to form a gel state at 37 ◦ C.
tion rate with 0.08 wt% GO was similar to that of 0.1 wt% GO until
day 3, but after 4 days, a noticeable difference started to be observed 3.2. Sol–gel transition phase diagram
(Fig. 1c). Therefore, time for complete degradation was prolonged
from 5 days to 10 days, and more than 15 days by increasing the con- The effect of the addition of GO (0.1 wt%) on the sol–gel transi-
centration from 0.04, 0.08 to 0.1 wt%, respectively, in contrast to the tion phase boundary of Pluronic or Pluronic-based nanogel in PBS
complete dissolution of Pluronic hydrogel itself within 1 day. These was analyzed by using a vial tilting method with 3 ◦ C interval of
results indicate that the addition of a small amount of GO com- temperature increase. In the case of Pluronic, the phase bound-
pared to Pluronic (17 wt%) could greatly improve the stability of ary of sol-to-gel transition by increasing temperature at a fixed
D.-A. Won et al. / Colloids and Surfaces B: Biointerfaces 128 (2015) 515–521 519

concentration of Pluronic did not show significant change by adding


GO. No significant change of the lower phase boundary of Pluronic
with the addition of GO implies that the addition of GO did not
significantly affect the micelle geometry or packing state, which is
reasonable considering the very small amount of GO compared to
Pluronic. On the other hand, by adding GO, 13 wt% Pluronic solution
showed a sol-to-gel transition at ∼28 ◦ C, thus formed a gel state,
whereas Pluronic solution at 13 wt% without GO did not form a gel
state at any temperature. More interestingly, the phase boundary
of gel-to-sol transition at high temperature disappeared by adding
GO (Fig. 2a), which means that at a given concentration, once the
solution becomes a gel state by increasing temperature, it does not
become a sol state again by further increasing temperature. The gel-
to-sol transition of Pluronic solution at a fixed concentration results Fig. 3. In vitro release profiles of VEGF from Pluronic (17 wt%) or Pluronic-based
from the shrinkage of micelles by the shortening of the PEO part of nanogel (11 wt%) with carboxylated GO containing (GO: 0.1 wt%, n = 3) at 37 ◦ C in
Pluronic as temperature increases, which induces the reduction of 100 rpm environment (n = 3).
micelle packing state. The disappearance of gel-to-sol phase bound-
ary by adding GO suggests that GO might strongly interact with release rate of VEGF from nanogel, the release of VEGF from GO-
Pluronic micelles in a packed state, so it could prevent the reduc- containing hydrogel was mainly determined by the degradation of
tion of micelle packing by further increasing temperature, thus to the hydrogel, which was also valid for hydrogel formed by Pluronic
maintain the physical gel state. Overall, the results of phase dia- micelles. In summary, hydrogels with significantly enhanced sta-
gram of Pluronic micelle indicates that by adding GO, the sol-to-gel bility by adding a small amount of GO were also useful for achieving
transition (and the gel state formation) by Pluronic micelle packing sustained release of loaded proteins.
could occur at a lower concentration of Pluronic, and the gel-to-
sol transition at high temperature disappeared, all supporting the 3.4. The effect of GO on the mechanical properties of hydrogel
strong interaction between GO and Pluronic micelles.
The solution of Pluronic-based nanogel also showed the sol–gel The gelation kinetics and storage moduli of hydrogels were
transition by the packing of nanogels. Compared to Pluronic solu- observed by using a rheometer. Hydrogels as a sol state were
tion, the sol-to-gel transition was observed at lower concentration loaded on a parallel plate geometry holder at 4 ◦ C. The gelation
and lower temperature. Besides, the gel state of nanogel did not was achieved instantaneously by increasing temperature to 37 ◦ C.
become a sol state even at high temperature (no gel-to-sol tran- By adding GO, moduli of hydrogels were increased in all cases.
sition by further increasing temperature at a fixed concentration); And, at the same polymer concentration, the moduli of hydrogels
as a chemically cross-linked state, the packing of nanogel did not increased by increasing concentration of GO from 0.04 to 0.1 wt%
seem to become loose significantly by increasing temperature. By (Fig. 4). Therefore, the addition of GO was effective for increasing
adding GO (0.1 wt%) to the nanogel solution, the sol-to gel phase the mechanical properties of hydrogels formed by Pluronic micelle
boundary shifted slightly to a lower concentration and tempera- or Pluronic-based nanogels, due to making more efficient physi-
ture; at a fixed concentration, the sol state became a gel state at a cal crosslinks induced by the added GO. The enhanced mechanical
slightly lower temperature, and at a fixed temperature, the sol-to- properties of the hydrogels were similar to the case of other
gel transition was observed at a lower concentration by adding GO. polymers with GO composite [19,20]. The increase in mechani-
Also, at 7 wt%, nanogel with GO showed a sol-to-gel transition at cal properties of the hydrogel by adding GO, however, was not
∼28 ◦ C, but nanogel itself did not form a gel state at any temperature dramatic compared to the increased stability of the hydrogels.
(Fig. 1b). Thus, similar to Pluronic micelle solution, Pluronic-based
nanogel solution also showed that the gelation could occur at a 3.5. In vivo stability and biocompatibility of GO-containing
lower concentration of nanogel by adding GO due to the interac- hydrogels
tion between GO and nanogel. The enhanced interaction among
nanogels mediated by GO might result in the structural stabilization To demonstrate the feasibility of GO-containing hydrogels as an
and also lower the critical gelation concentration and temperature injectable, long-lasting hydrogel in vivo, hydrogels with or without
[6]. GO was injected subcutaneously at the back of the mice. Since both
bare hydrogels and GO-containing hydrogels maintain the charac-
3.3. In vitro release profiles of a growth factor from hydrogels teristic of sol-to-gel transition by temperature change, they were
injected as a sol state at a low temperature. After injection, gel state
The VEGF release profiles from hydrogels with 0.1 wt% car- was formed within 3 min post-injection in mice. To monitor the sta-
boxylated GO were characterized. By forming hydrogels with bility of hydrogels in vivo, the size of hydrogel was monitored after
much enhanced stability, hydrogels with GO showed the sustained removing the skin of the injected mice at several time points: 1 day,
release of loaded VEGF (Fig. 3). Similar to the order of degradation 1 week, 2 weeks, and 4 weeks.
rate, hydrogel formed by Pluronic-based nanogel (11 wt%) showed Bare Pluronic hydrogel disappeared within 1 day (Fig. 5b) in
a more sustained release of VEGF compared to that formed by accordance with the result of in vitro degradation. However, GO-
Pluronic (17 wt%). Except the initial burst (∼10% for nanogel and containing Pluronic gel was degraded very slowly; 84% of the
∼30% for Pluronic), the release patterns of VEGF from hydrogels injected hydrogel was remained after 7 days, and 50% of the hydro-
were similar to the degradation rates of the hydrogels. And, the gel stably existed even after 28 days (Fig. 5a). In the case of
release rates of VEGF were a little faster than the degradation rates Pluronic-based nanogel, it disappeared within 7 days (Fig. 5d), but
of the hydrogels. The smaller initial burst from hydrogel formed GO-containing nanogel was even more slowly degraded than GO-
by the nanogel seemed to be associated with the characteristics containing Pluronic; 91% of the injected hydrogel was remained
of the nanogel that can efficiently encapsulate proteins and slowly after 7 days, and 71% of the hydrogel stably existed after 28 days
release then [30]. However, since the degradation rate of hydro- (Fig. 5c). The degradation rates of GO-containing hydrogels in vivo
gel by the dissociation among nanogel is much faster than the were slower than those in vitro, presumably due to the lack of
520 D.-A. Won et al. / Colloids and Surfaces B: Biointerfaces 128 (2015) 515–521

Fig. 4. Gelation of hydrogel composed of (a) 17 wt% Pluronic and (b) 11 wt%
Pluronic-based nanogel with different concentration of GO from 0 to 0.1 wt%.

shaking and the limited supply of physiological fluid in vivo com-


pared to the in vitro experimental condition. In any case, hydrogels
without GO were degraded too fast to be useful in vivo as well as
in vitro. The formation and long existence of GO-containing hydro-
gels verified that the instability of Pluronic-based hydrogel, which
prevented the practical biomedical application of Pluronic hydro-
gels, could be efficiently overcome by adding a small amount of GO
both in vitro and in vivo.
In addition to the remarkably enhanced in vivo stability, bio-
compatibility of these hydrogels needs to be characterized for
biomedical applications. The safety and toxicity of graphene-
based materials have not been completely resolved [33]. However,
graphene oxide is regarded to be relatively biocompatible com-
pared to other nano-structured carbon like CNT, and there
have been many studies reporting the biomedical application of
graphene oxide with apparently no (or significant) short term tox-
icity issue [34]. For example, PEG-coated graphene oxide showed
non-toxicity for 3 months via several biological analyses in mice
with 20 mg/kg [35]. We also verified no acute cytotoxicity of
graphene oxide coated with Pluronic in vitro [28] and in vivo [25].
Fig. 5. (a) In vivo degradation of hydrogels composed of 17 wt% Pluronic and 0.1 wt%
The histological images of tissues at the injection sites for bare
GO (n = 3). (b) Photos of remaining hydrogel at the injected skin: Pluronic after 1
hydrogels and GO-containing hydrogels were obtained after 3 day (left), and GO-containing Pluronic after 4 weeks (right). (c) In vivo degradation
weeks of injection (Suppl. Fig. 1). Pluronic and nanogel without of 11 wt% Pluronic-based nanogel and 0.1 wt% GO (n = 3). (d) Photos of remaining
GO were already completely degraded, so no sign of polymer was hydrogel at the injected skin: nanogel (upper) and GO-containing nanogel (lower)
after 4 weeks.
found. In the case of GO-containing hydrogels, dark spots were
found due to the presence of GO. However, neither the intense
localization of macrophages or noticeable inflammation response 4. Conclusion
was observed in all cases of GO-containing hydrogels by H&E stain-
ing. Thus, GO-containing hydrogels did not show any sign of acute We report a long-lasting, thermo-sensitive and injectable
biocompatibility issue in vivo. Since, a very small quantity of GO hydrogel formed by Pluronic or Pluronic-based nanogel by adding
(0.1 wt%) with abundant Pluronic, the present formulation may not a small amount (0.1 wt%) of graphene oxide (GO). Simple addition
cause significant toxicity without complete metabolism in vivo. of a small quantity of GO to Pluronic-based gel systems was very
D.-A. Won et al. / Colloids and Surfaces B: Biointerfaces 128 (2015) 515–521 521

efficient to increase the stability of the physical gel in an open [4] K.M. Park, S.Y. Lee, Y.K. Jeong, J.S. Na, M.C. Lee, K.D. Park, Acta Biomater. 5 (2009)
environment. Further carboxylated GO was more effective than 1956.
[5] Y. Lee, H.J. Chung, S. Yeo, C.H. Ahn, H. Lee, P.B. Messersmith, T.G. Park, Soft
bare GO for increasing the stability of the physical hydrogels. The Matter 6 (2010) 977.
in vitro dissolution rate of Pluronic-based hydrogels could be sys- [6] Y.Y. Chen, H.C. Wu, J.S. Sun, G.C. Dong, T.W. Wang, Langmuir 29 (2013) 3721.
tematically lowered by increasing the content of GO. The stable and [7] W.S. Shim, J.S. Yoo, Y.H. Bae, D.S. Lee, Biomacromolecules 6 (2005) 2930.
[8] M. Aliaghaie, H. Mirzadeh, E. Dashtimoghadam, S. Taranejoo, Soft Matter 8
injectable formulations of GO-containing Pluronic-based hydro- (2012) 7128.
gels were demonstrated in vivo by subcutaneous injection into the [9] L. Yu, J. Ding, Chem. Soc. Rev. 37 (2008) 1473.
mice. Existence of the injected GO-containing hydrogels over sev- [10] K. Mortensen, J.S. Pedersen, Macromolecules 26 (1993) 805.
[11] J. Elisseeff, K. Anseth, D. Sims, W. Mcintosh, M. Randolph, R. Langer, Proc. Natl.
eral weeks without noticeable acute inflammation response was
Acad. Sci. U.S.A. 96 (1999) 3104.
confirmed, in contrast to the complete disappearance of Pluronic- [12] D. Cohn, G. Lando, A. Sosnik, S. Garty, A. Levi, Biomaterials 27 (2006) 1718.
based hydrogels without GO in a short time. Therefore, the present [13] M. Morishita, J.M. Barichello, K. Takayama, Y. Chiba, S. Tokiwa, T. Nagai, Int. J.
Pharm. 212 (2001) 289.
approach could significantly increase the potential of practical
[14] S.Y. Lee, G. Tae, Y.H. Kim, J. Biomater. Sci. Polym. Edn. 18 (2007) 1335.
application of Pluronic systems, and we anticipate that the method [15] A.K. Geim, K.S. Novoselov, Nat. Mater. 6 (2007) 183.
of adding GO could also be applied to enhance the stability of other [16] Y. Wang, Z. Li, J. Wang, J. Li, Y. Lin, Trends Biotechnol. 29 (2011) 205.
physical hydrogel systems. [17] L. Feng, L. Wu, X. Qu, Adv. Mater. 25 (2013) 168.
[18] L. Zhang, Z. Wang, C. Xu, Y. Li, J. Gao, W. Wang, Y. Liu, J. Mater. Chem. 21 (2011)
10399.
Acknowledgements [19] K. Haraguchi, T. Takehisa, S. Fan, Macromolecules 35 (2002) 10162.
[20] X. Ma, Y. Li, W. Wang, Q. Ji, Y. Xia, EurPolym. J. 49 (2013) 389.
[21] H. Bai, C. Li, X. Wang, G. Shi, J. Phys. Chem. C 115 (2011) 5545.
Partial financial supports were provided by the Basic Science [22] H. Bai, C. Li, G. Shi, Chem. Commun. 46 (2010) 2376.
Research Program, NRF, MEST (2013R1A2A2A03068802), Korea, [23] Y. Xu, Q. Wu, Y. Sun, H. Bai, G. Shi, ACS Nano. 4 (2010) 7358.
and by the “GIST-Caltech Research Collaboration Project” through [24] D. Han, L. Yan, ACS Sustainable Chem. Eng. 2 (2014) 296.
[25] A. Sahu, W.I. Choi, G. Tae, Chem. Commun. 48 (2012) 5820.
a grant provided by GIST, Korea. [26] J. Liu, G. Chen, M. Jiang, Macromolecules 44 (2011) 7682.
[27] S.Z. Zu, B.H. Han, J. Phys. Chem. C 113 (2009) 13651.
Appendix A. Supplementary data [28] A. Sahu, W.I. Choi, J.H. Lee, G. Tae, Biomaterials 34 (2013) 6239.
[29] J.Y. Kim, W.I. Choi, Y.H. Kim, G. Tae, S.Y. Lee, K. Kim, I.C. Kwon, J. Controlled
Release 147 (2010) 109.
Supplementary data associated with this article can be [30] W.I. Choi, Y.H. Kim, G. Tae, Macromol. Res. 19 (2011) 639.
found, in the online version, at http://dx.doi.org/10.1016/j.colsurfb. [31] W. Zhang, Z. Guo, D. Huang, Z. Liu, X. Guo, H. Zhong, Biomaterials 32 (2011)
8555.
2015.03.002. [32] R.M. Silverstein, G.C. Bassler, T.C. Morrill, Spectrometric Identification of
Organic Compounds, 4th ed., John Wiley and Sons, New York, 1981.
References [33] K.H. Liao, Y.S. Lin, C.W. Macosko, C.L. Haynes, ACS Appl. Mater. Interfaces 3
(2011) 2607.
[34] C. Chung, Y.K. Kim, D. Shin, S.R. Ryoo, B.H. Hong, D.H. Min, Acc. Chem. Res. 46
[1] N.B. Graham, Med. Device Technol. 9 (1998) 18.
(2013) 2211.
[2] T.R. Hoare, D.S. Kohane, Polymer 49 (2008) 1993.
[35] K. Yang, J. Wan, S. Zhang, Y. Zhang, S.T. Lee, Z. Liu, ACS Nano 5 (2011) 516.
[3] W.I. Choi, G. Tae, Y.H. Kim, J. Mater. Chem. 18 (2008) 2769.

You might also like