Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Journal of

Materials Chemistry C
View Article Online
PAPER View Journal | View Issue
Published on 05 June 2013. Downloaded by Korea National University of Transportation on 23/06/2017 06:35:41.

Fabrication of highly fluorescent graphene quantum


dots using L-glutamic acid for in vitro/in vivo imaging
Cite this: J. Mater. Chem. C, 2013, 1,
4676 and sensing
Xu Wu,a Fei Tian,a Wenxue Wang,b Jiao Chen,a Min Wub and Julia Xiaojun Zhao*a

A facile bottom-up method for the synthesis of highly fluorescent graphene quantum dots (GQDs) has
been developed using a one-step pyrolysis of a natural amino acid, L-glutamic acid, with the assistance
of a simple heating mantle device. The developed GQDs showed strong blue, green and red
luminescence under irradiation with ultra-violet, blue and green light, respectively. Moreover, the GQDs
emitted near-infrared (NIR) fluorescence in the range 800–850 nm with an excitation-dependent
manner. This NIR fluorescence has a large Stokes shift of 455 nm, providing a significant advantage for
the sensitive determination and imaging of biological targets. The fluorescence properties of the GQDs,
such as the quantum yields, fluorescence life times, and photostability, were measured and the
fluorescence quantum yield was as high as 54.5%. The morphology and composites of the GQDs were
characterized using TEM, SEM, EDS, and FT-IR. The feasibility of using the GQDs as a fluorescent
biomarker was investigated through in vitro and in vivo fluorescence imaging. The results showed
that the GQDs could be a promising candidate for bioimaging. Most importantly, compared to the
traditional quantum dots (QDs), the GQDs are chemically inert. Thus, the potential toxicity of the
Received 1st May 2013
Accepted 4th June 2013
intrinsic heavy metal in the traditional QDs would not be a concern for GQDs. In addition, the GQDs
possessed an intrinsic peroxidase-like catalytic activity that was similar to graphene sheets and carbon
DOI: 10.1039/c3tc30820k
nanotubes. Coupled with 2,20 -azino-bis(3-ethylbenzothiazoline-6-sulphonic acid) (ABTS), the GQDs can
www.rsc.org/MaterialsC be used for the sensitive detection of hydrogen peroxide with a limit of detection of 20 mM.

Introduction traditional QDs would not be a concern for GQDs. Even though
the mechanism of photoluminescence needs further investiga-
Recently, graphene and its derivative graphene oxide have tion, a wide variety of applications for GQDs have been devel-
attracted remarkable attention due to their extraordinary oped, such as photovoltaic devices,29 biosensing,11,30 drug
optical and electronic properties and biocompatibility.1–6 delivery,31 cellular imaging,19 and interactions with biological
However, applications of graphene in the elds of bioimaging systems.32
and optoelectronics are limited because graphene is a zero- A number of methods have been developed for the synthesis
bandgap semiconductor.7,8 In order to generate the bandgap in of GQDs. Generally, the approaches can be divided into two
graphene, a zero-dimensional derivative, graphene quantum groups: top-down and bottom-up methods. The top-down
dots (GQDs) have been developed.9–17 GQDs have strong methods are based on cutting large carbon materials, such as
quantum connement and edge effects when their sizes are in carbon nanotubes, graphene sheets and carbon bers, to form
the range 3–10 nm.10,15,18,19 Compared to zero-bandgap graphene small GQDs. For example, Pan et al. synthesized blue GQDs by a
sheets, GQDs exhibit stronger photoluminescence for bio- hydrothermal method using a preoxidized graphene sheet.10
imaging,20 light-emitting diodes,21 and biosensors.11 The pho- Zhou et al. broke C–C bonds of graphene oxide using a photo-
toluminescence is not only attributed to the quantum Fenton reaction to produce GQDs.32 With control of the photo-
connement effect, but also to the edge effects of the small Fenton reaction conditions, nanoporous graphene sheets and
graphene quantum dots. Most importantly, compared to GQDs were generated in that order. Using a microwave as an
traditional quantum dots (QDs),22–28 GQDs are chemically inert. energy source, Li et al. synthesized greenish-yellow luminescent
Thus, the potential toxicity of the intrinsic heavy metals in GQDs by cutting graphene oxide under acidic conditions.33
Furthermore, several other methods were developed such as
a
acid treatment coupled with chemical exfoliation of carbon
Department of Chemistry, University of North Dakota, Grand Forks, ND58202, USA.
E-mail: jzhao@chem.und.edu; Fax: +1 7017772331; Tel: +1 7017773610 bers,15 the electrochemical etching of a graphene electrode,34
b
Department of Biochemistry and Molecular Biology, The School of Medicine and and ultrasonication breaking of graphene.35,36 The luminescent
Health Sciences, University of North Dakota, Grand Forks, ND58202, USA GQDs could be obtained from carbon sources by these

4676 | J. Mater. Chem. C, 2013, 1, 4676–4684 This journal is ª The Royal Society of Chemistry 2013
View Article Online

Paper Journal of Materials Chemistry C

top-down methods; however, the reaction conditions and operating voltage of 200 kV. The size distribution of the GQDs
purication processes are complicated and time-consuming. was measured and plotted using Image J soware. Fourier
In contrast to the top-down approaches, bottom-up methods transform infrared (FTIR) spectra were obtained on a FT-IR
fabricate GQDs from small carbon precursors including spectrometer (Peking). An Energy-dispersive X-ray spectroscope
glucose,21 hexa-peri-hexabenzocoronene,37 citric acid,12 and etc.38 (EDS) (Oxford), which was attached to a Hitachi SU8010 eld
The bottom-up methods produced GQDs with controllable emission scanning electron microscope (SEM) with an oper-
Published on 05 June 2013. Downloaded by Korea National University of Transportation on 23/06/2017 06:35:41.

morphologies and well-distributed sizes. However, harsh reac- ating voltage of 5 kV, was used to analyze the relative content of
tion conditions, such as using a strong acid or alkali, long carbon, nitrogen and oxygen in L-glutamic acid and the GQDs.
treatment times, and necessary separation processes, are limi- X-ray diffraction (XRD) proles were obtained with a Bruker AXS
tations of the bottom-up methods. In order to overcome these D8 advanced powder X-ray diffraction system using Cu Ka
limitations, in this work we have developed a simple and rapid radiation. Raman spectra were performed on an Aramis labRAM
bottom-up synthetic method by carbonizing a common amino Raman spectrometer (Horiba JobinYvon) using the helium–
acid, L-glutamic acid (Glu), using a heating mantle for pyrolysis. neon laser at 542 nm. A Shimadzu UV-250 UV/vis spectrometer
This is the rst report using L-glutamic acid for the synthesis of was used for measurement of the UV/vis absorption of GQDs.
highly uorescent nitrogen-doped GQDs by a simple heating The uorescence spectra, lifetime, and quantum yields were
process. The obtained GQDs emitted various wavelengths of measured using a Jobin-Yvon-Horiba Fluorometer 3 Model (FL
uorescence including near-infrared uorescence (NIR). In the 3-11 spectrouorometer). The zeta potential was detected using
NIR region, biological samples have low background uores- a Zetasizer (Marlwen, model of Nano-ZS) by adjusting the pH of
cence signals, providing a high signal to noise ratio. Meanwhile, the GQDs from 2.0 to 12.0 using PBS buffer.
NIR radiation can penetrate into sample matrices deeply due to
low light scattering. Thus, the GQDs are promising labeling Detection of quantum yields
reagents for the sensitive determination and imaging of bio- Fluorescein in 0.1 M NaOH (QY ¼ 92%) was chosen as a refer-
logical targets. The uorescence quantum yield of the developed ence.39,40 The quantum yield was calculated according to the
GQDs was about 5 times higher than that of GQDs from citric following equation:
acids.12 We have demonstrated the feasibility of using the GQDs    2 
in in vitro and in vivo uorescence imaging. Besides bioimaging AST IX hX
FX ¼ FST
applications, the GQDs showed a peroxidase-like catalytic AX IST hST 2
activity that was used for the sensitive detection of H2O2. where the subscripts ST and X denote standard and sample,
respectively. F is the quantum yield. I represents the measured
Experimental section integrated uorescence intensity. A is the absorbance at the
Materials excitation wavelength, and h is the refractive index of the
solvent. The refractive index for the standard and sample is
L-Glutamic acid, Ludox SM-30 colloidal silica (30 wt% suspen-
1.33. The GQDs were excited at 360 nm with an emission range
sion in water), hydrogen peroxide, 2,20 -azino-bis(3-ethyl-
of 380–650 nm.
benzothiazoline-6-sulphonic acid) (ABTS), uorescein
isothiocyanate (FITC) isomer I, rhodamine-101 and phosphate
Detection of uorescence lifetime
buffered saline (PBS) tablets were purchased from Sigma-
Aldrich. The deionized water (18.3 MU cm) was produced from a The uorescence lifetime of the GQDs was measured using
Millpore water purication system. time-correlated single photon counting (TCSPC) with the LED
(370 nm) equipped on a Jobin-Yvon-Horiba Fluorometer 3. A
Ludox SM-30 colloidal silica was used as the reference. The
Synthesis of the GQDs
emission wavelength of the GQDs changed from 455 to 505 and
The GQDs was synthesized by pyrolyzing L-glutamic acid. 650 nm. An emission wavelength of 815 nm was used for the
Briey, 2.0 g L-glutamic acid was added into a glass bottle and NIR uorescence lifetime measurement.
heated to 210  C with a heating mantle. Aer the solid L-gluta-
mic acid changed to liquid, the boiling colorless liquid turned In vitro uorescence imaging of cells
to brown in 45 s, which indicated the formation of GQDs. Then,
The uorescence imaging of cells treated with GQDs was con-
10.0 mL water was added into the solution followed by stirring
ducted at various wavelengths (Zeiss LSM-510 Meta Confocal
for 30 min. When the solution cooled to room temperature, the
Microscope). Briey, murine alveolar macrophage cells (MH-S)
solution was centrifuged at 10 000g for 30 min. The supernatant
were cultured overnight at 37  C on plates in RPMI 1640
was collected aer centrifuging. The resulted GQD solution
medium containing 10% fetal bovine serum in a 5% CO2
could be stored at room temperature for at least 5 months
environment. Then, the cultured cells were washed twice with
without precipitation.
PBS buffer (pH 7.4). Aerwards, a 2 mL aliquot of 10 mg mL1
GQDs in PBS was added to the cell culture medium followed by
Characteristics of the GQDs 1 hour culture at 37  C in a 5% CO2 environment. Finally, the
A JEOL JEM-2100 high-resolution transmission electron MH-S cells were washed twice using the PBS buffer and imaged
microscope (HRTEM) was used to characterize the GQDs at an using the Zeiss LSM-510 Meta confocal microscope.

This journal is ª The Royal Society of Chemistry 2013 J. Mater. Chem. C, 2013, 1, 4676–4684 | 4677
View Article Online

Journal of Materials Chemistry C Paper

In vivo uorescence imaging of mice solid L-glutamic acid was rst directly heated to 210  C using a
heating mantle. This temperature is the boiling point of L-glu-
The in vivo uorescence imaging of mice treated with GQDs was
tamic acid. Under these conditions, L-glutamic acid melted and
carried out using an IVIS imaging system (IVIS Lumina XR,
boiled very soon. The colorless liquid changed to pale yellow
Caliper). The athymic BALB/c-nu mice with weights 20–25 g
and brown within 45 s, indicating the carbonization of L-glu-
were obtained from Charles River (Wilmington, MA), and
tamic acid during pyrolysis. At that moment, distilled water was
maintained in the University of North Dakota animal center. All
Published on 05 June 2013. Downloaded by Korea National University of Transportation on 23/06/2017 06:35:41.

added into the brown liquid to dissolve and disperse the GQDs
animal operations were in accord with the institutional animal
use and care regulations, according to protocol no. 1204-4 and with vigorous stirring to avoid aggregation. Unlike other
1204-5. The feasibility of the GQDs for in vivo uorescence methods, surface passivation agents or other inorganic addi-
tives were not needed for the generation of high quantum yield
imaging was investigated by subcutaneous and intramuscular
GQDs. The residue L-glutamic acid could act as a surface
injection. The mice were anesthetized and injected with 100 mL
passivation agent, resulting in the extremely stable and highly
of 25 mg mL1 GQD PBS solution (pH 7.4) subcutaneously on
uorescent GQDs in solution. The GQDs were formed in one
the back, and 100 mL of 25 mg mL1 GQD intramuscularly in the
step by pyrolysis of the precursor. Moreover, the amino group
right back leg. The uorescence images of the mice were taken
on the L-glutamic acid surface provided active groups for surface
immediately with different excitation and emission lter
combinations. modication of GQDs.

Detection of H2O2 based on the GQD's catalytic activity Characterization of the morphology and composites of the
5.4 mg mL1 GQDs was dissolved in 110 mL of 10 mM Tris–HCl GQDs
buffer (pH ¼ 5.0), followed by the addition of 7.2 mM ABTS. The morphology of the obtained GQDs was rst characterized
The control samples were prepared using the same condi- using high resolution TEM (Fig. 1A). The high crystallinity of the
tions except without GQDs or H2O2, respectively. The absorp- GQDs was shown as in the inset of Fig. 1A. The GQDs' lattice
tion spectra of these three samples were recorded aer 2 min of spacing was accurately measured to be 0.246 nm from the
the reaction. Then, the reaction conditions, such as pH, HRTEM image, which is comparable to that of graphitic
temperature, concentration of ABTS, and concentration of carbon.44 The average diameter of the GQDs was 4.66  1.24 nm
GQDs, were optimized. Different concentrations of H2O2 were according to the statistical calculation of more than 100 dots
added into the solution for the investigation of sensitivity. The (Fig. 1B). The size distribution showed the uniform dimensions
absorbance at 416 nm against a reaction time of 120 s was of the developed GQDs.
measured. The elemental content of the GQDs was analyzed using EDS.
As a control, the relative atomic percentages for L-glutamic acid
Results and discussion were 48.73% (C), 9.54% (N), and 41.73% (O) (Fig. 2A and C),
which is similar to the theoretically relative atomic percentages
Design of the highly uorescent GQDs
of L-glutamic acid (50% C, 10% N and 40% O). Aer the
The objective of this work was to develop a simple bottom-up pyrolysis of L-glutamic acid, GQDs were formed. Their carbon
method for the synthesis of highly uorescent GQDs. The key content increased from 48.73% to 60.01%. Meanwhile, the
factor of our design was the usage of a green and natural oxygen content decreased from 41.73% to 34.60%, and the
L-glutamic acid as the precursor. Traditionally the “bottom up” nitrogen content decreased from 9.54% to 5.39% (Fig. 2B
methods for the preparation of GQDs used some special organic and C). This result indicated that L-glutamic acid was carbon-
precursors, such as 3-iodo-4-bromoaniline,38 resulting in a ized aer pyrolysis.
complicated synthetic process. However, in the synthesis of To conrm the formation of GQDs, XRD patterns and Raman
carbon nanodots, some common organic molecules, including spectroscopy were used to characterize the product. As shown in
citric acid,41 histidine42 and glucose,43 were used as precursors Fig. 3A, the Raman spectrum of the GQDs has two typical peaks
with the assistance of a microwave. These studies revealed the at ca. 1355 cm1 and 1580 cm1. The G band (1580 cm1) was
possibility of using common organic molecules to generate related to the vibration of the sp2-bonded C atoms in the gra-
GQDs. Most recently, using citric acid as the precursor, Chen phene material; and the D band (1355 cm1) corresponded to
et al. successfully developed uorescent GQDs and graphene the destruction of the sp2 network by the sp3-bonded C atoms.
oxide.12 Inspired by this pioneering work, we considered using a Fig. 3B shows a typical XRD prole of the GQDs. It had a broad
natural amino acid as a precursor to synthesize N-doped GQDs
based on pyrolysis of the precursor. Our criteria for the selection
of the precursor were non-toxic, common/natural materials,
containing nitrogen groups for easy surface modication of the
GQDs during their applications, and no further capping
processes needed. Considering these criteria, we selected
L-glutamic acid. Our results demonstrated that L-glutamic acid
was an excellent precursor for making the GQDs. The formation Scheme 1 Schematic diagram of the formation of GQDs through the pyrolysis
of a GQD from L-glutamic acid was illustrated in Scheme 1. The of L-glutamic acid.

4678 | J. Mater. Chem. C, 2013, 1, 4676–4684 This journal is ª The Royal Society of Chemistry 2013
View Article Online

Paper Journal of Materials Chemistry C

spectra (Fig. 4A). Comparing curve a with b in Fig. 4A, the C]C
stretching of graphite was formed as indicated by the appear-
ance of the 3000 cm1 and 1500 cm1 peaks, which demon-
strated the formation of GQDs.45 The broad peak area from
2500–2800 cm1 was from the carboxylic acid O–H stretch.
Meanwhile, the strong peak from 1600 to 1730 cm1 resulted
Published on 05 June 2013. Downloaded by Korea National University of Transportation on 23/06/2017 06:35:41.

from the combination of the carboxylic acid C]O and amide


C]O stretches in the GQDs. The carboxylic acid O–H stretching
Fig. 1 (A) The HRTEM image of GQDs with a scale bar of 5 nm. Inset: a typical
and C]O stretching peaks demonstrated the presence of the
single GQD with a lattice parameter of 0.246 nm. The scale bar is 1 nm; (B) the size carboxyl groups. The amide C]O stretching and newly formed
distribution of GQDs calculated from more than 100 dots. amide N–H stretching peak (3700–3500 cm1) indicated the
formation of the amides between the amine group and the
carboxyl group. The stretching vibrations of amine N–H (ca.
3360 cm1) in both curves indicated the presence of amine
groups in the L-glutamic acid and GQDs. The oxygen-containing
and nitrogen-containing groups indicated the high water solu-
bility and stability in solution.
The zeta potentials of the GQDs at different pHs were also
investigated. As shown in Fig. 4B, the zeta potential changed
from positive 6.81 mV to negative 23.2 mV by increasing the pH
value from 2 to 12. This large change of the zeta potential with
pH was related to the presence of the amino and carboxyl
groups in the GQDs. Because of the reaction activity of the
amino and carboxyl groups, biomolecules, such as antibodies
and aptamers, could be easily conjugated onto GQDs for further
applications.

Fluorescence properties of the GQDs


The signicant features of the developed GQDs rely on their
uorescence properties. We have investigated the uorescence
properties using both spectroscopy and the microscopy. The
obtained UV-vis absorption spectra and photoluminescence
spectra are shown in Fig. 5A. The GQDs showed two obvious
absorption peaks at 238 nm and 335 nm (Fig. 5A, curve a). The
peak around 238 nm was attributed to the p–p* transitions of
C]C bonds. The apparent peak at 335 nm indicated the major
Fig. 2 EDS spectra of L-glutamic acid (A) and GQDs (B), and the relative atomic uniform size of the sp2 clusters in the GQDs even though these
percentages of L-glutamic acid and the GQDs (C). sp2 clusters were doped in the sp3 matrix. The color of the GQDs
in the aqueous solution was yellow (inset of Fig. 5A). When the
GQDs were excited at 360 nm (Fig. 5A, curve b), the GQDs
showed a strong uorescence peak at 440 nm (Fig. 5A, curve c).
More importantly, the uorescence is strong enough to be
observed by the naked eye (inset of Fig. 5A), which has

Fig. 3 (A) Raman spectrum of the GQDs. (B) XRD pattern of GQDs.

(002) peak centered at ca. 20.0 , which is similar to GQDs


synthesized by other methods.21
The structural changes of the L-glutamic acid during the Fig. 4 (A) FT-IR spectra of L-glutamic acid (a) and GQDs (b). (B) The zeta potential
formation of the GQDs were further characterized using FT-IR of GQDs in different pH solutions.

This journal is ª The Royal Society of Chemistry 2013 J. Mater. Chem. C, 2013, 1, 4676–4684 | 4679
View Article Online

Journal of Materials Chemistry C Paper


Published on 05 June 2013. Downloaded by Korea National University of Transportation on 23/06/2017 06:35:41.

Fig. 5 (A) The absorption (curve a), excitation (curve b) and emission spectra Fig. 6 Lifetime measurement curves for reference (a) and GQDs at different
(curve c) of GQDs. Inset: from left to right are the GQDs solutions under bright emission wavelengths with 370 nm LED irradiation. Emission wavelengths: (b)
light, violet, blue and green irradiation, respectively; (B) fluorescence emission 445 nm; (c) 505 nm; (d) 650 nm; and (e) 815 nm.
spectra of GQDs with different excitation wavelengths; (C–E) the confocal fluo-
rescence images of the GQDs under different excitation wavelengths. (C) 359 nm
excitation, DAPI channel; (D) 488 nm excitation, LP 505 nm; and (E) 514 nm
excitation, BP 530–600 nm. the GQDs still showed a strong intensity (47% was bleached)
even aer 18 min of the same irradiation (Fig. 7B and C). This
increased photostability provides the GQDs with great potential
for in vitro and in vivo uorescence imaging applications.
promising applications for uorescence imaging. In addition, We also investigated the stability of the GQDs at different pH
the uorescence emission spectra showed an excitation- values (Fig. 7D). The results showed that at pH 7.0 the GQDs
dependent manner. The emission peak of GQDs shied from had the highest uorescence intensity. In the range of pH 5.0–
415 nm (violet) to 580 nm (red) when the excitation wavelength 9.0, the changes in the uorescence intensity were less than
was changed from 320 nm to 540 nm (Fig. 5B). 1.4%. This range covered the majority of pH values of living
We then investigated the uorescence properties of GQDs systems. When the pH value was below 5.0 or higher than 9.0,
using a uorescence microscope. Upon irradiation with violet the uorescence intensity of the GQDs was reduced by 32.3%.
(330–385 nm), blue (460–500 nm) and green (535–585 nm) light, The signicant stability of the GQDs over a broad pH range
the GQDs emitted strong blue, green and red uorescence, might come from the surface amino groups and carboxyl
respectively (Fig. 5C–E). It should be noted that the precursor of groups, which could resist the pH effect on the uorescence
the GQDs showed no absorption in the violet region and no intensity of GQDs.
uorescence in the range of violet to visible light. Thus, all these Besides the uorescence emission in the blue, green and red
uorescence properties should be derived from the formation regions, the GQDs emitted a relatively lower NIR uorescence
of GQDs. around 815 nm upon 360 nm irradiation. The intensity ratio of
To further explore the uorescence features of the developed 815 nm to 440 nm was 0.2 according to the uorescence spectra
GQDs, we measured their uorescence quantum yield and (Fig. 8A). The NIR peak was signicantly higher than the 2l
lifetime. Using uorescein as a reference dye the quantum yield diffraction peak as shown in Fig. 8B. Similar to the uorescence
of GQDs was measured to be 54.5% with excitation at 360 nm. properties in the visible region, the peak in the NIR region
This relative high quantum yield might be due to the protection showed a red shi if the excitation wavelength became longer
of the residue L-glutamic acid as the surface passivation agent. (Fig. 8B). Meanwhile, the uorescence intensity of the GQDs in
The uorescence lifetime of the GQDs was analyzed using the the NIR region was proportional to the concentration of the
TCSPC method at different emission wavelengths with 370 nm GQDs in the range 0.468–15.000 mg mL1 (Fig. 8C). This rela-
excitation. All three uorescence emissions exhibited well tted tionship indicated that the NIR uorescence must be attributed
triple-exponential functions as shown in Fig. 6. The observed to the GQDs even though the mechanism is not yet clear.
lifetimes of the GQDs at 445 nm, 505 nm, 650 nm and 815 nm
are summarized in Table 1. Their average lifetime was in the
range 1.48 to 2.40 ns. The nanosecond lifetime of the GQDs Table 1 Fluorescence lifetime of the GQDs at different emission wavelengths
demonstrated their potential for optoelectronic and biological with 370 nm excitation
applications.
Compared to traditional organic dyes, such as FITC, the Ex (nm) Em (nm) s1 (ns) s2 (ns) s3 (ns) Av. s (ns) Chi
GQDs not only could emit uorescence at different wavelengths, 370 445 0.81 3.64 11.55 2.40 1.14
but also showed much higher photostability. As shown in Fig. 7, 370 505 3.78 0.80 12.04 2.22 1.27
the uorescence of FITC was bleached by 70% under a confocal 370 650 2.35 6.77 0.47 1.48 1.13
laser within 2 min (Fig. 7A and C). However, the uorescence of 370 815 3.23 0.47 8.70 1.91 0.97

4680 | J. Mater. Chem. C, 2013, 1, 4676–4684 This journal is ª The Royal Society of Chemistry 2013
View Article Online

Paper Journal of Materials Chemistry C

value eliminated the possibility of phosphorescence of the


815 nm peak.

In vitro uorescence imaging using GQDs


Based on these excellent uorescence properties, the bioimag-
ing applications of the uorescent GQDs were further investi-
Published on 05 June 2013. Downloaded by Korea National University of Transportation on 23/06/2017 06:35:41.

gated in vitro and in vivo. The in vitro uorescence imaging


using the developed GQDs was explored at different excitation
wavelengths. Aer the GQDs were incubated with MH-S cells for
1 h, uorescence images of the cells were taken using a confocal
microscope. As shown in Fig. 9B, the MH-S cells incubated with
GQDs in PBS buffer showed strong green (LP475) and red (LP
530) colors upon excitation at 458 nm and 514 nm, respectively.
However, no apparent uorescence was observed from the
control group that was not incubated with GQDs under the
same conditions (Fig. 9A). Our results demonstrated that
the GQDs can emit strong uorescence in the green to red range
for in vitro uorescence imaging. The strong uorescence on the
Fig. 7 Comparison of the photostability of the GQDs with a fluorescent dye MH-S cells incubated with GQDs suggested that GQDs had
molecule FITC under the irradiation of a confocal laser. (A) The confocal fluores- penetrated the cells and were able to label both the cell
cence images of FITC; (B) the confocal fluorescence images of GQDs; (C) the membrane and the cytoplasm of the MH-S cells. The uptake of
photostability of these two fluorescent materials; and (D) the effect of pH on the
GQDs by the MH-S cells would be similar to the pathway for
fluorescence intensity of the GQDs.
nanoparticles, which is endocytosis.46

In vivo uorescence imaging using GQDs


The feasibility of the GQDs for in vivo bioimaging was investi-
gated in mice. A 100 mL aliquot of 25 mg mL1 GQDs was
injected subcutaneously into the back of nude mice and intra-
muscularly into the right back leg. The uorescence images of
the mice under different excitation and emission lters were
captured. As shown in Fig. 10, various excitations from blue to
red centered at 430, 465, 500, 535 and 605 nm were applied for
the in vivo uorescence imaging. The emission bandpass lters
for each excitation wavelength are indicated on the top of each
image in Fig. 10. As the excitation wavelength increased, the
uorescence intensity decreased. Moreover, the detectable
uorescence region extended with the longer excitation and
emission wavelengths from the intramuscular injection spot of
the right back leg. This result demonstrated that longer wave-
lengths had better penetration ability than the short

Fig. 8 (A) Fluorescence emission spectrum of the GQDs in the visible-NIR region.
(B) NIR fluorescence spectra of the GQDs at different excitation wavelengths. (C)
The NIR fluorescence intensity vs. the concentration of GQDs.

Notably, the Stokes shi of this NIR uorescence is as large as


455 nm. The large Stokes shi provided the GQDs with signif-
icant advantages for uorescence bioimaging and biosensing.
The lifetime of the uorescence at 815 nm with 370 nm exci- Fig. 9 Laser scanning confocal fluorescence microscopy images of MH-S cells.
tation was examined (Fig. 6e). The result showed a similar (A) Control MH-S cells without GQDs and (B) the MH-S cells treated with GQDs for
value to that of the visible uorescence. The short lifetime 1 hour. All the scale bars are 20 mm.

This journal is ª The Royal Society of Chemistry 2013 J. Mater. Chem. C, 2013, 1, 4676–4684 | 4681
View Article Online

Journal of Materials Chemistry C Paper

We then optimized the detection conditions for catalyzing the


H2O2 reaction using GQDs. The conditions included pH, reaction
temperature, and concentrations of ABTS and GQDs. The pH was
a critical factor for the detection of H2O2. It was reported that the
optimal pH for a carbon nanoparticle-catalyzed H2O2 reaction
was 4.0.49 Our results showed that, in a low pH range (3.0–5.0),
Published on 05 June 2013. Downloaded by Korea National University of Transportation on 23/06/2017 06:35:41.

the reaction rate was very stable (Fig. 12A). However, the reaction
rate decreased above pH 5.0. Therefore, the pH 5.0 was chosen as
the optimal pH. Meanwhile, the reaction temperature also affect
the catalytic reaction rate. As shown in Fig. 12B, high tempera-
ture increased the reaction rate. Considering normal mamma-
lian body temperature, 37  C was selected as the standard
condition for analyzing H2O2. Apparently, the ABTS as a
substrate could affect the sensitivity of the colorimetric detection
of H2O2. As shown in Fig. 12C, the reaction rate increased as the
concentration of ABTS increased until a concentration of 7.2 mM
was reached. Therefore, 7.2 mM of ABTS was chosen as the
Fig. 10 In vivo fluorescence imaging of mice injected GQDs subcutaneously
(spot a) and intramuscularly (spot b). The images were taken at various excitation optimal concentration. Moreover, the concentration of GQDs
wavelengths and emission wavelengths indicated at the top of each image. A was proportional to the reaction rate (Fig. 12D). However, a high
100 mL aliquot of 25 mg mL1 GQDs was injected into each spot. The following concentration of GQDs would affect the absorption of ABTS.
emission bandpass filters were used: GFP: 515–575 nm; DsRed: 575–650 nm; Therefore, a concentration of 5.40 mg mL1 GQDs was chosen as
Cy5.5: 695–770 nm.
the optimal concentration for detection.
Under the above optimal conditions, the detection of H2O2
was carried out using this GQD-based label-free colorimetric
wavelengths for in vivo uorescence imaging. Considering the
system. As shown in Fig. 13A, with the increase in the concen-
decreased uorescence intensity in the longer wavelengths, the
trations of H2O2, the absorbance increased in a concentration-
optimal excitation and emission bandpass to obtain the highest
dependent manner. This assay provided rapid detection (2 min)
signal-to-background ratio should be investigated in the future
of H2O2 with a linear range from 0.1 mM to 10 mM (Fig. 13B).
for further imaging applications. Overall, in vivo uorescence
The limit of detection was 20 mM according to the calibration
imaging in mice under subcutaneous and intramuscular
curve at a signal-to-background ratio of 3. Thus, our studies
injections indicated that the GQDs could be used as a uores-
suggest that the GQDs may have the potential for applications
cent labeling agent for bioimaging in both the visible and NIR
in biosensing elds.
regions.

Colorimetric detection of H2O2 using GQDs Conclusions


To extend the applications of the GQDs, in addition to the In conclusion, GQDs with excellent uorescence properties and
optical properties, we have investigated the catalytic ability of peroxidase-like activity were developed using the direct and
the GQDs. It was reported that materials containing aromatic
sp2 carbon clusters had intrinsic peroxidase-like catalytic
activity to produce a blue color reaction in the presence of H2O2
and a peroxidase substrate, 3,3,5,5-tetramethylbenzidine
(TMB). For example, both carbon nanotubes and graphene
sheets were used for the label-free detection of glucose and
single-nucleotide polymorphism through their intrinsic perox-
idase-like catalytic activity.47,48 Inspired by this work, we
explored whether our GQDs could be used for label-free color-
imetric sensing as a catalyst. With the addition of H2O2, the
GQDs catalyzed the reduction of hydrogen peroxide as indicated
by the color change of the solution to green in the presence of
the peroxidase substrate ABTS (Fig. 11 inset). The absorbance
change of ABTS at 416 nm could be used to monitor the reaction
rate and the peroxidase-like catalytic activity (Fig. 11).
Compared to the control groups including the solution without
GQDs or H2O2, the signal-to-background ratio was higher than
9. Thus, the peroxidase-like catalytic activity of the GQDs could Fig. 11 UV-vis absorption spectra of the ABTS solution containing (a) 9 mM
be used for the detection of H2O2, an important reactive oxygen H2O2; (b) 2.7 mg mL1 GQDs with 9 mM H2O2; and (c) 2.7 mg mL1 GQDs. Inset:
species in a biomedical context. the corresponding photographs of these three samples.

4682 | J. Mater. Chem. C, 2013, 1, 4676–4684 This journal is ª The Royal Society of Chemistry 2013
View Article Online

Paper Journal of Materials Chemistry C

University of North Dakota for providing the confocal uores-


cence microscope and Mr Blaise Mibeck for XRD prole at UND
EERC.

Notes and references


Published on 05 June 2013. Downloaded by Korea National University of Transportation on 23/06/2017 06:35:41.

1 H. Li, X. He, Z. Kang, H. Huang, Y. Liu, J. Liu, S. Lian,


C. H. A. Tsang, X. Yang and S.-T. Lee, Angew. Chem., Int.
Ed., 2010, 49, 4430.
2 Q. Mei, K. Zhang, G. Guan, B. Liu, S. Wang and Z. Zhang,
Chem. Commun., 2010, 46, 7319.
3 Z. Liu, J. T. Robinson, S. M. Tabakman, K. Yang and H. Dai,
Mater. Today, 2011, 14, 316.
4 Y. Song, K. Qu, C. Zhao, J. Ren and X. Qu, Adv. Mater., 2010,
22, 2206.
5 I. V. Lightcap and P. V. Kamat, J. Am. Chem. Soc., 2012, 134,
Fig. 12 Optimization of the reaction conditions for the detection of H2O2. (A) 7109.
pH; (B) temperature; (C) concentration of ABTS; and (D) concentration of GQDs. 6 X. Yang, G. Niu, X. Cao, Y. Wen, R. Xiang, H. Duan and
Y. Chen, J. Mater. Chem., 2012, 22, 6649.
7 G. Eda, Y.-Y. Lin, C. Mattevi, H. Yamaguchi, H.-A. Chen,
I. S. Chen, C.-W. Chen and M. Chhowalla, Adv. Mater.,
2010, 22, 505.
8 K. P. Loh, Q. Bao, G. Eda and M. Chhowalla, Nat. Chem.,
2010, 2, 1015.
9 S. Kim, S. W. Hwang, M. K. Kim, D. Y. Shin, D. H. Shin,
C. O. Kim, S. B. Yang, J. H. Park, E. Hwang, S. H. Choi,
G. Ko, S. Sim, C. Sone, H. J. Choi, S. Bae and B. H. Hong,
ACS Nano, 2012, 6, 8203.
10 D. Pan, J. Zhang, Z. Li and M. Wu, Adv. Mater., 2010, 22, 734.
11 Y. Dong, G. Li, N. Zhou, R. Wang, Y. Chi and G. Chen, Anal.
Fig. 13 (A) The relationship of absorption with concentration of H2O2 (from the Chem., 2012, 84, 8378.
bottom to the top: 0, 0.1, 0.2, 0.3, 0.4, 0.5, 1.0, 5.0 and 10.0 mM) and (B) the
12 Y. Dong, J. Shao, C. Chen, H. Li, R. Wang, Y. Chi, X. Lin and
calibration curve of the label-free colorimetric method for the detection of H2O2.
G. Chen, Carbon, 2012, 50, 4738.
13 J. Shen, Y. Zhu, X. Yang and C. Li, Chem. Commun., 2012, 48,
3686.
simple pyrolysis of an amino acid, L-glutamic acid. The GQDs
14 Y. Li, Y. Zhao, H. Cheng, Y. Hu, G. Shi, L. Dai and L. Qu,
not only showed a stable, strong excitation-dependent photo-
J. Am. Chem. Soc., 2012, 134, 15.
luminescence with a high quantum yield (54.5%) in the blue to
15 J. Peng, W. Gao, B. K. Gupta, Z. Liu, R. Romero-Aburto, L. Ge,
red range but also provided an excitation-dependent NIR uo-
L. Song, L. B. Alemany, X. Zhan, G. Gao, S. A. Vithayathil,
rescence emission around 815 nm. Even though the mechanism
B. A. Kaipparettu, A. A. Marti, T. Hayashi, J. J. Zhu and
of this NIR uorescence emission is not well understood, the
P. M. Ajayan, Nano Lett., 2012, 12, 844.
application potential of the NIR uorescence is signicant. By
16 J. Shen, Y. Zhu, X. Yang, J. Zong, J. Zhang and C. Li, New J.
using strong visible light irradiation, the GQDs showed the
Chem., 2012, 36, 97.
potential to be used for in vitro and in vivo uorescence imaging.
17 M. Li, W. Wu, W. Ren, H.-M. Cheng, N. Tang, W. Zhong and
Furthermore, the GQDs possessed a peroxidase-like activity
Y. Du, Appl. Phys. Lett., 2012, 101, 103107.
which could catalyze the reduction of H2O2 in the presence of an
18 K. A. Ritter and J. W. Lyding, Nat. Mater., 2009, 8, 235.
ABTS substrate. Using the labelling-free colorimetric method,
19 S. Zhu, J. Zhang, C. Qiao, S. Tang, Y. Li, W. Yuan, B. Li,
the GQDs may have broad applications in detecting H2O2, an
L. Tian, F. Liu, R. Hu, H. Gao, H. Wei, H. Zhang, H. Sun
important reactive oxygen species in a biomedical context.
and B. Yang, Chem. Commun., 2011, 47, 6858.
20 S. Zhu, J. Zhang, S. Tang, C. Qiao, L. Wang, H. Wang, X. Liu,
Acknowledgements B. Li, Y. Li, W. Yu, X. Wang, H. Sun and B. Yang, Adv. Funct.
Mater., 2012, 22, 4732.
This work was supported by the National Science Foundation 21 L. Tang, R. Ji, X. Cao, J. Lin, H. Jiang, X. Li, K. S. Teng,
Grants CHE 0911472 and CHE 0947043 to J. Zhao; Flight C. M. Luk, S. Zeng, J. Hao and S. P. Lau, ACS Nano, 2012,
Attendant Medical Research Institute (FAMRI, Grant #103007), 6, 5102.
National Institute of Health AI101973-01, and AI097532-01A1 to 22 R. Freeman, T. Finder and I. Willner, Angew. Chem., Int. Ed.,
M. Wu. We thank the Edward C. Carlson Imaging Center at the 2009, 48, 7818.

This journal is ª The Royal Society of Chemistry 2013 J. Mater. Chem. C, 2013, 1, 4676–4684 | 4683
View Article Online

Journal of Materials Chemistry C Paper

23 H. Dong, W. Gao, F. Yan, H. Ji and H. Ju, Anal. Chem., 2010, 35 S. Zhuo, M. Shao and S. T. Lee, ACS Nano, 2012, 6, 1059.
82, 5511. 36 J. T. Robinson, S. M. Tabakman, Y. Liang, H. Wang,
24 R. Di Corato, N. C. Bigall, A. Ragusa, D. Dorfs, A. Genovese, H. S. Casalongue, D. Vinh and H. Dai, J. Am. Chem. Soc.,
R. Marotta, L. Manna and T. Pellegrino, ACS Nano, 2011, 5, 2011, 133, 6825.
1109. 37 R. Liu, D. Wu, X. Feng and K. Mullen, J. Am. Chem. Soc., 2011,
25 G. Hong, J. T. Robinson, Y. Zhang, S. Diao, A. L. Antaris, 133, 15221.
Published on 05 June 2013. Downloaded by Korea National University of Transportation on 23/06/2017 06:35:41.

Q. Wang and H. Dai, Angew. Chem., Int. Ed., 2012, 51, 9818. 38 X. Yan, X. Cui and L. S. Li, J. Am. Chem. Soc., 2010, 132,
26 X. Gao, Y. Cui, R. M. Levenson, L. W. K. Chung and S. Nie, 5944.
Nat. Biotechnol., 2004, 22, 969. 39 D. Magde, R. Wong and P. G. Seybold, Photochem. Photobiol.,
27 L. Ye, K.-T. Yong, L. Liu, I. Roy, R. Hu, J. Zhu, H. Cai, 2002, 75, 327.
W.-C. Law, J. Liu, K. Wang, J. Liu, Y. Liu, Y. Hu, X. Zhang, 40 J. Shen and R. D. Snook, Chem. Phys. Lett., 1989, 155, 583.
M. T. Swihart and P. N. Prasad, Nat. Nanotechnol., 2012, 7, 41 X. Zhai, P. Zhang, C. Liu, T. Bai, W. Li, L. Dai and W. Liu,
453. Chem. Commun., 2012, 48, 7955.
28 B. Dubertret, P. Skourides, D. J. Norris, V. Noireaux, 42 J. Jiang, Y. He, S. Li and H. Cui, Chem. Commun., 2012, 48,
A. H. Brivanlou and A. Libchaber, Science, 2002, 298, 1759. 9634.
29 V. Gupta, N. Chaudhary, R. Srivastava, G. D. Sharma, 43 H. Zhu, X. Wang, Y. Li, Z. Wang, F. Yang and X. Yang, Chem.
R. Bhardwaj and S. Chand, J. Am. Chem. Soc., 2011, 133, 9960. Commun., 2009, 5118.
30 X. Ran, H. Sun, F. Pu, J. Ren and X. Qu, Chem. Commun., 44 Y. Baskin and L. Meyer, Phys. Rev., 1955, 100, 544.
2013, 49, 1079. 45 W. Kwon and S. W. Rhee, Chem. Commun., 2012, 48, 5256.
31 X. Sun, Z. Liu, K. Welsher, J. T. Robinson, A. Goodwin, 46 Y. Jin, S. Kannan, M. Wu and J. X. Zhao, Chem. Res. Toxicol.,
S. Zaric and H. Dai, Nano Res., 2008, 1, 203. 2007, 20, 1126.
32 X. Zhou, Y. Zhang, C. Wang, X. Wu, Y. Yang, B. Zheng, 47 Y. Guo, L. Deng, J. Li, S. Guo, E. Wang and S. Dong, ACS
H. Wu, S. Guo and J. Zhang, ACS Nano, 2012, 6, 6592. Nano, 2011, 5, 1282.
33 L.-L. Li, J. Ji, R. Fei, C.-Z. Wang, Q. Lu, J.-R. Zhang, L.-P. Jiang 48 Y. Song, X. Wang, C. Zhao, K. Qu, J. Ren and X. Qu, Chem.–
and J.-J. Zhu, Adv. Funct. Mater., 2012, 22, 2971. Eur. J., 2010, 16, 3617.
34 Y. Li, Y. Hu, Y. Zhao, G. Shi, L. Deng, Y. Hou and L. Qu, Adv. 49 X. Wang, K. Qu, B. Xu, J. Ren and X. Qu, Nano Res., 2011, 4,
Mater., 2011, 23, 776. 908.

4684 | J. Mater. Chem. C, 2013, 1, 4676–4684 This journal is ª The Royal Society of Chemistry 2013

You might also like