Accepted Manuscript
Easy synthesis of highly fluorescent carbon quantum dots from gelatin and their
luminescent properties and applications
Qinghua Liang, Wangjing Ma, Yao Shi, Zhi Li, Xinming Yang
PII: S0008-6223(13)00351-5
DOI: http://dx.doi.org/10.1016/j.carbon.2013.04.055
Reference: CARBON 7986
To appear in: Carbon
Received Date: 30 January 2013
Accepted Date: 17 April 2013
Please cite this article as: Liang, Q., Ma, W., Shi, Y., Li, Z., Yang, X., Easy synthesis of highly fluorescent carbon
quantum dots from gelatin and their luminescent properties and applications, Carbon (2013), doi: http://dx.doi.org/
10.1016/j.carbon.2013.04.055
This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Easy synthesis of highly fluorescent carbon quantum dots from
gelatin and their luminescent properties and applications
Qinghua Liang, Wangjing Ma, Yao Shi, Zhi Li* and Xinming Yang
Technical Institute of Physics and Chemistry, Chinese Academy of Sciences (CAS), 29
Zhongguancun East Road, Beijing, 100190, PR China.
Abstract
A simple approach for the synthesis of fluorescent carbon dots (CQDs) has been
developed by the hydrothermal treatment of gelatin in the presence only pure water.
The as-synthesized CQDs were found to emit blue photoluminescence (PL) with a
maximum quantum yield of 31.6%. Meanwhile, the CQDs exhibit
excitation-dependent, pH-sensitive and up-converted PL properties. Importantly, these
CQDs are demonstrated to be excellent bioimaging agents and fluorescent ink due to
their stable emission, well dispersibility, low toxicity, long emission life time, and
good compatibility with cells and macromolecules.
*
Corresponding author. Fax: +86 10 62554670
E-mail address: lizhi@mail.ipc.ac.cn (Zhi Li)
1
1. Introduction
During the past decades, due to their unique optical and electronic properties
caused by the quantum confinement and edge effects, luminescent quantum dots (QDs)
have attracted considerable interest for the promising candidates to replace the
conventional phosphor materials [1-7]. However, the traditional semiconductor QDs
based on metallic elements, such as CdS, CdSe, PbSe and Ag2S, more or less suffer
from the problems of toxicity, hydrophobicity and high cost, which limits their
practical application. As a new type of the nanocarbon family, carbon quantum dots
(CQDs) have been found to be an intriguing QDs for numerous applications owing to
their favorable luminescent properties, high chemical stability, low toxicity,
biocompatibility, and easy functionalization [8, 9]. Inspired by these superior
advantages, many works have been carried out on the simple synthesis routes and
intensive studies of the photoluminescence (PL) properties. Since its origin production
from carbon nanotubes during the process of electrophoresis in 2004 [10], a variety of
synthesis approaches including pyrolysis [11], electrochemical exfoliation [6, 12-14],
incomplete combustion oxidation [15], acidic oxidation [16], laser ablation [17],
hydrothermal treatments [18], microwave/ultrasonic passivation [7, 19], and plasma
treatment [20] in recent years have been developed to prepare CQDs with various
precursors (such as graphite oxide [21], citric acid [22], glycerol [23], coffee grounds
[24], soy milk [25], grass [26], egg [20], and so on). Despite many impressive
advances, it is still desirable and urgent to rapidly synthesize high-quality CQDs by an
easy and environmentally benign method with low-cost raw materials.
2
Recently, as a traditional soft chemical preparation route, the hydrothermal
synthesis route based on water system is considered to be one of the simplest and
most cost-effective methods owing to its cheap apparatus, simple manipulation,
low energy consumption, good selectivity and preparation can be achieved in a
single step without complex control. This method has been used to prepare CQDs by
many researchers. For instance, Liu and co-workers reported the synthesis of CQDs
with a PL quantum yield (QY) of 6.2% by hydrothermal treatment of grass [26].
Zhang et al. produced fluorescent CQDs with a PL quantum yield of 11% by
using bovine serum albumin and as carbon precursor under the passivation of
decanediamine [27]. Sahu and co-workers prepared luminescent CQDs with a PL
quantum yield of 26% from orange juice [18]. It can be seen that, apart from the
synthesis method, an appropriate carbon precursor is another essential factor for
considering. Very recently, Hsu and Chang found that compounds containing
both amino and carboxyl groups are beneficial to synthesize CQDs with high PL
quantum yield [28]. Inspired by this, we assumed that the gelatin that is abundant in
amino and carboxyl groups may be a suitable precursor.
Herein, with a main aim of developing a more sustainable and greener
approach for CQDs production, we report an easy route to synthesize highly
luminescent CQDs by the hydrothermal treatment of gelatin without the
assistance of any chemicals but only pure water. As a relative inexpensive and
water-soluble polymer mainly derived from animal skin and bones, gelatin has
been long used in photographic coating, film-making and food industry. To the
3
best of our knowledge, this is the first time that using the gelatin as the reactants
to fabricate CQDs. The strategy of the synthetic process is illustrated in Fig. 1. A
significant advantage of this method is that neither a strong acid solvent nor a surface
passivation reagent or a complicated post-treatment process is demanded.
Figure 1 A schematic of the synthesis process of photoluminescent CQDs by
hydrothermal treatment of commercial gelatin.
2. Experimental
2.1 Chemicals
Chemical reagents were purchased from the Beijing Chemical Company and
used as received unless otherwise specified. Inertia gelatin (No.55667) was obtained
from Rousselot in France, and was used without further purification. The incubation
media for cells were purchased from Thermo Fisher. Distilled water was used
throughout the experiments.
2.2 Sample preparation
The typical experimental procedure is described as follows: first, 0.8 g gelatin
was added to 40 mL water. Subsequently, the above admixture was poured into a
stainless steel autoclave with a Teflon liner of 50 mL capacity and heated at 200 ℃
4
for 3 h. Finally, the reactor was automatically cooled to room temperature. The
resulting light yellow solution was centrifuged at 16000 rpm for 30 min to remove
weight precipitate and agglomerated particles and then yielded a light brown aqueous
solution of CQDs for further characterization. The yield determined by freeze-dried
method was calculated to be approximate 38.6%.
2.3 Characterization
Transmission electron microscope (TEM) images were taken on a JEOL JEM
2100 TEM at an accelerating voltage of 200 kV. Prior to TEM analysis, the sample
was dropped on a Cu grid coated with an ultrathin amorphous carbon film, and
then dried under ambient condition. UV-Vis absorption was carried out using a Varian
Cary 500 ultraviolet-visible-near infrared (UV-Vis-NIR) spectrophotometer.
Fluorescence spectroscopy was performed using a fluorescence spectrophotometer
(Model Cary eclipse) equipped with a 120 W xenon lamp as the excitation source. The
PL decay curves were recorded on a combined fluorescence lifetime and steady state
spectrometer (F900) using a time-corrected single photon counting system. The
excitation light source for measuring the decay profile of up-converted emission was
the 680 nm line of a laser. Fourier transform infrared spectra (FTIR) investigation was
accomplished by an Excalibur 3100 infrared spectrophotometer using KBr pellets as
the sample matrix in the frequency range 500-4000 cm-1. X-ray photoelectron
spectroscopy (XPS) analysis was characterized by an ESCALab220i-XL electron
spectrometer using 300 W Al Kα as the X-ray source (1486.6 eV), and the energy step
5
size was set as 0.100 eV. Curve fitting of the C1s spectra was achieved by a
Gaussian-Lorentzian peak shape. Powder X-ray diffraction (XRD) was analyzed by a
German Bruker D8 Focus XRD with a graphite monochromatized Cu-Ka radiation
source (λ = 1.5405 Å). The fluorescence microscopy images were taken with an
Olympus BX51TRF microscope. The light source for fluorescence microscopy
observation was a mercury lamp with a fluorescent filter. All measurements were
performed at room temperature.
2.4 Quantum yield measurements
Quantum yield determination was achieved according to previously established
procedure by comparing their integrated PL intensities. Quinine sulfate in 0.1 M
H2SO4 (quantum yield 54%) was chosen as a standard as described [15, 29]. The
absorbance of the aqueous solution of CQDs and reference sample were kept below
0.10 at 350 nm. The calculation of quantum yield was achieved by the following
equation:
( )( ) ( )
φ c = φ q Ac Aq Ic Iq η q η c
Where the subscript “c” and “q” refer to CQDs and quinine sulfate, ф is the PL
quantum yield, A is the absorbance at the excitation wavelength, I is the refractive
index of the solvent.
2.5 Live cell imaging experiments
100 μL suspensions of A549 cells seeded on 96-well plates supplemented with
6
Dulbecco's Modified Eagle Medium (DMEM) that contained 1%
penicillin/streptomycin and 10% fetal bovine serum was carefully cultured in a 5%
CO2 humidified incubator at 37 for 24 h. After removing the incubation media
and then rinsing three times with phosphate buffer solution (PBS), the cells were
further incubated in 1.0 mL fresh DMEM containing 0.3 mL aqueous solution of
CQDs at 37 ℃ for some time. Finally, the cells were washed with PBS for three
times and then imaged using an Olympus BX51TRF microscope.
3. Results and discussion
Fig. 2 (A) shows the TEM image and the diameter distribution of CQDs. It can
be clearly seen that the as-synthesized CQDs are uniform in size and possess a nearly
spherical shape. The Gaussian fitting curve reveals that the average size of the CQDs
is about 1.7 nm, as determined by statistical analysis of more than five hundred
particles by using the ImageJ software (Fig. 2B).
Figure 2 (A) Low and high-magnified (inset) TEM images and (B) the diameter
distribution of the CQDs.
7
In order to determine the chemical composition of these CQDs, XPS
technique was characterized in detail. Four dominant peaks at 168.2, 284.8, 399.1
and 531.5 eV of the XPS survey spectrum depicted in Fig. 3 (A) were attributed
to S2p, C1s, N1s and O1s [25, 30], suggesting the existence of sulfur, carbon,
nitrogen, and oxygen elements. As shown in Fig. 3 (B), the high resolution of the
C1s spectrum displays the peaks at 284.1, 285.7, 287.7 and 289.6 eV, which
demonstrates that the CQDs were functionalized with C−C, C−N, C=O and O−C
groups [25, 26, 31].
Figure 3 (A) XPS survey spectrum, (B) the high-resolution C1s peaks and the fitting
curves, (C) the FTIR spectrum and (D) the XRD pattern of the as-synthesized CQDs.
Meanwhile, the FTIR spectrum was characterized to obtain further structural
insights about the CQDs, as shown in Fig. 3 (C). The characteristic absorption bands
8
of O−H at 3500 cm-1, the stretching vibration band of C=O at 1700 cm-1, and the
stretching vibration bands of C−O at 1100 and 870 cm-1 indicate the presence of
carboxylic acid and other oxygen-containing functional groups [25]. Furthermore, a
broad absorption around the range of 3050-3250 cm-1 and one sharp peak at 1510 cm-1
are assigned to the vibration and deformation bands of N−H, suggesting the existence
of amino-containing functional groups [19, 28]. In addition, two peaks at 2620 and
680 cm-1 ascribed to stretching vibration band of S−H and S−C, respectively, indicate
the containing of sulfur-containing functional groups [16]. Moreover, three obvious
absorption peaks at 2950, 1420 and 1330 cm-1 are associated with the stretch of
vibration C−H, C=C and C−C, suggesting the presence of alkyl and aryl groups [26].
The FTIR studies are consistent well with the XPS investigation. The XRD profile
depicted in Fig. 3 (D) exhibits a weak broad reflection peak centered at around 22°,
suggesting the interlayer spacing of (002) diffraction peak is 0.41 nm. The larger
interlayer spacing than that of graphite (0.34 nm) is ascribed to the existence of
more functional groups [18]. On the basis of the above analysis, a possible
mechanism for the formation of CQDs from gelatin was proposed, as shown in
Fig.S1. During the hydrothermal treatment process, gelatin was initially pyrolyzed to
small protein that was subsequently hydrolyzed into amino acids with abundant amino
and carboxy groups. With continued hydrothermal processing, the amino acids
can be partially decomposed and polymerized, and then carbonized into the
defect-containing CQDs functionalized with abundant oxygen, sulfur and amino
groups [20, 28]. Meanwhile, the hydrophilic groups make the CQDs freely dispersed
9
in water and other polar solvents, such as methanol, ethanol, dimethylformamide
(DMF) and acetonitrile (see Fig. S2 in the Supplementary data). The good
hydrophilicity of the CQDs is favorable for their functionalization and the
applications in detection and diagnostics.
Figure 4 (A) UV-Vis absorption spectrum and the digital image, (B) the PL emission
and excitation spectrum and (C) the fluorescence decay curve of the aqueous solution
of CQDs. The instrument response function (IRF) is also listed for comparison. (D)
The PL spectra of the CQDs excited at different excitation wavelengths, and the inset
shows the corresponding normalized PL emission. (E-G) Fluorescent images taken
10
under the excitation of UV, blue and green light (scale bars: 50 µm).
To study the optical properties of the CQDs, UV-Vis absorption and PL studies
were carried out in detail. As illustrated in Fig. 4 (A), the absorption spectrum exhibits
a broad peak around the range of 250-290 nm, which is ascribed to the typical
absorption of an aromatic π system or the n–p* transition of the carbonyl [19, 25, 28].
The well-dispersed aqueous suspension of CQDs exhibits bright blue emission under
UV light, which could be easily observed with the naked eye and taken with a digital
camera (inset in Fig. 4B). Furthermore, the suspensions of the CQDs in other polar
solvents also exhibited intense blue fluorescence under UV light (Fig. S3). A broad
emission centered at 430 nm was observed in the emission spectrum with an
excitation wavelength at 350 nm. Two peaks at 248 and 350 nm appeared in the
excitation spectrum reveals that the emission may be related to two kinds of
transitions. The quantum yield measured by using quinine bisulfate as standard
sample was 31.6%, which is higher than the previous report (typically < 20%) [11, 23,
26, 27]. The higher quantum yield may be ascribed to the chemical nature and
abundant surface defects of the as-synthesized CQDs. The time-resolved fluorescence
decay curve measured by time-correlated single photon counting method is illustrated
in Fig. 4 (C). The decay curve can be very well fitted to a triple-exponential function.
The lifetimes are τ1 = 10.68, τ2 = 4.15 and τ3 = 0.81 ns, and the mean lifetime is
calculated to be 6.76 ns, which is longer than previous reports [10, 21]. Meanwhile,
almost no bleaching was observed after continuous irradiation with UV light for 8 h,
suggesting the emission was quite stable (Fig. S4). Moreover, the PL spectra also
11
display the characteristic feature that the maximum emission moves to a longer
wavelength with reduced intensity as the excitation wavelength increased. As shown
in Fig. 4 (D), the strongest emission peak shifts from 430 nm to 580 nm and the
intensity decrease gradually as the excitation alters from 340 nm to 500 nm. Thus, the
CQDs exhibited different emission colors including blue, green and red when they
were excited by different excitation wavelengths, as investigated under a fluorescent
microscope (Fig. 4E-4G). This phenomenon has been widely observed in the
luminescent carbon nanomaterials, which may be caused by the optical selection
of different surface defect states near the Fermi level of CQDs [17]. To further
investigate the chemical environment dependence PL behavior, the qualitative
relationship between pH values and emission was studied. Interestingly, the maximum
intensity of the emission spectra obviously decreases under overly acidic or basic
environment (Fig. S4). The reason may be that too many H+ or OH– ions induce the
significant change of functional groups, and then the electronic transition of some
defects would be disrupted or even prohibited.
Significantly, in addition to the down-converted PL properties, the CQDs exhibit
a distinct up-converted characteristic. Inset in Fig. 5 (A) displays a typical PL
spectrum of CQDs excited by 680 nm light with a maximum peak located at 430 nm.
The up-converted emission of the CQDs should be ascribed to the multi-photon active
process, in which the simultaneous absorption of two or more photons leads to the
emission of light at a shorter wavelength than that of the excitation [17]. The
up-converted emission decay profile was also recorded by monitoring at 430 nm with
12
the excitation wavelength at 680 nm. As illustrated in Fig. 5 (A), the lifetime decay
curve is close, but not identical to the instrument response function (IRF), which can
be very well fitted to a double-exponential function, the lifetimes are τ1 = 0.382 and τ2
= 2.57 ns, and the average lifetime is determined to be 0.49 ns that is much shorter
that that of down-converted emission. Most interestingly, similar to the
down-converted emission, the up-converted PL emission exhibits an
excitation-dependent feature as well. Fig. 5 (B) demonstrates the up-converted
luminescence spectra of the CQDs excited by long wavelength ranging from 550 to
1100 nm with the emission located in the range from 330 to 610 nm. The maximum
up-converted emission and the chromaticity coordinates that were calculated from the
corresponding PL spectra were summarized in Table S1. The results imply that the
up-converted emission color of the CQDs can be finely tuned by adjusting the
excitation wavelength, as illustrated in the inset in Fig. 5 (B).
Figure 5 (A) The PL decay profile of the solution of CQDs measured by monitoring
the emission at 440 nm with 680 nm excitation, and the inset exhibits the lifetime data
and the parameter generated by the exponential fitting. The instrument response
function (IRF) is also listed for comparison. (B) The up-converted emission of CQDs
13
excited by different excitation wavelengths and their corresponding color coordinates.
To evaluate the potential of using the highly luminescent CQDs as optical
labels for in vitro cell imaging, A549 cells were chosen as an example, and were
cultivated in the presence of the as-prepared CQDs. As depicted in Fig. 6, strong
blue fluorescence of the A549 cells can be seen after incubation with CQDs for 6
h. Moreover, no autofluorescence emitting from the cells indicates the CQDs
remain exhibited intense PL emission after being internalized into the cells (Fig.
S5). A more careful observation found that the luminescent spots widely
appeared in the membrane and cytoplasmic area of the A549 cell, whereas
fluorescence at the central region involved with nucleus was very weak. In this
case, genetic disruption of the cells may be avoided, which agrees well with the
results of previous investigation on the interaction of living cells with CQDs and
other QDs [8, 32, 33]. And most of the available cell imaging studies of CQDs
generally suggests a possible endocytosis mechanism accounting for the
inhomogeneous distribution of CQDs inside the cell [8, 28, 34]. Elucidating the
exact mechanism of C-dot uptake by cells is still under way. Furthermore, to
evaluate their biocompatibility, the cytotoxicity of the CQDs was subsequently
assessed through the MTT assay, which is described in the Supplementary data
in detail. As can be seen from Fig. S6 and S7, more than 90% of the A549 cells
were remaining alive and exhibited intense blue fluorescence even they were
incubated with the as-prepared CQDs at a high dose of ~240 µg mL-1 for 24 h.
These results firmly demonstrate that the as-synthesized CQDs are of low
14
cytotoxicity. Although the exact mechanism for the endocytosis of CQDs into the
cell is still not clear at this stage, these results safely indicate that such CQDs
possess good biocompatibility and can be effectively applied for in vivo cell
imaging and biological labeling [18, 34, 35].
Figure 6 (A) Bright field image, (B) fluorescence microscopy image obtained
through a band-pass filter of 420 nm under irradiation of UV light (340 nm), (C)
their merged image of A549 cells incubated in the presence of CQDs for 6 h
(scale bar: 50 µm). Objective used was 50× /0.75 NA. The exposure time is 150
ms. Digital photograph of the gelatin-CQDs composites (D, E) and dip pen
writing pattern of the gelatin-CQDs composites on a microscope slide (F, G)
under visible and UV light excitations.
Furthermore, CQDs has been recently pursued for potential application in optical
detectors by dispersing the CQDs in polymer matrixes. For this purpose, an attempt
that 5 mL aqueous solution of CQDs (~0.8 mg mL-1) was added into the solution of
gelatin (50 wt %) at 45 under agitation was conducted to investigate the
15
compatibility with macromolecules without further chemical modifications. Fig. 6 (D,
E) depicts the photograph of the resultant composites under the irradiation of
ordinary and UV light after gelling process. It can be clearly seen that the
composites still exhibited bright blue emission under the irradiation of UV light,
suggesting the CQDs retained their fluorescent properties upon integration into gelatin
matrices. Meanwhile, there is not any evident change of their PL emission spectra
(Fig. S8), indicating no significant surface deterioration or aggregation of the
CQDs occurred. Meanwhile, due to the high transmittance of composites, it can be
hardly seen any information under visible light when the composite was written on a
glass slide through a dip pen (Fig. 6F). However, three characters read “ABC” can be
clearly seen under the irradiation of UV light (Fig. 6G). These results safely
demonstrate that the as-synthesized CQDs exhibit nice compatibility with
polymers, which enables them possible incorporation into solid matrices for
potential applications as optoelectronic devices, anti-counterfeiting ink and
gelatin-based optical waveguides [9, 20, 35, 36]. By the way, highly luminescent
CQDs were also obtained by exploiting protease as the reactants and using
ethanol as the solvent (the reaction temperature was kept at 180 ℃). The TEM
images and luminescence properties of them were shown in Fig. S9 and S10.
Further investigation is still under way.
4. Conclusions
The present work describes a green, cheap and simple approach to synthesize
16
high quality fluorescent CQDs by hydrothermal treatment of commercial gelatin
without any additive and complicated post-treatment. The as-synthesized CQDs were
found to emit strong blue emission with a maximum quantum yield of 31.6% at the
excitation wavelength of 350 nm. Meanwhile, these CQDs exhibit typical
excitation-dependent, pH-sensitive and up-converted PL properties. The results
demonstrate that, coupling with the excellent dispersibility, stable emission, relative
long emission life time, pH-sensibility, low toxicity, good compatibility with cells and
macromolecules, and the distinctive up-converted emission, such CQDs may provide
promising applications in the fields of bioscience, optical device, information
encryption and energy conversion.
Acknowledgments
This work was financially supported by the National Natural Science
Foundation of China (Grant No. 20873169) and the Laboratory of Special
Photographic Materials, Technical Institute of Physics and Chemistry, Chinese
Academy of Sciences (CAS).
Appendix A. Supplementary data
Supplementary data associated with this article can be found in the online version.
References
[1] Hong G, Robinson JT, Zhang Y, Diao S, Antaris AL, Wang Q, et al. In-Vivo
17
fluorescence imaging with Ag2S quantum dots in the second near-infrared region.
Angew Chem Int Ed. 2012;51(39):9818-21.
[2] Zhao Y, Riemersma C, Pietra F, Koole R, de Mello Donegá C, Meijerink A.
High-temperature luminescence quenching of colloidal quantum dots. Acs Nano.
2012;6(10):9058-67.
[3] Zhuo S, Shao M, Lee S-T. Upconversion and downconversion fluorescent
graphene quantum dots: ultrasonic preparation and photocatalysis. Acs Nano.
2012;6(2):1059-64.
[4] Song WS-S, Yang H. Efficient white-light-emitting diodes fabricated from highly
fluorescent copper indium sulfide core/shell quantum dots. Chem Mater.
2012;24(10):1961-7.
[5] Lightcap IV, Kamat PV. Fortification of CdSe quantum dots with graphene oxide.
excited state interactions and light energy conversion. J Am Chem Soc.
2012;134(16):7109-16.
[6] Li HT, He XD, Kang ZH, Huang H, Liu Y, Liu JL, et al. Water-soluble
fluorescent carbon quantum dots and photocatalyst design. Angew Chem Int Edit.
2010;49(26):4430-4.
[7] Li HT, He XD, Liu Y, Huang H, Lian SY, Lee ST, et al. One-step ultrasonic
synthesis of water-soluble carbon nanoparticles with excellent photoluminescent
properties. Carbon. 2011;49(2):605-9.
[8] Li H, Kang Z, Liu Y, Lee S-T. Carbon Nanodots: Synthesis, properties and
applications. J Mater Chem. 2012;22(46):24230-53
18
[9] Baker SN, Baker GA. Luminescent carbon nanodots: emergent nanolights.
Angew Chem Int Ed. 2010;49(38):6726-44.
[10] Xu XY, Ray R, Gu YL, Ploehn HJ, Gearheart L, Raker K, et al. Electrophoretic
analysis and purification of fluorescent single-walled carbon nanotube fragments.
J Am Chem Soc. 2004;126(40):12736-7.
[11] Jia X, Li J, Wang E. One-pot green synthesis of optically pH-sensitive carbon
dots with upconversion luminescence. Nanoscale. 2012;4(18):5572-5.
[12] Long Y-M, Zhou C-H, Zhang Z-L, Tian Z-Q, Bao L, Lin Y, et al. Shifting and
non-shifting fluorescence emitted by carbon nanodots. J Mater Chem.
2012;22(13):5917-20.
[13] Ming H, Ma Z, Liu Y, Pan K, Yu H, Wang F, et al. Large scale electrochemical
synthesis of high quality carbon nanodots and their photocatalytic property.
Dalton Trans. 2012;41(31):9526-31.
[14] Li HT, Ming H, Liu Y, Yu H, He XD, Huang H, et al. Fluorescent carbon
nanoparticles: electrochemical synthesis and their pH sensitive
photoluminescence properties. New J Chem. 2011;35(11):2666-70.
[15] Liu HP, Ye T, Mao CD. Fluorescent carbon nanoparticles derived from candle
soot. Angew Chem Int Edit. 2007;46(34):6473-5.
[16] Liu S, Tian J, Wang L, Luo Y, Sun X. A general strategy for the production of
photoluminescent carbon nitride dots from organic amines and their application
as novel peroxidase-like catalysts for colorimetric detection of H2O2 and glucose.
Rsc Adv. 2012;2(2):411-3.
19
[17] Cao L, Wang X, Meziani MJ, Lu FS, Wang HF, Luo PJG, et al. Carbon dots for
multiphoton bioimaging. J Am Chem Soc. 2007;129(37):11318-9.
[18] Sahu S, Behera B, Maiti TK, Mohapatra S. Simple one-step synthesis of highly
luminescent carbon dots from orange juice: application as excellent bio-imaging
agents. Chem Commun. 2012;48(70):8835-7.
[19] Ma Z, Ming H, Huang H, Liu Y, Kang ZH. One-step ultrasonic synthesis of
fluorescent N-doped carbon dots from glucose and their visible-light sensitive
photocatalytic ability. New J Chem. 2012;36(4):861-4.
[20] Wang J, Wang C-F, Chen S. Amphiphilic egg-derived carbon dots: rapid plasma
fabrication, pyrolysis process, and multicolor printing patterns. Angew Chem Int
Ed. 2012;124(37):9431-5.
[21] Wang Q, Zheng H, Long Y, Zhang L, Gao M, Bai W. Microwave–hydrothermal
synthesis of fluorescent carbon dots from graphite oxide. Carbon.
2011;49(9):3134-40.
[22] Dong Y, Shao J, Chen C, Li H, Wang R, Chi Y, et al. Blue luminescent graphene
quantum dots and graphene oxide prepared by tuning the carbonization degree of
citric acid. Carbon. 2012;50(12):4738-43.
[23] Lai CW, Hsiao YH, Peng YK, Chou PT. Facile synthesis of highly emissive
carbon dots from pyrolysis of glycerol; gram scale production of carbon
dots/SiO2 for cell imaging and drug release. J Mater Chem. 2012;22(29):14403-9.
[24] Hsu PC, Shih ZY, Lee CH, Chang HT. Synthesis and analytical applications of
photoluminescent carbon nanodots. Green Chem. 2012;14(4):917-20.
20
[25] Zhu C, Zhai J, Dong S. Bifunctional fluorescent carbon nanodots: green synthesis
via soy milk and application as metal-free electrocatalysts for oxygen reduction.
Chem Commun. 2012;48(75):9367-9.
[26] Liu S, Tian J, Wang L, Zhang Y, Qin X, Luo Y, et al. Hydrothermal treatment of
grass: a low-cost, green route to nitrogen-doped, carbon-rich, photoluminescent
polymer nanodots as an effective fluorescent sensing platform for label-free
detection of Cu(II) ions. Adv Mater. 2012;24(15):2037-41.
[27] Zhang Z, Hao J, Zhang J, Zhang B, Tang J. Protein as the source for synthesizing
fluorescent carbon dots by a one-pot hydrothermal route. Rsc Adv.
2012;2(23):8599-601.
[28] Hsu P-C, Chang H-T. Synthesis of high-quality carbon nanodots from
hydrophilic compounds: role of functional groups. Chem Commun.
2012;48(33):3984-6.
[29] Kwon W, Rhee S-W. Facile synthesis of graphitic carbon quantum dots with size
tunability and uniformity using reverse micelles. Chem Commun.
2012;48(43):5256-8.
[30] Zhang Y-Q, Ma D-K, Zhuang Y, Zhang X, Chen W, Hong L-L, et al. One-pot
synthesis of N-doped carbon dots with tunable luminescence properties. J Mater
Chem. 2012;22(33):16714-8.
[31] Li Y, Zhao Y, Cheng H, Hu Y, Shi G, Dai L, et al. Nitrogen-doped graphene
quantum dots with oxygen-rich functional groups. J Am Chem Soc.
2011;134(1):15-8.
21
[32] Yildiz I, Deniz E, McCaughan B, Cruickshank SF, Callan JF, Raymo FM.
Hydrophilic CdSe-ZnS core-shell quantum dots with reactive functional
groups on their surface. Langmuir. 2010;26(13):11503-11.
[33] Yildiz I, McCaughan B, Cruickshank SF, Callan JF, Raymo FM.
Biocompatible CdSe-ZnS sore-shell quantum dots coated with hydrophilic
polythiols. Langmuir. 2009;25(12):7090-6.
[34] Luo PG, Sahu S, Yang S-T, Sonkar SK, Wang J, Wang H, et al. Carbon
"quantum" dots for optical bioimaging. J Mater Chem B.
2013;1(16):2116-27.
[35]Fowley C, McCaughan B, Devlin A, Yildiz I, Raymo FM, Callan JF. Highly
luminescent biocompatible carbon quantum dots by encapsulation with an
amphiphilic polymer. Chem Commun. 2012;48(75):9361-3.
[36] Manocchi AK, Domachuk P, Omenetto FG, Yi H. Facile fabrication of
gelatin-based biopolymeric optical waveguides. Biotechnol Bioeng.
2009;103(4):725-32.
22