The World Scientific Encyclopedia of Nanomedicine and Bioengineering II, 3 Volumes Set
The World Scientific Encyclopedia of Nanomedicine and Bioengineering II, 3 Volumes Set
Published
Vol. 6 Handbook of Immunological Properties of Engineered Nanomaterials
(In 3 Volumes)
edited by Marina A. Dobrovolskaia and Scott E. McNeil
(Leidos Biomedical Research Inc., USA)
Forthcoming titles
World Scientific
NEW JERSEY • LONDON • SINGAPORE • BEIJING • SHANGHAI • HONG KONG • TA I P E I • CHENNAI • TOKYO
For photocopying of material in this volume, please pay a copying fee through the Copyright Clearance Center,
Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. In this case permission to photocopy is not required from
the publisher.
Printed in Singapore
Preface
vi Preface
hybrid system are lower than those of the individual nanoparticles, and usually
decrease over time. In addition, the sizes of the hybrids significantly increase
compared with those of the individual nanoparticles. Therefore, one type of
nanoparticle exhibiting multifunctional properties may be more useful than the
single system incorporated with different types of nanoparticles.
Semiconductor quantum dot (QD) is one of the good candidates which
meet the above requirements: small size and multifunctional properties.
These nanocrystals are spherical in shape with only 1–10 nm in diameter.
QDs have attracted great interest in biomedical research fields mostly due to
the excellent optical properties such as size-dependent fluorescence, high
fluorescent quantum yield and long-term photostability. They have been
widely applied for in vitro and in vivo fluorescent labeling and imaging. QDs
are also excellent photodynamic therapy (PDT) and photothermal therapy
(PTT) agents which can significantly inhibit mouse tumor growth under red
or near-infrared (NIR) laser irradiation. However, the components of QDs
are mainly from groups II–VI, III–V, or IV–VI, which are composed of toxic
atoms (e.g. cadmium).
Graphene quantum dot (GQD) is another QD which has similar proper-
ties as the semiconductor QD but has no toxic heavy metal atoms. GQDs are
multifunctional nanomaterials with less than 10 nm in size, which are usually
called zero-dimensional (0D) graphene nanosheets.
Both “top-down” and “bottom-up” techniques have been introduced to
prepare such small nanosheets. For example, large-sized graphene can be
physically cut into pieces through strong acid treatment; small aromatic
molecules after pyrolysis or carbonization can be built into GQDs.
Since GQDs are only several nanometers and have only one or several
layer(s) of carbon atoms, GQDs have huge specific surface areas for loading
drugs. Both the two faces and edge of the GQDs can be loaded with drug
through π–π stacking interaction between the drugs and GQDs and/or
through electrostatic adsorption. Due to the small size, the drug-loaded
GQDs after intravenous injection may conveniently accumulate in tumor
tissue through EPR effect. Therefore, GQDs may be excellent drug delivery
tools. For the semiconductor QDs, however, they cannot be efficiently
loaded with drugs. As for the large-sized graphene nanosheets, they may not
be suitable for intravenous injection compared with the GQDs.
Although GQDs are small-sized graphene nanosheets, some of their
physical properties are different from those of the large-sized graphene. For
example, large graphene sheets are semimetals or specific semiconductors
with zero bandgap energy. However, when the graphene size decreases to less
than 10 nm (has been changed to GQDs), these small graphene sheets have
Preface vii
remarkable quantum confinement and edge effects, which have great poten-
tial applications in o ptical and electronic devices. It has been demonstrated
that GQDs have stable fluorescence ranging from blue to red band. Therefore,
GQDs can be used as fluorescent probes for in vitro labeling and in vivo tar-
geted imaging. The fluorescence of most organic fluorescent dyes and semi-
conductor QDs is not stable in physiological environment, whereas GQDs
usually exhibit resistance to photobleaching in vitro and in vivo, which make
them as potential nanoprobes for long-term tracking the intracellular activi-
ties, biodistribution and metabolism of drugs in vivo.
GQDs are not only electron donors but also electron acceptors.
Therefore, resonance energy transfer (RET) usually occurs between the
GQDs and molecules or particles around the GQDs. Chemical sensor and
biosensor of GQDs have been recently developed based on the fluorescence
RET or luminescence RET. These GQD-based sensors have great potential
for medical, pharmaceutical and food detections.
Upon blue or red laser irradiation, GQDs absorb laser light energy and
produce reactive oxygen species (ROS), especially singlet oxygen (1O2).
Bacterials can be killed by these ROS produced by the laser-triggered GQDs.
Although GQDs have broad light absorption spectrum, the absorption band
mainly locates ranging from ultraviolet to visible wavelength. GQDs under
NIR laser irradiation may therefore produce little ROS. However, PDT in
clinic has been usually applied for the treatment of superficial tumors such as
skin cancer, nasopharyngeal carcinoma and esophageal cancer, and red laser
has been successfully used for irradiating the photosensitizers in tumors.
Therefore, GQDs may have potential for in vivo cancer PDT.
In addition, GQD aqueous solution with a suitable concentration can
convert laser light energy into heat. This means that GQD is also a
photothermal agent which can be used for cancer PTT. This is an interesting
phenomenon since GQDs are both PDT and PTT agents. The growth of
mouse tumors was significantly inhibited after the tumors were injected with
high-dose GQDs and irradiated with a red (e.g. 671 nm) laser, which may be
due to the synergistic effect of PDT and PTT.
The toxicity of the GQDs to cells and animals may be mainly dependent
on the GQD concentration. Under the same dose, GQDs are safer than other
interesting nanoparticles such as semiconductor QDs and graphene oxide
(GO). For introducing the GQDs into future clinical uses, the in vivo
metabolic pathways and potential toxicity should be further investigated.
In this book, the synthesis strategies and optical properties of the GQDs,
plasmon behavior in GQDs, GQDs for medical and pharmaceutical analysis
as well as food detections, in vitro and in vivo fluorescent imaging, drug
viii Preface
delivery and the toxicity of the GQDs to cells and animals were summarized
and discussed. The front line researchers, who are focusing on the research
of graphene-based nanomaterials and nanobiomedicine, were specially
invited by me, and have written six chapters. I also contributed two chapters
and designed the book cover. I would like to take this opportunity to grate-
fully acknowledge all invited authors for their excellent contributions to this
book. I hope this book will provide useful information on the preparation
and properties as well as the biomedical applications of GQDs.
ix
Contents
Prefacev
About the Editorix
Chapter 1
Properties and Synthesis Strategies of Graphene
Quantum Dots 1
Jiajia Zhang and Hongbin Lu
Chapter 2
Synthesis and Amino-Functionalization of the Graphene
Quantum Dots 19
Limin Dong and Kejia Wu
Chapter 3
Plasmons in Graphene Quantum Dots 39
Haifeng Yin
Chapter 4
Application of Graphene Quantum Dots in Medical
and Pharmaceutical Analyses 57
Xiaolei Zhang, Jing Wang and Gongjun Yang
Chapter 5 Graphene Quantum Dots for Food Analysis 77
Yongkang Ye and Xiaodong Cao
Chapter 6 Fluorescent Graphene Quantum Dots for Bioimaging 97
Shuhua Li, Zetan Fan, Fanglong Yuan and Louzhen Fan
Chapter 7 Graphene Quantum Dots for Drug Delivery 115
Yang Chen and Maoquan Chu
Chapter 8 Toxicity of Graphene Quantum Dots 127
Maoquan Chu
Index139
xi
Chapter 1
Quantum dots (QDs) are usually referred to as semiconductor nanoparticles with their
sizes in the quantum-confined regime, in which the excitons are confined in all the
three spatial dimensions. Typical QDs are inorganic semiconductor nanocrystals from
the group II–VI elements in the periodic table. However, the applications of these QDs
are limited by their internal disadvantages, including intrinsic toxicity (e.g. in the case
of widely studied CdSe QDs) and problems caused by the colloidal stability. Therefore,
developing new QDs and relevant nanomaterials is necessary. Consequently, graphene
quantum dots (GQDs), a class of zero-dimensional (0D) graphitic nanomaterials, have
attracted increasing attention recently. Compared with semiconductor QDs, GQDs
are superior in terms of low cytotoxicity, high dispersity in water and some polarity
organic solvents, resistance to photobleaching and biocompatibility. In addition, the
properties of GQDs are easier to be tuned through surface chemistry, indicating that
researchers can design GQDs with various functionalities for different applications.
1. Introduction
Graphene, a single carbon atom thick carbon film, has attracted strong atten-
tion from both academia and industry due to its extraordinary optical,
Figure 1. (A) and (B) TEM and AFM images of GQDs. Reprinted with permission from
Ref. [21]. Copyright (2012) American Chemical Society. (C), (D) and (F) Raman, XRD and
XPS spectra of GQDs. Reprinted with permission from Ref. [8]. Copyright (2012) American
Chemical Society. (E) FTIR spectra of GQDs. Reprinted with permission from Ref. [9].
Copyright (2012) WILEY-VCH Verlag GmbH & Co.KGaA. (G) MALDI–TOF mass spec-
trum of GQDs. Reprinted with permission from Ref. [10]. Copyright (2010) American
Chemical Society.
(HR-TEM) images, both (100) in-plane lattice spacing and (002) interlayer
spacing co-exist.6
X-ray diffraction (XRD) is also used to characterize the crystal structure
of GQDs.8 In most cases, the XRD patterns of GQDs show a broad peak with
an interlayer space larger than 0.34 nm (the interlayer space of graphite)
(Figure 1(D)). The broad diffraction peaks arise from the small size effect of
GQDs whereas the increased interlayer space primarily originates from the
introduction of oxygen-containing functional groups and doped heter-
oatoms. Raman spectroscopy is another powerful, non-destructive tool to
characterize GQDs.8 The Raman spectrum of GQDs shows two peaks
centered around 1360 and 1580 cm–1 (Figure 1(C)), corresponding to
D band and G band, respectively. The D band results from the breathing
mode of the sp2 carbon atoms, which is caused by defects such as functional
groups, doped heteroatoms and structural disorders. The G band is the
result of first-order scattering of the E2g mode, corresponding to the in-plane
vibration of sp2 hybridized C–C bonds.
Fourier transform infrared (FTIR)9 and X-ray photoelectron spectroscopy
(XPS) techniques8 are commonly employed to analyze the chemical compo-
nent of GQDs (Figures 1(E) and 1(F)). Matrix-assisted laser desorption/
ionization time of flight mass spectrometry (MALDI–TOF-MS) is thought
to be the best way to survey GQDs on an ensemble level (Figure 1(G)).10 The
spectra of MALDI–TOF-MS can yield reasonable, yet no quantitative,
estimation for the chemical composition of GQDs, which can be taken as a
supplement of FTIR and XPS.
5
(2014) WILEY-VCH Verlag GmbH & Co.KGaA. (D) Mechanism for the broadband emission in the N-GQDs. Reprinted with permission from
Ref. [18]. Copyright (2013) Nature Publishing Group.
b2227-P2-V1 Synthesis and Biomedical Applications of Graphene Quantum Dots
3. Synthesis of GQDs
Similar to graphene, the synthetic strategies of GQDs can be divided into two
categories: top-down strategy and bottom-up strategy.36 The PL behavior of
GQDs depends on the synthesis strategies. To alter the fluorescence proper-
ties of GQDs and enhance their quantum yield, post-modification such as
chemical doping and surface functionalization is usually performed. In the
following, we summarize the synthesis and modification strategies of GQDs
with emphasis on the reaction mechanisms. Considering the difficulty in
revealing the internal mechanisms of GQDs, we highlight stepwise organic
synthesis strategy because well-defined GQDs can be prepared by this
method, which is important to reveal the origin of PL of GQDs and unravel
new properties and applications.
cutting oxidized graphene sheets into nanosize GQDs (Figure 2(A)).16 In this
method, graphite oxide was prepared first by Hummers or modified Hummers
method. The graphite oxide obtained was thermally deoxidized in tube fur-
nace to give graphene sheets (GSs). Subsequently, the GSs were re-oxidized in
the mixture of H2SO4 and HNO3 (volume ratio = 1:3) under mild sonication
and the oxidized GSs were dispersed in water to form a pH = 8 suspension.
The solution was then treated at 200°C for a period of time to yield GQDs.
Fluorescence spectra prove that the as-synthesized GQDs emit bright blue
fluorescence even in the neutral solution. The oxidized GSs are considered to
have more carbonyl groups rather than epoxy groups, as shown in FTIR spec-
tra in which the oxidized GSs exhibit a strong carbonyl signal and a weak epoxy
signal. The cutting mechanism was presented on the basis of the above results,
namely, the presence of the linear defects makes the oxidized GSs easier to be
attacked under hydrothermal condition. Besides, the portion of ultrafine pieces
surrounded by epoxy lines and edges are further broken down when the bridg-
ing oxygen atoms in epoxy chains are removed. This results in the formation
of GQDs. In addition, it is likewise possible that the fluorescence-emitting
might arise from free zigzag sites in a carbene-like triplet ground state. This
mechanism is consistent with the pH-dependent fluorescence property, that is,
the GQDs emit bright fluorescence under alkaline and neutral conditions, but
under acidic conditions, the fluorescence is nearly completely quenched.
Modified hydrothermal treatment. After the above seminal work, many
further studies regarding hydrothermal synthesis of GQDs through top-down
strategy have been published.48 In view of the fact that the GQDs fabricated
in the previous work are highly disordered, Pan et al. improved their cutting
route. They employed the high-temperature thermally reduced graphite
oxide as the precursor and then treated it under strongly alkaline condition.
The resulting GQDs are very small, with a lateral size of about 3 nm, well
crystallized. They emit bright, stable green fluorescence, and have good
reversibility under different pH conditions. Despite the difference in crystal-
linity, however, the fluorescence emitting of these GQDs is not tunable. Thus,
Tetsuka et al.49 designed another hydrothermal approach to synthesize GQDs
with edge-terminated primary amines. In this work, the GO without thermal
reduction was used as the precursor and hydrothermal treatment was per-
formed in ammonia solution. The ammonia can react with the epoxy group
on GO sheets by nucleophilic substitution, which can not only extract sp2
domains through the ring-opening reaction of epoxy groups but also decorate
the edge of GQDs due to the bonding of amines. The fluorescence of the
extracted GQDs can be simply controlled by adjusting the temperature of
hydrothermal treatment and the concentration of ammonia. As a result, PL
can be tuned from violet to yellow. Ab initio calculations were used to
Figure 3. (A) Hydrothermal synthesis of GQDs from oxidized graphene sheets. Reprinted
with permission from Ref. [16]. Copyright (2010) WILEY-VCH Verlag GmbH & Co.KGaA.
(B) Microwave synthesis of GQDs from GO. Reprinted with permission from Ref. [9].
Copyright (2012) WILEY-VCH Verlag GmbH & Co.KGaA. (C) Electrochemistry synthesis
of GQDs from 3D graphene. Reprinted with permission from Ref. [17]. Copyright (2014)
WILEY-VCH Verlag GmbH & Co.KGaA. (D) Oxidation synthesis of GQDs from coal.
Reprinted with permission from Ref. [18]. Copyright (2013) Nature Publishing Group.
different alkaline aqueous solutions (Figure 4(D)). High quality GQDs can also
be synthesized by rupturing C60 molecules (Figure 4(B)).57,58
Formation mechanisms. Although considerable efforts have been con-
ducted in the synthesis of GQDs via pyrolysis or carbonization of organic pre-
cursors, few efforts were dedicated to the study of the formation mechanisms.
Dong et al.52 attributed the formation of GQDs to the carbonization of citric
acid based on the observation in which the GO grows with extended reaction
times. Tang et al. proposed another different mechanism. Glucose is first dehy-
drated to form the nucleus composed of sp2 carbon atoms, followed by the edge
growth of GQDs. The glucose molecules generate new sp2 carbon by dehydra-
tion at the edge of GQDs, resulting in increasing lateral sizes in the extended
reaction time. Nevertheless, this hypothesis could work only for the specific
process. To further elucidate the formation mechanism, effectively monitoring
the whole reaction process is of great importance. Several points need to be
considered in this regard. (1) How does the transformation of carbon from sp3
to sp2 hybridization occur? (2) How do the functional groups in the raw mate-
rial molecules impact the formation process? (3) How does the re-organization
of the species take place? (4) Whether or not the pyrolysis or carbonization of
different organic precursors has something in common? Apparently, to answer
these questions, there is still a large amount of work to be conducted.
debate, which inhibits to some extent the exploitation of GQDs with new
properties and applications. Apparently, the research for GQDs is presently
still at its early stage and more systematic efforts will largely boost their fur-
ther development and practical applications.
References
1. Zhu YW, Murali S, Cai WW, Li XS, Suk JW, Potts JR, Ruoff RS. Graphene and graphene
oxide: Synthesis, properties, and applications. Adv Mater 2010; 22(35): 3906–3924.
2. Eda G, Lin YY, Mattevi C, Yamaguchi H, Chen HA, Chen IS, Chen CW, Chhowalla M.
Blue photoluminescence from chemically derived graphene oxide. Adv Mater 2010;
22(4): 505–509.
3. Cao L, Meziani MJ, Sahu S, Sun YP. Photoluminescence properties of graphene versus
other carbon nanomaterials. Accounts Chem Res 2013; 46(1): 171–180.
4. Shen JS, Zhu YH, Yang XL, Li CZ. Graphene quantum dots: Emergent nanolights for
bioimaging, sensors, catalysis and photovoltaic devices. Chem Commun 2012; 48(31):
3686–3699.
5. Zhu SN, Song YB, Zhao XH, Shao JR, Zhang JH, Yang B. The photoluminescence
mechanism in carbon dots (graphene quantum dots, carbon nanodots, and polymer dots):
Current state and future perspective. Nano Res 2015; 8(2): 355–381.
6. Zheng XT, Ananthanarayanan A, Luo KQ, Chen P. Glowing graphene quantum dots and
carbon dots: Properties, syntheses, and biological applications. Small 2015; 11(14):
1620–1636.
7. Peng J, Gao W, Gupta BK, Liu Z, Romero-Aburto R, Ge L, Song L, Alemany LB, Zhan
X, Gao G, et al. Graphene quantum dots derived from carbon fibers. Nano Lett 2012;
12(2): 844–849.
8. Li Y, Zhao Y, Cheng H, Hu Y, Shi G, Dai L, Qu L. Nitrogen-doped graphene quantum
dots with oxygen-rich functional groups. J Am Chem Soc 2012; 134(1): 15–18.
9. Li L-L, Ji J, Fei R, Wang C-Z, Lu Q, Zhang J-R, Jiang L-P, Zhu J-J. A Facile Microwave
Avenue to electrochemiluminescent two-color graphene quantum dots. Adv Funct Mater
2012; 22(14): 2971–2979.
10. Yan X, CuX, Li BS. Large, solution-processable graphene quantum dots as light absorbers
for photovoltaics. Nano Lett 2010; 10(5): 1869–1873.
11. Colherinhas G, Fileti EE, Chaban VV. Can inorganic salts tune electronic properties of
graphene quantum dots? Phys Chem Chem Phys 2015; 17(26): 17413–17420.
12. Li C, Yue Y. Fluorescence spectroscopy of graphene quantum dots: Temperature effect at
different excitation wavelengths. Nanotechnology 2014; 25(43): 435703.
13. Tang L, Ji R, Cao X, Lin J, Jiang H, Li X, Teng KS, Luk CM, Zeng S, Hao J, et al. Deep
ultraviolet photoluminescence of water-soluble self-passivated graphene quantum dots.
ACS Nano 2012; 6(6): 5102–5110.
14. Deng X, Sun J, Yang S, Shen H, Zhou W, Lu J, Ding G, Wang Z. The emission wave-
length dependent photoluminescence lifetime of the N-doped graphene quantum dots.
Applied Phys Lett 2015; 107(24): 241905.
15. Wang L, Wang Y, Xu T, Liao H, Yao C, Liu Y, Li Z, Chen Z, Pan D, Sun L, et al. Gram-
scale synthesis of single-crystalline graphene quantum dots with superior optical proper-
ties. Nat Commun 2014; 5: 5357.
16. Pan D, Zhang JC, Li Z, Wu MH. Hydrothermal route for cutting graphene sheets into
blue-luminescent graphene quantum dots. Adv Mat 2010; 22(6): 734–738.
17. Ananthanarayanan A, Wang X, Routh P, Sana B, Lim S, Kim D-H, Lim K-H, Li J, Chen P.
Facile Synthesis of graphene quantum dots from 3D graphene and their application for
Fe3+ sensing. Adv Funct Mater 2014; 24(20): 3021–3026.
18. Ye R, Xiang C, Lin J, Peng Z, Huang K, Yan Z, Cook NP, Samuel EL, Hwang CC, Ruan G,
et al. Coal as an abundant source of graphene quantum dots. Nat Commun 2013; 4: 2943.
19. Shen J, Zhu Y, Chen C, Yang X, Li C. Facile preparation and upconversion luminescence
of graphene quantum dots. Chem Commun 2011; 47(9): 2580–2582.
20. Li M, Wu W, Ren W, Cheng H-M, Tang N, Zhong W, Du Y. Synthesis and upconversion
luminescence of N-doped graphene quantum dots. Appl Phys Lett 2012; 101(10): 103107.
21. Zhuo S, Shao M, Lee S-T. Upconversion and downconversion fluorescent graphene quan-
tum dots: Ultrasonic preparation and photocatalysis. ACS Nano 2012; 6(2): 1059–1064.
22. Gan Z, Wu X, Zhou G, Shen J, Chu PK. Is there real upconversion photoluminescence
from graphene quantum dots? Adv Opt Mate 2013; 1(8): 554–558.
23. Wen X, Yu P, Toh YR, Ma X, Tang J. On the upconversion fluorescence in carbon nano-
dots and graphene quantum dots. Chem Commun 2014; 50(36): 4703–4706.
24. Volk C, Neumann C, Kazarski S, Fringes S, Engels S, Haupt F, Muller A, Stampfer C.
Probing relaxation times in graphene quantum dots. Nat Commun 2013; 4: 1753.
25. Röding M, Bradley SJ, Nydén M, Nann T. Fluorescence lifetime analysis of graphene
quantum dots. J Phys Chem C 2014; 118(51): 30282–30290.
26. Tang L, Ji R, Li X, Teng KS, Lau SP. Size-dependent structural and optical characteristics
of glucose-derived graphene quantum dots. Part Part Systs Char 2013; 30(6): 523–531.
27. Kim S, Hwang SW, Kim M-K, Shin DY, Shin DH, Kim CO, Yang SB, Park JH, Hwang
E, Choi S-H, et al. Anomalous behaviors of visible luminescence from graphene quantum
dots: Interplay between size and shape. ACS Nano 2012; 6(9): 8203–8208.
28. Lingam K, Podila R, Qian H, Serkiz S, Rao AM. Evidence for edge-state photoluminescence
in graphene quantum dots. Adv Funct Mater 2013; 23(40): 5062–5065.
29. Ritter KA, Lyding JW. The influence of edge structure on the electronic properties of
graphene quantum dots and nanoribbons. Nat mater 2009; 8(3): 235–242.
30. Wang S, Cole IS, Zhao D, Li Q. The dual roles of functional groups in the photolumi-
nescence of graphene quantum dots. Nanoscale 2016; 8(14): 7449–7458.
31. Tam TV, Trung NB, Kim HR, Chung JS, Choi WM. One-pot synthesis of N-doped gra-
phene quantum dots as a fluorescent sensing platform for Fe3+ ions detection. Sensor
Actuat B-Chem 2014; 202: 568–573.
32. Sk MA, Ananthanarayanan A, Huang L, Lim KH, Chen P. Revealing the tunable photo-
luminescence properties of graphene quantum dots. J Mater Chem 2014; 2(34): 6954.
33. Permatasari FA, Aimon AH, Iskandar F, Ogi T, Okuyama K. Role of C-N configurations
in the photoluminescence of graphene quantum dots synthesized by a hydrothermal
route. Sci Rep 2016; 6: 21042.
34. Liu F, Jang MH, Ha HD, Kim JH, Cho YH, Seo TS. Facile synthetic method for pristine
graphene quantum dots and graphene oxide quantum dots: Origin of blue and green
luminescence. Adv Mater 2013; 25(27): 3657–3662.
35. Kumar GS, Roy R, Sen D, Ghorai UK, Thapa R, Mazumder N, Saha S, Chattopadhyay
KK. Amino-functionalized graphene quantum dots: Origin of tunable heterogeneous
photoluminescence. Nanoscale 2014; 6(6): 3384–3391.
36. Lin LP, Rong MC, Luo F, Chen DM, Wang YR, Chen X. Luminescent graphene quan-
tum dots as new fluorescent materials for environmental and biological applications. Trend
Anal Chem 2014; 54: 83–102.
37. Kwon W, Kim YH, Lee CL, Lee M, Choi HC, Lee TW, Rhee SW. Electroluminescence
from graphene quantum dots prepared by amidative cutting of tattered graphite. Nano
Lett 2014; 14(3): 1306–1311.
38. Habiba K, Makarov VI, Avalos J, Guinel MJF, Weiner BR, Morell G. Luminescent gra-
phene quantum dots fabricated by pulsed laser synthesis. Carbon 2013; 64: 341–350.
39. Zhang M, Bai L, Shang W, Xie W, Ma H, Fu Y, Fang D, Sun H, Fan L, Han M, et al.
Facile synthesis of water-soluble, highly fluorescent graphene quantum dots as a robust
biological label for stem cells. J Mater Chem 2012; 22(15): 7461.
40. Zhu Y, Wang G, Jiang H, Chen L, Zhang X. One-step ultrasonic synthesis of graphene
quantum dots with high quantum yield and their application in sensing alkaline phos-
phatase. Chem Commun 2015; 51(5): 948–951.
41. Shin Y, Lee J, Yang J, Park J, Lee K, Kim S, Park Y, Lee H. Mass production of graphene
quantum dots by one-pot synthesis directly from graphite in high yield. Small 2014;
10(5): 866–870.
42. Shin Y, Park J, Hyun D, Yang J, Lee JH, Kim JH, Lee H. Acid-free and oxone oxidant-
assisted solvothermal synthesis of graphene quantum dots using various natural carbon
materials as resources. Nanoscale 2015; 7(13): 5633–5637.
43. Lin L, Zhang S. Creating high yield water soluble luminescent graphene quantum dots
via exfoliating and disintegrating carbon nanotubes and graphite flakes. Chem Commun
2012; 48(82): 10177–10179.
44. Sun Y, Wang S, Li C, Luo P, Tao L, Wei Y, Shi G. Large scale preparation of graphene
quantum dots from graphite with tunable fluorescence properties. Phys Chem Chem Phys
2013; 15(24): 9907–9913.
45. Dong Y, Chen C, Zheng X, Gao L, Cui Z, Yang H, Guo C, Chi Y, Li CM. One-step and
high yield simultaneous preparation of single- and multi-layer graphene quantum dots
from CX-72 carbon black. J Mater Chem 2012; 22(18): 8764.
46. Minati L, Torrengo S, Maniglio D, Migliaresi C, Speranza G. Luminescent graphene
quantum dots from oxidized multi-walled carbon nanotubes. Mater Chem Phys 2012;
137(1): 12–16.
47. Dong Y, Pang H, Ren S, Chen C, Chi Y, Yu T. Etching single-wall carbon nanotubes into
green and yellow single-layer graphene quantum dots. Carbon 2013; 64: 245–251.
48. Pan DY, Guo L, Zhang JC, Xi C, Xue Q, Huang H, Li JH, Zhang ZW, Yu WJ, Chen ZW,
Li Z, Wu MH. Cutting sp2 clusters in graphene sheets into colloidal graphene quantum
dots with strong green fluorescence. J Mater Chem 2012; 22(8): 3314–3318.
49. Tetsuka H, Asahi R, Nagoya A, Okamoto K, Tajima I, Ohta R, Okamoto A. Optically
tunable amino-functionalized graphene quantum dots. Adv Mater 2012; 24 (39),
5333–5338.
50. Zhu SJ, Zhang JH, Qiao CY, Tang SJ, Li YF, Yuan WJ, Li B, Tian L, Liu F, Hu R, Gao
HN, Wei HT, Zhang H, Sun HC, Yang B. Strongly green-photoluminescent graphene
quantum dots for bioimaging applications. Chem Commun 2011; 47(24): 6858–6860.
51. Zhou XJ, Zhang Y, Wang C, Wu XC, Yang YQ, Zheng B, Wu HX, Guo SW, Zhang JY.
Photo-Fenton reaction of graphene oxide: A new strategy to prepare graphene quantum
dots for DNA cleavage. ACS Nano 2012; 6(8): 6592–6599.
52. Dong Y, Shao J, Chen C, Li H, Wang R, Chi Y, Lin X, Chen G. Blue luminescent gra-
phene quantum dots and graphene oxide prepared by tuning the carbonization degree of
citric acid. Carbon 2012; 50(12): 4738–4743.
53. Qu D, Zheng M, Zhang L, Zhao H, Xie Z, Jing X, Haddad RE, Fan H, Sun Z. Formation
mechanism and optimization of highly luminescent N-doped graphene quantum dots. Sci
Rep 2014; 4: 5294.
54. Zhang B-X, Gao H, Li X-L. Synthesis and optical properties of nitrogen and sulfur
co-doped graphene quantum dots. New J Chem 2014; 38(9): 4615.
55. Tang L, Ji B, Li XM, Bai GX, Liu CP, Hao JH, Lin JY, Jiang HX, Teng KS, Yang ZB, Lau
SP. Deep ultraviolet to near-infrared emission and photoresponse in layered N‑doped
graphene quantum dots. ACS Nano 2014; 8(6): 6312–6320.
56. Liu R, Wu D, Feng X, Mullen K. Bottom-up fabrication of photoluminescent graphene
quantum dots with uniform morphology. J Am Chem Soc 2011; 133(39):
15221–15223.
57. Chen G, Zhuo Z, Ni K, Kim NY, Zhao Y, Chen Z, Xiang B, Yang L, Zhang Q, Lee Z, et
al. Rupturing C60 molecules into graphene-oxide-like quantum dots: Structure, photo-
luminescence, and catalytic application. Small 2015; 11(39): 5296–5304.
58. Chua CK, Sofer Z, Simek P, Jankovsky O, Klimova K, Bakardjieva S, Kuckova SH,
Pumera M. Synthesis of strongly fluorescent graphene quantum dots by cage-opening
buckminsterfullerene. ACS Nano 2015; 9(3): 2548–2555.
59. Yan X, Cui X, Li LS. Synthesis of large, stable colloidal graphene quantum dots with tun-
able size. J Am Chem Soc 2010; 132(17): 5944–5945.
60. Gupta V, Chaudhary N, Srivastava R, Sharma GD, Bhardwaj R, Chand S. Luminscent
graphene quantum dots for organic photovoltaic devices. J Am Chem Soc 2011; 133(26):
9960–9963.
61. Li N, Than A, Wang X, Xu S, Sun L, Duan H, Xu C, Chen P. Ultrasensitive profiling of
metabolites using tyramine-functionalized graphene quantum dots. ACS Nano 2016;
10(3): 3622–3629.
62. Luo P, Ji Z, Li C, Shi G. Aryl-modified graphene quantum dots with enhanced photolu-
minescence and improved pH tolerance. Nanoscale 2013; 5(16): 7361–7367.
63. Novak TG, Kim J, Song SH, Jun GH, Kim H, Jeong MS, Jeon S. Fast P3HT exciton
dissociation and absorption enhancement of organic solar cells by PEG-functionalized
graphene quantum dots. Small 2016; 12(8): 994–999.
64. Sun H, Wu L, Gao N, Ren J, Qu X. Improvement of photoluminescence of graphene
quantum dots with a biocompatible photochemical reduction pathway and its bioimaging
application. ACS Appl Mater Interfaces 2013; 5(3): 1174–1179.
65. Yang HB, Dong YQ, Wang X, Khoo SY, Liu B. Cesium carbonate functionalized gra-
phene quantum dots as stable electron-selective layer for improvement of inverted poly-
mer solar cells. ACS Appl Mater Interfaces 2014; 6(2): 1092–1099.
66. Yang S, Sun J, He P, Deng X, Wang Z, Hu C, Ding G, Xie X. Selenium doped graphene
quantum dots as an ultrasensitive redox fluorescent switch. Chem Mater 2015; 27(6):
2004–2011.
67. Jin SH, Kim DH, Jun GH, Hong SH, Jeon S. Tuning the photoluminescence of graphene
quantum dots through the charge transfer effect of functional groups. ACS Nano 2013;
7(2): 1239–1245.
68. Sun H, Gao N, Wu L, Ren J, Wei W, Qu X. Highly photoluminescent amino-functional-
ized graphene quantum dots used for sensing copper ions. Chem Eur J 2013; 19(40):
13362–13368.
69. Yan X, Li BS, Li LS. Colloidal graphene quantum dots with well-defined structures.
Accounts Chem Res 2013; 46(10): 2254–2262.
Chapter 2
N,N-dimethyl formamide (DMF) was used as the ammonia source, and amino-
functionalized graphene quantum dots (af-GQDs) were prepared by the microwave-
solvothermal method. The samples were characterized in terms of morphological,
structural, and optical properties which are evaluated by X-ray diffraction (XRD),
transmission electron microscopy (TEM), the photoluminescence (PL) spectrum
and Fourier transform infrared (FTIR) spectra correspondingly. The graphene
oxide (GO) goes through with microwave expansion treatment first, then followed
by solvothermal process. Amino-functionalization has significant influence on the
crystallization degree and the layer spacing of graphene quantum dots (GQDs), and
the size was less than 10 nm, which was closed to zero-dimension state. The af-GQDs
have strong absorption in the ultraviolet region (λ ≤ 276 nm), but it has no significant
characteristic absorption peak, which fits with the semiconductor nanomaterials
ultraviolet absorption spectrum. The fluorescence emission spectrum of GQDs
that synthesized by this method perform excitation-dependent PL behaviors. The
excitation-dependent PL behaviors may be associated with electronic transitions
between electronic surface state level and the lowest unoccupied molecular orbital
(LUMO) energy level.
19
1. Introduction
Nowadays, the main methods for preparing graphene quantum dots (GQDs)
included top-down1,2 and bottom-up strategies.3 Their principles were con-
trolling the size, but in the choice of raw materials was slightly different. The
top-down approach referred to cut the graphene sheets (GSs) (the large size)
by physical or chemical ways into the GQDs (the small size). It consisted of
hydrothermal methods,4 electrochemical strategies5 and chemical exfoliation6
of carbon fibers. It had the advantages of relatively simple steps, higher yield,
and defect of not being able to precisely control the morphological structure
and size of the GQDs. The bottom-up approach referred that using small
molecules carbon sources as precursors through a series of chemical reactions
prepared GQDs. The method mainly involved ultrasonic methods,7 solution
chemistry methods8 and microwave preparation.9 Most of the methods were
highly controllable, but were also complex and complicated. Moreover, the
strict preparation conditions of some special methods limited the promotion
of these methods. So, presently, the study of GQDs mainly was concentrated
on the improvement of preparation methods.
In recent years, the hydrothermal method in the preparation of nanom-
eter materials showed more obvious advantages. More and more researchers
are trying to adopt this method to prepare GQDs. But graphene as essential
material in these methods was complex, costly, and expensive. The yield of
GQDs is generally low with the graphene as the raw materials wasting
resources, so the choice of cheaper and easier materials to prepare GQDs is
currently a research focus in the field of synthetic process.
In the field of biomedical imaging, the luminous intensity of GQDs
directly affects the observation of labeled cells. Yang10 and his team found
that there are a lot of amino groups in the PEG graphene, which showed
good optical absorption and longer blood half-life in solution. According to
this conclusion, we considered bringing in amino in the GQDs, which obvi-
ously improves the PL intensity of the GQDs. It is beneficial to improve the
sensitivity of fluorescence analysis in biological studies and the observation of
labeled cells in the organism.
The fluorescence properties of the GQDs which take the amino-
functionalization were significantly increased, and the level of cytotoxicity
was also the key factor to the af-GQDs application. Whether bringing in
amino enhances the combination of the QDs and non-specific cell or not, it
is a question worth considering. Chen11,12 used modified Photo-Fenton
hydrothermal method to prepare the fluorescence GQDs. He added
(001)
(001)
10 20 30 40 50
2u (°)
Figure 2. TEM images of GO before and after processing. Reprinted with permission from
Ref. [15]. Copyright (2016) Virtual Institute of Physics.
transparent film, and part of the GO wrinkles, these results indicate that the
oxidation of the graphite sheet is more thorough. As shown in Figure 2(C),
the reduction of GO is obtained after microwave heat reduction. It can be
found that agglomeration occurs since the hydrophilic oxygen-containing
groups on the surface of the graphene were removed. The stretch degree of
graphene is weaker than GO. Figure 2(D) shows the TEM spectrum of the
precursor after secondary oxidation process, its size is far smaller than the GO
before secondary oxidation process.
From Figure 2, we can see that with the improvement of the experimental
method, the layer thickness of GO changed significantly. Graphite was oxi-
dized and exfoliated more thoroughly after ultrasonic-assisted Hummers, and
the thinner GO is obtained. The external energy was added in the graphite
oxidation reaction stage, and the ultraphonic provides simultaneous, homoge-
neous, and fast heating. With the intercalation reactions of oxidants, the vibra-
tion frequency of sulfate and metal ions is increased, thus making them easier
to carry out intercalation between the graphite layers. The intercalation agents
were hydrolyzed to oxygen-containing group such as –COOH or –CHO
during the latter part of the high-temperature hydrolysis.
The reduction of graphene oxide (RGO) is obtained after microwave heat
reduction. Compared with the GO, the RGO is nearly transparent, it means
the thickness is further decreased. This is because the microwave treatment
provided enough energy to the GO in a short time, and the unstable oxygen-
containing groups between the layers as well as at the edge of the graphite
were removed in the form of gases. After the secondary oxide, the size of the
GO decreased significantly. As shown in Figure 2(D), samples are in the form
of fragmented distribution, size of the GO is less than 100 nm. The reaction
process is in line with unzipping mechanism.16 In the process of acid oxida-
tion, bridging oxygen atom (C–O–C) tends to be arranged in line, this is easy
to disconnect the C–O bond. Once the epoxy bond was opened, it is easy to
form a more stable C=O/COOH group at room temperature. These groups
form epoxy chain on graphene nanosheets17 (Figure 2). These linear defects
make the GO easily breakable.
1200 (002)
900
b
Intensity (a.u.)
(002)
600
a
300
10 20 30 40 50 60 70 80
2u (°)
Figure 3. XRD spectra of GQDs prepared by different methods. Reprinted with permission
from Ref. [15]. Copyright (2016) Virtual Institute of Physics.
carried out at 190°C for 6 h, then we get the XRD pattern of GQDs
(curve b). From Figure 3, we can find that the intensity of the GQDs (002)
crystal plane diffraction peak increased significantly after microwave-assisted
hydrothermal method, and the crystallinity is also increased, this is consistent
with the TEM. At the same time, the diffraction peak of (002) crystal face is
from 27° to 24° to the left, the interplanar spacing is increased to 0.37 nm.
Microwave provides simultaneous, homogeneous, and fast heating in a short
time. Since the polarization effect of the microwave reaction reduces the acti-
vation energy of the reactants, thereby the rate of reaction is increased.
TEM characterization of the above samples was obtained as shown in
Figure 4. Figure 4(A) shows the TEM pattern of GQDs which were prepared
by the hydrothermal reaction at 190°C for 10 h; when the GO sheets were
prepared by microwave-assisted hydrothermal method, 800 W for 9 min, and
then the hydrothermal reaction was carried out at 190°C for 6 h, then we get
the TEM pattern of GQDs as shown in Figure 4(B). From Figure 4(B), we
can find the size of the GQDs is between 7 and 24 nm with microwave-
assisted, and the average size is about 20 nm, samples approximate spherical
particles, and we can see the lattice fringes. Without microwave processing,
the size of the GQDs is about 24 nm, the lattice fringe is not clear, the
crystallinity is not high. Microwave processing can improve the crystallinity of
the sample, and the smaller graphene nanosheets can be obtained, these facili-
tate the formation of the GQDs.
Figure 4. TEM images of GQDs prepared by different methods: (A) hydrothermal method,
(B) microwave-assisted hydrothermal method. Reprinted with permission from Ref. [15].
Copyright (2016) Virtual Institute of Physics.
GO
3431
Intensity (a.u.)
GQDs
1324 1025
1448
Figure 5. FT-IR spectrums of GO and GQDs. Reprinted with permission from Ref. [15].
Copyright (2016) Virtual Institute of Physics.
800
314 nm 430 nm
600
257 nm
Intensity (a.u.)
400
200
0
200 250 300 350 400 450 500 550 600
Wavelength (nm)
bending vibration peak of the –CH2 appears at 1448 cm–1, while the stretch-
ing vibration peak of aldehyde group at 1729 cm–1 disappeared.
From the figure, we can see the disappearance of the aldehyde group, we
take the sodium hydroxide adjust the pH and introduce the hydroxyl , the
reaction of hydroxyl and hydrogen ions of aldehyde group to produce water,
and the –CH2–CO– is formed. Furans epoxy groups in the carbon–oxygen
bonds were broken, and the bridging oxygen atoms were removed, which is
the key to cut the GO into small-sized GQDs. During the experiment,
hydroxyl functional group introduced by sodium hydroxide has changed the
functional groups types of GQDs. The deprotonation process has a great
influence on the fluorescence properties of the GQDs.
Figure 7. Electron transition diagram of GQDs triplet state. Reprinted with permission from
Ref. [15]. Copyright (2016) Virtual Institute of Physics.
σ→π* and π→π* transition. UV absorption spectrum shows that GQDs have
absorption in the ultraviolet region, which is due to the σ→π* transition of
sp2 aromatic group.
The fluorescence properties of GQDs are related to the carbene structure
of triplet electron transition. For triplet σ1π1 structure, σ and π orbital were
occupied by unpaired electrons. In the process of triplet electron transition,
the electrons from the σ and π orbital track (highest occupied molecular
orbital, HOMO) jump into the π* orbital (lowest unoccupied molecular
orbital, LUMO). In GQD emission process, electrons from π orbital jump
into the LUMO energy levels, and the energy required is 4.15 eV (314 nm).
Electrons from the σ orbital jump into the LUMO energy levels, the energy
required is 5.07 eV (257 nm). As shown in Figure 7, the energy gap (δE)
between the σ and π is 0.92 eV, and the value is consistent with the triplet
carbene theoretical value δE (<1.5 eV).18 Therefore, when the energy reaches
the energy required for electron transitions from σ and π orbital to the
LUMO, electron transitions will occur; fluorescence is generated when the
energy changed from the LUMO to the lower energy state.
(001) (002)
(002)
af-GQDs
Intensity (a.u.)
GQDs
GO
10 20 30 40 50
2 θ (°)
Figure 8. XRD spectra of the precursor, GQDs and af-GQDs. Reprinted with permission
from Ref. [15]. Copyright (2016) Virtual Institute of Physics.
like bread peak, and the peak is not obvious. The diffraction peak of GO
(001) crystal face is located at about 10°, that’s a strong and sharp peak. The
figure shows that the intensity of the af-GQDs (002) crystal plane diffraction
peak is greater than GQDs, the corresponding 2θ diffraction angle is right
shift to 25.7°, and the interplanar spacing is reduced to 0.35 nm. Amino-
functionalization has significant influence on the crystallization degree and
the layer spacing of GQDs. We analyze that when the nitrogen source reacted
with oxygen-containing functional groups in the GS, and it easily reacted with
the functional groups on the edge. Unreacted oxygen-containing functional
groups and bridging oxygen atom are transformed into water or gas, etc., and
they are removed under the high-temperature and high-pressure atmosphere.
This makes the graphene defects decreased, the graphene crystal structure
tends to complete. Af-GQDs interplanar spacing (0.35 nm) is less than the
GO (002) interplanar spacing (0.431 nm) because the oxygen-containing
groups between the graphite layers are removed during the reaction at high-
temperature and pressure.
Figure 9 shows the TEM images of af-GQDs. Figure 9(A) is an enlarged
10,000 times TEM image of the af-GQDs, and Figure 9(B) is an enlarged
60,000 times TEM image of the af-GQDs. The af-GQDs dispersed uniformly
without agglomeration. From the high-resolution electron microscope photo-
graph (HRTEM), we can see that the size of the prepared af-GQDs are about 5–7
nm, it is close to the zero-dimensional (0D) state. The (002) crystal face lattice
fringes of the af-GQDs are legible, and the interplanar spacing is about 0.35 nm.
Figure 9. TEM images of af-GQDs. Reprinted with permission from Ref. [15]. Copyright
(2016) Virtual Institute of Physics.
benzene
GQDs nuclear
Intensity (a.u.)
–OH
af-GQDs C=O
R–NH
Ammonium ion
Absorption band
CH2 C–N
NH
R-N NH
NH2 (wagging
–OH R=O CH3 vibration)
Figure 10. FT-IR spectrums of GQDs and af-GQDs. Reprinted with permission from Ref.
[15]. Copyright (2016) Virtual Institute of Physics.
is also reduced. After a series of reaction, the NH2 and C–N bending or
stretching vibration peaks appear at 2362, 1492, 1263, and 698 cm–1. It
means that nitrogen-containing functional groups were introduced.
416 nm 494 nm
PLE
383 nm
Abs
337 nm PL
Intensity (a.u.)
281 nm
Figure 11. UV–vis absorption and fluorescence spectra of af-GQDs. Reprinted with permis-
sion from Ref. [15]. Copyright (2016) Virtual Institute of Physics.
1000
600 420 nm
800
440 nm
500
460 nm
Intensity (a.u.)
Intensity (a.u.)
0 0
400 450 500 550 600 650 450 500 550 600 650 700
Wavelength (nm) Wavelength (nm)
(A) (B)
Figure 12. The fluorescence emission spectra of the af-GQDs excited by different excitation
wavelengths. Reprinted with permission from Ref. [15]. Copyright (2016) Virtual Institute of
Physics.
structure of GQDs changed, the bandgap between the HOMO and the
LUMO is shortened. We analyze that amino groups were introduced into the
GQDs edge. While reacting with amino group electron orbit, graphene mol-
ecule delocalized π electrons HOMO orbital is moved to the higher energy
orbits, and the bandgap is shortened consequently.
Af-GQDs excitation spectra have four excitation peaks, and they are
located at 281, 337, 383, and 416 nm, respectively. Figure 12(A) shows the
emission spectra of af-GQDs excited by different excitation wavelengths.
Although the excitation wavelength changes, the corresponding af-GQDs
emission peak wavelength does not change. The maximum emission peak
wavelength is about 490 nm. The research shows that with the increase of the
intensity of each excitation peak, the intensities of emission peaks are also
increased. When excited by the 416 nm, the highest emission peak intensity
is obtained. GQDs have an emission characteristic of multiwavelength excita-
tion, we analyze that apart from the σ → π * and π → π * electronic transition
of GQDs carbene structure itself, since the introduction of the nitrogen-
containing functional groups, the electron distribution of the af-GQDs sur-
face is changed.
To study the relationship between the emission spectrum and excitation
wavelength, the af-GQDs were tested by the different excitation wavelengths.
Figure 12(B) shows the emission spectra of af-GQDs which were excitated
by different wavelengths. The inset at the right shows the af-GQDs emitting
light colors which were excitated by the 420, 440, 460, 480, and 500 nm,
respectively. With the excitation wavelength changed from 420 to 500 nm,
the GQDs emission peak wavelength from 497 nm red-shift to 546 nm, and
the emission intensity is followed by decline. The fluorescence emission
spectrum of GQDs that synthesized by this method perform excitation-
dependent PL behaviors. The excitation-dependent PL behaviors may be
associated with electronic transitions between electronic surface state level
and the LUMO energy level.
RGO
300 W
500 W
800 W
1000 W
20 40 60 80
2u (°)
Figure 13. XRD spectra of GQDs which conducted at different microwave power. Reprinted
with permission from Ref. [14]. Copyright (2016) Virtual Institute of Physics.
450
483 nm
400
350 485 nm
300 500 w
Intensity (a.u.)
457 nm
250 466 nm 300 w
200 438 nm 0
150 800 w
100 1000 w
50
0
400 450 500 550 600 650
Wavelength (nm)
Figure 14. The fluorescence emission spectra of GQDs under different microwave power.
Reprinted with permission from Ref. [14]. Copyright (2016) Virtual Institute of Physics.
Intensity (a.u.)
RGO
3 min
5 min
7 min
9 min
10 20 30 40 50 60 70 80
2u (°)
Figure 15. XRD spectra of RGO and GQDs under different microwave times. Reprinted
with permission from Ref. [14]. Copyright (2016) Virtual Institute of Physics.
5 min
3 min
300
0
Intensity (a.u.)
7 min
200
9 min
100
0
350 400 450 500 550 600 650
Wavelength (nm)
Figure 16. The fluorescence emission spectra of GQDs under different microwave times.
Reprinted with permission from Ref. [14]. Copyright (2016) Virtual Institute of Physics.
times. Figure 16, shows the emission spectrum of samples. With the increase
in the microwave time, the fluorescence intensity of GQDs first increased and
then decreased; when the microwave time was set to 5 min, the intensity got
the strongest. The appropriate microwave power and time can improve the
fluorescence properties of GQDs. Besides, it can also improve the quantum
yield, so the fluorescence intensity of GQDs increased. But the high concen-
tration of GQDs in the solution leads to concentration quenching effect, and
the fluorescence intensity decreased instead.
450
483 nm
400
481 nm
250 190˚C
160˚C
200
150
100
50
0
400 450 500 550 600 650
Wavelength (nm)
Figure 17. The fluorescence emission spectra of GQDs at different solvothermal tempera-
tures. Reprinted with permission from Ref. [14]. Copyright (2016) Virtual Institute of Physics.
5. Conclusion
This paper focuses on the preparation and fluorescence properties of GQDs,
and the preparation of the af-GQDs to enrich fluorescence properties. Finally,
the paper reaches the following conclusions:
References
1. Li Y, Zhao Y, Cheng H, et al. Nitrogen-doped graphene quantum dots with oxygen-rich
functional groups. J Am Chem Soc 2011; 134(1): 15–18.
2. Shen J, Zhu Y, Chen C, et al. Facile preparation and upconversion luminescence of
graphene quantum dots. Chem Commun 2011; 47(9): 2580–2582.
3. Liu R, Wu D, Feng X, et al. Bottom-up fabrication of photoluminescent graphene
quantum dots with uniform morphology. J Am Chem Soc 2011; 133(39):
15221–15223.
4. Lu J, Yeo PSE, Gan CK, et al. Transforming C60 molecules into graphene quantum dots.
Nat Nanotechnol 2011; 6(4): 247–252.
5. Li Y, Hu Y, Zhao Y, et al. An electrochemical avenue to green-luminescent graphene
quantum dots as potential electron-acceptors for photovoltaics. Adv Mater 2011; 23(6):
776–780.
6. Schnez S, Molitor F, Stampfer C, et al. Observation of excited states in a graphene
quantum dot. Appl Phys Lett 2009; 94(1): 012107.
7. Zhuo S, Shao M, Lee ST. Upconversion and downconversion fluorescent graphene quan-
tum dots: ultrasonic preparation and photocatalysis. Acs Nano 2012; 6(2): 1059–1064.
8. Yan X, Cui X, Li B, et al. Large, solution-processable graphene quantum dots as light
absorbers for photovoltaics. Nano Lett 2010; 10(5): 1869–1873.
9. Li LL, Ji J, Fei R, et al. A facile microwave avenue to electrochemiluminescent two-color
graphene quantum dots. Adv Funct Mater 2012; 22(14): 2971–2979.
10. Yang K, Zhang S, Zhang G, et al. Graphene in mice: ultrahigh in vivo tumor uptake and
efficient photothermal therapy. Nano Lett 2010; 10(9): 3318–3323.
Chapter 3
Haifeng Yin
39
40 H. Yin
2. Plasmons in GQDs
2.1. Plasmons in pristine GQDs
Collective electron oscillations in QDs form surface plasmon resonances,7–9
which exhibit unusual properties in energy, lifetime, optical spectroscopy, and
short-time dynamics. Surface plasmon can be excited by the electromagnetic
field of light. In the case of metallic nanoparticles, where the electrons are
confined in the three dimensions, the electron oscillations induce an electric
field around the nanoparticles that can be much larger than the incident light
one. Surface plasmon resonances are one of the best examples that things are
different at the nanoscale. When the size of a particle is reduced to a few
nanometres, the optical properties are dramatically modified by the appear-
ance of surface plasmons and its resulting behavior is completely different
from the bulk one. Surface plasmons open the possibility to amplify, concen-
trate and manipulate light at the nanoscale, overcoming the diffraction limit
42 H. Yin
Figure 1. The dipole response (optical) of the rectangular GQDs to an impulse excitation
polarized in the X-axis (A) and Y-axis (B) direction. The inset is the schematic diagram of the
GQDs. Fourier transform of the induced charge density for GQD (A) and (B) at the energy
resonance points 5.13 and 3.37 eV, respectively, as shown in the direction of the arrow.20
44 H. Yin
46 H. Yin
48 H. Yin
direction, for N-doped hexagonal GQDs whose side length is 1 nm, there
are greater intensity plasmon resonances throughout the near-infrared spec-
tral region, as shown in Figure 6. The electrons which participate in these
near-infrared plasmon resonances oscillate back and forth between the
center and edge regions of the hexagonal QDs. The formation of near-
infrared plasmon resonance mode depends on nitrogen-doped position and
scale size of the GQD. It is only when the nitrogen-doped location is close
to the edge of the QDs that near-infrared plasmon resonance mode of the
GQD will be formed. For N-doped hexagonal GQDs whose side length is
less than 1 nm, there is no plasmon resonance in the near-infrared spectral
region.
50 H. Yin
Figure 7. Optical absorption of the hexagonal GQD and hexagonal GQD dimers to an
impulse excitation polarized in the X-axis directions.50
Figure 8. Optical absorption of the rectangle GQD and rectangle GQD dimers to an
impulse excitation polarized in the X-axis directions.50
In the low energy spectrum area, induced current intensity of the greater
intensity plasmon resonance mode is also larger.
Thongrattanasiri et al. performed realistic, quantum-mechanical calcula-
tions of the optical response of graphene dimers formed by nanodisks and
nanotriangles, showing a strong sensitivity of the level of enhancement to the
type of carbon edges near the gap region with armchair edges favoring
stronger interactions than zigzag edges.52 Their quantum-mechanical descrip-
tion automatically incorporates non-local effects that are absent in classical
electromagnetic theory, leading to over an order of magnitude higher
52 H. Yin
4. Conclusions
Surface plasmon of the graphene has become one of the world’s research
hotspots due to its novel optical properties. This section discusses the physical
nature of plasmon excitation which is modulated in GQDs and GQD dimers.
Compared with the plasmon in the macroscopic graphene, the plasmon in
GQDs has some different properties due to the effects of the size and the
dimensional confinement. In lower-energy resonance zone, the spectral band
is greatly broadened, and the photoabsorption strength line splits. Because of
the electromagnetic coupling between the QDs, GQD dimers exhibit differ-
ent optical properties. For plasmon regulation and control, these results pro-
vide a solid theoretical guidance.
References
1. Grigorenko AN, Polini M, Novoselov KS. Graphene plasmonics. Nat Photon 2012; 6:
749–758.
54 H. Yin
23. Wang W, Li BH, Stassen E, Mortensen NA, Christensen J. Localized surface plasmons in
vibrating graphene nanodisks. Nanoscale 2016; 8: 3809–3815.
24. García de Abajo FJ. Graphene plasmonics: Challenges and opportunities. ACS Photon
2014; 1: 135–152.
25. Cox JD, García de Abajo FJ. Plasmon-enhanced nonlinear wave mixing in nanostructured
graphene. ACS Photon 2015; 2: 306−312.
26. Manjavacas A, Marchesin F, Thongrattanasiri S, Koval P, Nordlander P, Sánchez-Portal D,
García de Abajo FJ. Tunable molecular plasmons in polycyclic aromatic hydrocarbons.
ACS Nano 2013; 7: 3635−3643.
27. Manzoni MT, Silveiro I, García de Abajo FJ, Chang DE. Second-order quantum nonlin-
ear optical processes in graphene nanostructures. arXiv:1406.4360.
28. Cox JD, García de Abajo FJ. Electrically tunable nonlinear plasmonics in graphene nanois-
lands. Nat Commun 2014; 5: 5725–5732.
29. Schiros T, Nordlund D, Pálová L, Prezzi D, Zhao L, Kim KS, Wurstbauer U, Gutiérrez
C, Delongchamp D, Jaye C, Fischer D, Ogasawara H, Pettersson L, Reichman DR, Kim
P, Hybertsen MS, Pasupathy AN. Connecting dopant bond type with electronic structure
in N-doped graphene. Nano Lett 2012; 12: 4025–4031.
30. Wang Y, Shao Y, Matson DW, Li J, Lin Y. Nitrogen-doped graphene and its application
in electrochemical biosensing. ACS Nano 2010; 4: 1790–1798.
31. Lee SU, Belosludov RV, Mizuseki H, Kawazoe Y. Designing nanogadgetry for nanoelec-
tronic devices with nitrogen-doped capped carbon nanotubes. Small 2009; 5:
1769–1775.
32. Zhao L, He R, Rim KT, Rim KT, Kim KS, Zhou H, Gutierrez C, Chockalingam S,
Arguello C, Palova L, Reichman DR, Heinz TF, Kim P, Pinczuk A, Flynn GW,
Pasupathy AN. Visualizing individual nitrogen dopants in monolayer graphene. Science
2011; 333: 999–1003.
33. Gao H, Song L, Guo W, Huang L, Yang D, Wang F, Zuo Y, Fan X, Liu Z, Gao W, Vajtai
R, Hackenberg K, Ajayan PM. A simple method to synthesize continuous large area
nitrogen-doped graphene. Carbon 2012; 50: 4476–4482.
34. Li M, Wu W, Ren W, Cheng H, Tang N, Zhong W, Du YW. Synthesis and upconversion
luminescence of n-doped graphene quantum dots. Appl Phys Lett 2012; 101: 103107.
35. Guo B, Liu Q, Chen E, Zhu H, Fang L, Gong JR. Controllable N-doping of graphene.
Nano Lett 2010; 10: 4975–4980.
36. Xiang HJ, Huang B, Li ZY, Wei SH, Yang JL, Gong XG. Ordered semiconducting nitrogen-
graphene alloys. Phys Rev X 2012; 2: 011003.
37. Ritter KA, Lyding JW. The influence of edge structure on the electronic properties of
graphene quantum dots and nanoribbons. Nat Mater 2009; 8: 235–242.
38. Yan X, Cui X, Li B, Li L. Large solution-processable graphene quantum dots as light
absorbers for photovoltaics. Nano Lett 2010; 10: 1869–1873.
39. Yan X, Cui X, Li L. Synthesis of large, stable colloidal graphene quantum dots with tun-
able size .J Am Chem Soc 2010; 132: 5944–5945.
40. Li Y, Zhao Y, Cheng H, Hu Y, Shi G, Dai L, Qu L. Nitrogen-doped graphene quantum
dots with oxygen-rich functional groups. J Am Chem Soc 2012; 134: 15–18.
41. Jin SH, Kim DH, Jun GH, Hong SH, Jeon S. Tuning the photoluminescence of graphene
quantum dots through the charge transfer effect of functional groups. ACS Nano 2013;
7: 1239–1245.
42. Yin HF, Chen G, Xiang G, Zhang H. Plasmon excitation in nitrogen-doped graphene
nanostructures. Chin J Lumin 2014; 35: 1297–1306.
43. Yin HF, Zhang H,Yue L. Near-infrared plasmon study on N-doped hexagonal graphene
nanostructures. Acta Phys-Chim Sin 2014; 30: 1049–1054.
44. Zuloaga J, Prodan E, Nordlander P. Quantum description of the plasmon resonances of
a nanoparticle dimer. Nano Lett 2009; 9: 887–891.
45. Lim DK, Jeon KS, Kim HM, Nam JM, Suh YD. Nanogap-engineerable Raman-active
nanodumbbells for single-molecule detection. Nat Mater 2010; 9: 60–67.
46. Kim S, Jin J, Kim YJ, Park IY, Kim Y, Kim SW. High-harmonic generation by resonant
plasmon field enhancement. Nat 2008; 453: 757–760.
47. Giannini V, Fernández-Domínguez AI, Heck SC, Maier SA. Plasmonic nanoantennas:
fundamentals and their use in controlling the radiative properties of nanoemitters. Chem
Rev 2011; 111: 3888–3912.
48. Liu N, Tang ML, Hentschel M, Giessen H, Alivisatos AP. Nanoantenna-enhanced gas
sensing in a single tailored nanofocus. Nat Mater 2011; 10: 631–637.
49. Halas NJ, Lal S, Chang WS, Link S, Nordlander P. Plasmons in strongly coupled metallic
nanostructures. Chem Rev 2011; 111: 3913–3961.
50. Yin HF, Zhang H, Liu D. Plasmon excitation in graphene quantum dot dimers. Journal
of Sichuan Normal University (Natural Science) 2014; 37: 691–696.
51. Yin HF, Zeng CH, Yue L. Plasmon excitation in graphene quantum dots connected by
carbon atomic chains. Journal of University of Jinan (Science and Technology) 2016; 30: 1–6.
52. Thongrattanasiri S, Javier Garcı´a de Abajo F. Optical field enhancement by strong plas-
mon interaction in graphene nanostructures. Phys Rev Lett 2013; 110: 187401.
53. Rosolen G, Maes B. Asymmetric and connected graphene dimers for a tunable plasmonic
response. Phys Rev B 2015; 92: 205405.
Chapter 4
1. Introduction
Graphene quantum dots (GQDs) have attracted extensive attention of scien-
tists owing to their excellent chemical/physical qualities arising from the
pronounced quantum confinement and edge effects. In contrast to other
semiconductor quantum dots and organic dyes, GQDs show low cytotoxicity,
good biocompatibility, stable photoluminescence (PL), high electron mobil-
ity, better surface grafting, and so on, thus making them promising in opto-
electronic devices, sensors, bioimaging, etc.1–3 At present, applications of
57
Figure 1. Schematic representation of the ECL aptamer ATP sensor. Reprinted with
permission from Ref. [23]. Copyright © 2013 Elsevier B.V.
Figure 3. Schematic illustration of the preparation processes of ECL electrode and the detec-
tion procedures of the ECL sensor. Reprinted with permission from Ref. [33]. Copyright ©
2015 Elsevier Ltd.
2.4. Antigen
Certain antigens such as specific antigen and carcinoembryonic antigen (CEA)
have been proved to be the indicator of diseases. Accurate analysis of such
antigen plays an important role in early detection and therapy monitoring.
Heart attack (myocardial infarction) is one of the major reasons of death
among people suffering from cardiovascular diseases. Cardiac Troponin I
(cTnI) marker has been considered as the “gold standard” for determining
heart attack in patients. Thus, cTnI level in blood serum is directly associated
with death risk from heart attack. Bhatnagar et al.39 reported a simple, rapid
and sensitive FRET — based immunosensor for detection of cTnI using
GQDs and Gr. The mechanism of immunosensing was shown in Figure 5. The
anti-cTnI/af-GQDs nanoprobe was constructed with monoclonal antibody
cardiac Troponin I (anti-cTnI) and amine-functionalized graphene quantum
dots (af-GQDs) by covalent conjugation which absorbed UV radiations
(360 nm) and emitted blue radiations (437 nm). The fluorescence intensity
of the nanoprobe was quenched after the adsorption of anti-cTnI/af-GQDs
on the surface of Gr through electrostatic attraction and π–π stacking interac-
tion. Finally, the addition of target antigen (cTnI) resulted in fluorescence
recovery. The reason is that due to the specific antibody–antigen interaction,
the detachment of π–π interaction between GQDs conjugate and Gr sheet
occurs and then hinders the FRET process. The good linear relationship
between PL intensity (I) and logarithm of cTnI concentrations was obtained
in the range of 0.001–1000 ng/mL with a limit of detection of 0.192 pg/
mL. Therefore, the proposed immunosensor based on GQDs was founded to
apply in the early stage of myocardial infarction causing heart attack.
Zou et al.40 designed a novel dual-mode immunoassay to detect a tuber-
culosis (TB) antigen, CFP-10. They produced one-dimensionally aligned
magnetoplasmonic (MagPlas) nanowires (NWs), i.e. Fe3O4@PEI@Au nano-
particles as sensing platform which were aligned on the surfaces of silicon
wafers under the external magnetic field. Then GBP-anti CPF10 [G2] and
GBP-anti CPF10 [G3] antibodies were fixed separately on the surfaces of the
MagPlas NPs and GQDs. The target antigen (CFP10) bound with both anti-
bodies via the antibody−antigen interaction to form sandwich-type nanocom-
plex. The nanocomplex with specific antigens sent out stronger SERS signals
than those with no specific antigen. The proposed method was expected to
be further applied in high-throughput screening of target molecules.
Prostate-specific antigen (PSA) was detected with a fluorescent turn-on
nanoprobe by Pei et al.41 The GQDs@Ag core-shell nanocrystals were synthe-
sized and attached with detection anti-PSA antibodies (Ab2) to form the
immunosensing probe (GQDs@Ag–Ab2). Capture anti-PSA antibodies (Ab1)
were directly adsorbed onto magnetic beads (MBs) to form capture probe
(MBs–Ab1). A sandwich type of immunocomplex with PSA, MBs–Ab1 and
GQDs@Ag–Ab2 was constructed and then separated using an external mag-
net. Finally, when Ag shell was etched using H2O2, the incorporated GQDs
were released and the fluorescence of the GQDs was recovered.
Zhang et al.42 developed a novel strategy for ultrasensitive ECL
sandwich-type immunosensor based on the green-luminescence N-GQDs
oxidation using AuNFs modified paper as working electrode (Au-PWE) to
detect α-fetoprotein antigen (AFP). In this work, an interconnected gold
nanoflowers (AuNFs) layer was grown on the surfaces of the paper in order
to fabricate paper working electrode (PWE) and Ab1 was immobilized on
the Au-PWE. MWCNTs/N-GQDs/Ab2 labels were prepared by the
method that N-GQDs were absorbed on the surface of MWCNTs and then
Ab2 was immobilized on the MWCNTs/GQDs composites. Only when
adding of AFP, a strong anodic ECL signal could be observed due to sand-
wich immune complex formation through immunoreaction among Ab1,
AFP and Ab2. Li et al.43 fabricated a highly sensitive three-dimensional (3D)
origami device combined with ECL immunosensor for sensitive point-of-
care testing of CEA. In this sensing system, nanoporous gold/chitosan
modified PWE was used as sensor platform and GQDs functionalized Au@
Pt core–shell nanoparticles (GQDs/Au@Pt) were designed as signal labels.
The role of GQDs, gold nanoflowers and gold nanoparticles employed in
above two types of sensors was to amplify ECL signal and further improve
the sensitivity.
2.5. Trypsin
Trypsin, which is highly correlated with the occurrence inflammation and
related to the onset of pancreatitis, is a pancreatic digestive enzyme. Thus, it
is very important for the clinical detection of trypsin in urine samples for self-
checking by the suspect patient. Poon et al.44 selected GQDs as signal donor
and coumarin fluorophore as an acceptor for the detection of trypsin urine.
BSA was chosen for the linking of the donor and acceptor. The florescence
quenched as a result of FRET when the donor and acceptor were in close
proximity after being linked. While in the presence of trypsin, BSA was
cleaved into small fragments and the florescence recovered. On the same
principle, Li et al.45 demonstrated a novel fluorescent biosensor for trypsin.
GQDs were self-assembled in the presence of cytochrome c (Cyt c) and then
the florescence quenched. However, trypsin cleaved the peptide bonds of Cyt
c to give lysine and arginine residues, both of them induced a subsequent
increase in GQDs florescence intensity.
2.6. Dopamine
Dopamine (DA) is a neurotransmitter that plays an essential role in the func-
tion of the central nervous and cardiovascular systems. The dysfunctions of
DA are related to several important nervous system diseases, such as
Parkinson’s disease, schizophrenia and anorexia. Facile and accurate analysis
of DA could be good for clinical diagnosis of special neurological diseases.
GODs were used mainly as fluorescent probe with the same turn-off
fluorescent strategy for the analysis of DA. For example, DA46 could be
adsorbed on the surface of GQDs and became pDA (polydopamine) through
the self-polymerization effect of DA in alkaline condition, which resulted in
an efficient fluorescence quenching of GQDs. In the same way,
Figure 6. Schematic illustration of the assay strategy for the detection of acetylcholinesterase
activity and inhibitor screening. Reprinted with permission from Ref. [59]. Copyright © 2015
Elsevier B.V.
4. Conclusion
GQDs’ exceptional optical and electronic properties make them powerful
tools in the construction of sensing system, such as optical sensor and elec-
trochemical sensor. At the same time, their optoelectronics properties make
them excellent nanomaterials for a very wide variety of applications.64 PEC
sensors and ECL sensors based on GQDs have been developed gradually.
We have discussed that these nanoparticles have great potential in the
development of analytical sensors and biosensors. However, there are many
problems that have not been resolved during the synthesis of GQDs, such as
high-quality GQDs, appropriate control of size distribution uniformity, and
few ways of optical and electric properties tuning, and so on. Moreover, some
sensing mechanisms of sensors and biosensors constructed on the basis of
GQDs, their derivates or combination with other materials are not fully
understood.
In the fields of medical and pharmaceutical science, GQDs have been
used in the detection of biomolecule (e.g. DNA, antigen and enzyme, etc.)
and drug for clinical diagnosis and molecular screening of drug. We expected
the application is not only for sensing and GQDs would have more develop-
ment by virtue of advantages comparing with other QDs. They should be
further applied in studying the mechanism of drug effect, therapeutical effect,
pathogenesis and so on.
Although optical sensor based on GQDs developed rapidly, electrochemi-
cal sensor still needed to be improved. In addition, optoelectronics conver-
sion capacity is very important for GQDs research, so more studies in
optoelectronics conversion properties would be needed to extend their appli-
cation in nanomedicine fields. There is still much work to explore and
research the application of GQDs thoroughly.
References
1. Shen J, Zhu Y, Yang X, Li C. Graphene quantum dots: emergent nanolights for bioimag-
ing, sensors, catalysis and photovoltaic devices. Chem Comm 2012; 48: 3686–3699.
2. Zhang Z, Zhang J, Chen N, Qu L. Graphene quantum dots: An emerging material for
energy-related applications and beyond. Energy Environ Sci. 2012; 5: 8869–8890.
3. Li L, Wu G, Yang G, Peng J, Zhao J, Zhu J. Focusing on luminescent graphene quantum
dots: Current status and future perspectives. Nanoscale 2013; 5: 4015–4039.
4. Ting SL, Ee SJ, Ananthanarayanan A, Leong KC, Chen P. Graphene quantum dots func-
tionalized gold nanoparticles for sensitive electrochemical detection of heavy metal ions.
Electrochim Acta 2015; 172: 7–11.
5. Huang H, Liao L, Xu X, Zou M, Liu F, Li N. The electron-transfer based interaction
between transition metal ions and photoluminescent graphene quantum dots (GQDs):
A platform for metal ion sensing. Talanta 2013; 117: 152–157.
6. Fan Z, Li Y, Li X, Fan L, Zhou S, Fang D, Yang S. Surrounding media sensitive photo-
luminescence of boron-doped graphene quantum dots for highly fluorescent dyed crys-
tals, chemical sensing and bioimaging. Carbon 2014; 70: 149–156.
7. Amjadi M, Shokri R, Hallaj T. A new turn-off fluorescence probe based on graphene
quantum dots for detection of Au (III) ion. Spectrochim Acta 2016; 153: 619–624.
8. Dong Y, Li G, Zhou N, Wang R, Chi Y, Chen G. Graphene quantum dot as a green and
facile sensor for free chlorine in drinking water. Anal Chem 2012; 84: 8378–8382.
9. Wang L, Zheng J, Yang S, Wu C, Liu C, Xiao Y, Li Y, Qing Z, Yang R. Two-photon sensing
and imaging of endogenous biological cyanide in plant tissues using graphene quantum
dot/gold nanoparticle conjugate. Appl Mater Interfaces 2015; 7: 19509–19515.
10. Chen Y, Dong Y, Wu H, Chen C, Chi Y, Chen G. Electrochemiluminescence sensor for
hexavalent chromium based on the graphene quantum dots/peroxodisulfate system.
Electrochim Acta 2015; 151: 552–557.
11. Punrat E, Maksuk C, Chuanuwatanakul S, Wonsawat W, Chailapakul O. Polyaniline/
graphene quantum dot-modified screen-printed carbon electrode for the rapid determina-
tion of Cr (VI) using stopped-flow analysis coupled with voltammetric technique. Talanta
2016; 150: 198–205.
12. Xu L, Mao W, Huang J, Li S, Huang K, Li M, Xia J, Chen Q. Economical, green route
to highly fluorescence intensity carbon materials based on ligninsulfonate/graphene
quantum dots composites: Application as excellent fluorescent sensing platform for detec-
tion of Fe3+ ions. Sensor ctuat B-Chem 2016; 230: 54–60.
13. Zhang C, Cui Y, Song L, Liu X, Hu Z. Microwave assisted one-pot synthesis of graphene
quantum dots as highly sensitive fluorescent probes for detection of iron ions and pH
value. Talanta 2016; 150: 54–60.
14. Sun H, Gao N, Wu L, Ren J, Wei W, Qu X. Highly photoluminescent amino-
functionalized graphene quantum dots used for sensing copper ions. Chem Eur J 2013;
19: 13362–13368.
15. Liu M, Liu T, Li Y, Xu H, Zheng B, Wang D, Du J, Xiao D. A FRET chemsensor based
on graphene quantum dots for detecting and intracellular imaging of Hg2+. Talanta 2015;
143: 442–449.
16. Chakraborti H, Sinha S, Ghosh S, Pal SK. Interfacing water soluble nanomaterials with
fluorescence chemosensing: Graphene quantum dot to detect Hg2+ in 100% aqueous
solution. Mater Lett 2013; 97: 78–80.
17. Shi B, Zhang L, Lan C, Zhao J, Su Y, Zhao S. One-pot green synthesis of oxygen-
rich nitrogen-doped grapheme quantum dots and their potential application in pH-sensi-
tive photoluminescence and detection of mercury (II) ions. Talanta 2015; 142:
131–139.
18. Hua M, Wang C, Qian J, Wang K, Yang Z, Liu Q, Mao H, Wang K. Preparation of gra-
phene quantum dots based core-satellite hybrid spheres and their use as the ratiometric
fluorescence probe for visual determination of mercury (II) ions. Anal Chim Acta 2015;
888: 173–181.
19. Qi Y, Zhang M, Fu Q, Liu R, Shi G. Highly sensitive and selective fluorescent detection
of cerebral lead (II) based on graphene quantum dot conjugates. Chem Commun 2013;
49: 10599–10601.
20. Park M, Ha HD, Kim YT, Jung JH, Kim SH, Kim DH, Seo TS. Combination of a sample
pretreatment microfluidic device with a photoluminescent graphene oxide quantum dot
sensor for trace lead detection. Anal Chem 2015; 87: 10969–10975.
21. Qian Z, Shan X, Chai L, Chen J, Feng H. A fluorescent nanosensor based on graphene
quantum dots–aptamer probe and graphene oxide platform for detection of lead (II) ion.
Biosens Bioelectron 2015; 68: 225–231.
22. Liu Z, Xiao J, Wu X, Lina L, Weng S, Chen M, Cai X, Lin X. Switch-on florescent strategy
based on N and S co-doped graphene quantum dots (N-S/GQDs) for monitoring
pyrophosphate ions in synovial fluid of arthritis patients. Sensor Actuat B-Chem 2016;
229: 217–224
23. Lu J, Yan M, Ge L, Ge S, Wang S, Yan J, Yu J. Electrochemiluminescence of blue-
luminescent graphene quantum dots and its application in ultrasensitive aptasensor for
adenosine triphosphate detection. Biosensor Bioelectron 2013; 47: 271–277.
24. Liu J, Zhang X, Cong Z, Chen Z, Yang H, Chen G. Glutathione-functionalized graphene
quantum dots as selective fluorescent probes for phosphate-containing metabolites.
Nanoscale 2013; 5: 1810–1815.
25. Wu Z, Li W, Chen J, Yu C. A graphene quantum dot-based method for the highly sensitive
and selective fluorescence turn on detection of biothiols. Talanta 2014; 119: 538–543.
26. Liu Z, Gong Y, Fan Z. Cysteine detection using a high-fluorescence sensor based on a
nitrogen-doped graphene quantum dot–mercury (II) system. J Lumin 2016; 175:
129–134.
27. Lin L, Song X, Chen Y, Rong M, Wang Y, Zhao L, Zhao T, Chen X. Europium-decorated
graphene quantum dots as a fluorescent probe for label-free, rapid and sensitive detection
of Cu2+ and L-cysteine. Anal Chim Acta 2015; 891: 261–268.
28. Zhang Q, Song C, Zhao T, Fu H, Wang H, Wang Y, Kong D. Photoluminescent sensing
for acidic amino acids based on the disruption of graphene quantum dots/europium ions
aggregates. Biosensor Bioelectron, 2015; 65: 204–210.
29. Achadu OJ, Uddin I, Nyokong T. The interaction between graphene quantum dots
grafted with polyethyleneimine and Au@Ag nanoparticles: Application as a fluorescence
“turn-on” nanoprobe. J Photoch Photo A 2016; 324: 96–105.
30. Jua J, Zhang R, Chen W. Photochemical deposition of surface-clean silver nanoparticles
on nitrogen-doped graphene quantum dots for sensitive colorimetric detection of
glutathione. Sensor Actuat B-Chemical 2016; 228: 66–73.
31. Wang L, Tricard S, Yue P, Zhao J, Fang J, Shen W. Polypyrrole and graphene quantum
dots @ Prussian Blue hybrid film on graphite felt electrodes: Application for amperometric
determination of L-cysteine. Biosensor Bioelectron 2016; 77: 1112–1118.
47. Zhao J, Zhao L, Lan C, Zhao S. Graphene quantum dots as effective probes for label-free fluo-
rescence detection of dopamine. Sensors Actuators B: Chem 2016; 223: 246–251.
48. Zhou X, Ma P, Wang A, Yu C, Qian T, Wu S, Shen J. Dopamine fluorescent sensors based
on polypyrrole/grapheme quantum dots core/shell hybrids. Biosensors Bioelectron 2015;
64: 404–410.
49. Li Y, Jiang Y, Mo T, Zhou H, Li Y, Li S. Highly selective dopamine sensor based on
graphene quantum dots self-assembled monolayers modified electrode. J Electroanal
Chem 2016; 767: 84–90.
50. Zhang Y, Wu C, Zhou X, Wu X, Yang Y, Wu H, Guo S, Zhang J. Graphene quantum
dots/gold electrode and its application in living cell H2O2 detection. Nanoscale 2013; 5:
1816–1819.
51. Wu X, Tian F, Wang W, Chen J, Wu M, Zhao JX. Fabrication of highly fluorescent
graphene quantum dots using L-glutamic acid for in vitro/in vivo imaging and sensing.
J Mater Chem C 2013; 1: 4676–4684.
52. Zheng A, Cong Z, Wang J, Li J, Yang H, Chen G. Highly-efficient peroxidase-like catalytic
activity of grapheme dots for biosensing. Biosensor Bioelectron 2013; 49: 519–524.
53. Lin L, Song X, Chen Y, Rong M, Zhao T, Wang Y, Jiang Y, Chen X. Intrinsic peroxidase-
like catalytic activity of nitrogen-doped grapheme quantum dots and their application in
the colorimetric detection of H2O2 and glucose. Anal Chim Acta 2015; 869: 89–95.
54. Nirala NR, Abraham S, Kumar V, Bansal A, Srivastava A, Saxena PS. Colorimetric detec-
tion of cholesterol based on highly efficient peroxidase mimetic activity of graphene
quantum dots. Sensor Actuat B: Chem 2015; 218: 42–50.
55. Amjadi M, Manzoori JL, Hallaj T. Chemiluminescence of graphene quantum dots and its
application to the determination of uric acid. J Lumin 2014; 153: 73–78.
56. Wang Y, Zhang L, Liang R, Bai J, Qiu J. Using graphene quantum dots as photolumines-
cent probes for protein kinase sensing. Analy Chem 2013; 85: 9148–9155.
57. Liang R, Qiu W, Zhao H, Xiang C, Qiu J. Electrochemiluminescence resonance energy
transfer between graphene quantum dots and graphene oxide for sensitive protein kinase
activity and inhibitor sensing. Anal Chim Acta 2016; 904: 58–64.
58. Liu j, He X, Wang K, He D, Wang Y, Mao Y, Shi H, Wen L. A highly sensitive electro-
chemiluminescence assay for protein kinase based on double-quenching of graphene
quantum dots by G-quadruplex–hemin and gold nanoparticles. Biosensor Bioelectron
2015; 70: 54–60.
59. He T, Qi L, Zhang J, Huang Y, Zhang Z. Enhanced graphene quantum dot fluorescence
nanosensor for highly sensitive acetylcholinesterase assay and inhibitor screening. Sensor
Actuat B: Chem 2015; 215: 24–29.
60. Liu Y, Yan K, Okoth OK, Zhang J. A label-free photoelectrochemical aptasensor based on
nitrogen-doped graphene quantum dots for chloramphenicol determination. Biosensor
Bioelectron 2015; 74: 1016–1021.
61. Zhao H, Chang Y, Liu M, Gao S, Yu H, Quan X. A universal immunosensing strategy
based on regulation of the interaction between graphene and graphene quantum dots.
Chem Comm 2013; 49: 234–236.
62. Patra S, Roy E, Choudhary R, Tiwari A, Madhuri R, Sharma PK. Graphene quantum dots
decorated CdS doped graphene oxide sheets in dual action mode: As initiator and plat-
form for designing of nimesulide imprinted polymer. Biosens Bioelectron 2015; 89:
627–635.
Chapter 5
Due to quantum size confinement and edge effects, graphene quantum dots (GQDs)
have presented extraordinary properties, attracting extensive attention from scientists
in the field. This chapter focuses in the recent progress made in GQDs nanotechnology
as applied to food analysis. Typical examples are presented for detected objectives
(e.g. microorganisms, pesticides, toxins, or heavy ions) and different analysis methods
(optical, electrochemical, and chromatographic) in food analysis.
77
2.2. Pesticides
Pesticides are compounds collection of insecticides, fungicides, herbicides,
molluscicides, rodenticides, plant growth regulators and repellents,8 which
contribute significantly to the contamination of the food chain. Due to the
high water solubility of most pesticides, their action mechanism, their high
level of toxicity and the widespread use in agriculture, there is a critical need
for highly sensitive and selective analytical methods for residue analysis of
these pollutants. Zor et al. reported on the synthesis of a multifunctional
composite material based on GODs, magnetic silica beads, and molecularly
imprinted polypyrrole, which enables a rapid, simple and sensitive platform
for pesticide tributyltin (as a model) detection, even in complex mediums
such as seawater without any sample treatment.37
2.3. Toxins
Toxins cover a large variety of molecules and can be grouped as plant/bacte-
rial toxins, mycotoxins and marine toxins. The consequences of consuming
food contaminated with toxins range from mild gastrointestinal problems to
death depending on the type of toxin and its concentration. Mycotoxins are
secondary metabolites naturally produced by fungal species that grow on
nuts, fruits and grains, both in the field, after harvest and during storage.
Their presence in food supply is due to direct contamination of plants or their
products or by contamination of animal feeds.4 Toxin detection methods fea-
tured with short analytical time, low cost, high sensitivity and easy adaptability
for in situ measurement are most wanted. Wang et al. developed a two-photo
excitation nanosensor using GQDs/AuNPs conjugate for sensing and imag-
ing endogenous biological CN−. With the benefit of the high quenching
efficiency of AuNPs and excellent two-photon properties of GQDs, the sens-
ing system can achieve a low detection limit and deeper penetration depth
without interference from background signals of a complex biological envi-
ronment, thus realizing sensing and imaging of CN− in different types of plant
tissues and even monitoring CN− removal in food processing.38
Figure 2. (A) Schematic illustration of AuNP–PEP–GQDs nanosensor for CN− assay; (B)
Two-photon excitation images of (A) the fresh cassava cell, (B) cassava cell pretreated by soak-
ing and washing, and (C) 200 mM CN− preincubated cassava cell after incubation with AuNP–
PEP–GQDs (2 nM) for 2 h. Scale bar: 10 mm. Reused with permission from Ref. [38].
American Chemical Society.
FL Intensity
0 600 0.8 1.04
Fluorescence Intensity
1
400
600 .
2
.. 200
0.6
9.24 k = 0.936
600 0.4
FL Intensity
0 b = –0.0155
0 2 4 6 8 10 12
0.2 R = 0.9989
12 Time (min) –4
LOD = 8.4 × 10 M
400 0
400 2 4 6 8 10
4
Concentration (×10 M)
200 200
0 0
400 450 500 550 600 400 450 500 550 600
Wavelength (nm) Wavelength (nm)
(A) (B)
Figure 3. (A) Fluorescence response of GQDs in the presence of 9.6 mM hydroquinone,
6.0 mg L-1 HRP and 1.04 × 10-4 M H2O2 in pH 7.4 PBS buffer at different times.
Inset: (1) time-dependent fluorescence changes of GQDs with hydroquinone at 452 nm;
(2) fluorescence and GQDs in the absence and presence of hydroquinone; (B) The fluores-
cence quenching effect of hydroquinone at different concentrations in the presence of
HRP and H2O2 (reaction time: 8 min). Inset: fluorescence quenching effect vs. concentra-
tion of added hydroquinone. Reused with permission from Ref. [47]. The Royal Society
of Chemistry.
Figure 6. The schematic illustration of the BPA detection principle. Reused with permission
from Ref. [40]. The Royal Society of Chemistry.
Figure 8. The sensing mechanism of the proposed GQDs–AuNPs FRET biosensor for
S. aureus gene detection. Reused with permission from Ref. [36]. Elsevier B.V.
Figure 9. Schematic illustration of the ECL biosensor based on GQDs combined with endo-
nuclease cleavage and bidentate chelation. Reused with permission from Ref. [50]. American
Chemical Society.
Figure 10. (A) DPVs of the GQD-modified GCE in PBS w/o ascorbic acid and acetami-
nophen; (B) CVs of the GQD-modified GCE in PBS (pH = 6.5) with ascorbic acid and aceta-
minophen at various scan rates. Reused with permission from Ref. [34]. The Royal Society of
Chemistry.
Figure 11. Schematic drawing of the synthesis of GQDs from pyrolysis citric acid and elec-
trochemical oxidize HQ and CC on GQDs/GCE. Reused with permission from Ref. [35].
Elsevier Ltd.
Figure 12. Chromatograms of (A) fresh mineral water; (B) mineral water kept in plastic bot-
tles and directly exposed to sunlight for one week; (C) mineral water kept in plastic bottles and
directly exposed to sunlight for one week and spiked with 20 ng mL-1 BPA. Reused with
permission from Ref. [39]. The Royal Society of Chemistry.
Figure 13. Effect of temperature on the chromatograms on the GQD coated capillary
column (22 m long × 0.32 mm i.d.) for the separation of: (A) ethylbenzene and styrene;
(B) ethylbenzene, p-Xylene and o-Xylene; (C) propylbenzene isomers; (D) dichlorobenzene
isomers. Reused with Permission from Ref. [41]. The Royal Society of Chemistry.
References
1. Picó Y. Chemical Analysis of Foods: Techniques and Applications. 1st Edn. Academic Press
is an imprint of Elsevier: 225 Wyman Street, Waltham, MA 02451, USA 2012.
2. Cifuentes A. Food analysis: Present, future, and foodomics. ISRN Anal Chem 2012;
2012: Article ID 801607.
3. Nielsen SS. Food Analysis. In Food Analysis, 4th Edn., Nielsen SS (Ed), Springer:
New York, 2010: p. 6.
4. Valdés MG, Valdés González AC, García Calzón JA, Díaz-García ME. Analytical nano-
technology for food analysis. Microchim Acta 2009; 166: 1–19.
5. Blasco C, Picó Y. Determining nanomaterials in food. TrAC Trend Anal Chem 2011; 30:
84–99.
6. Pérez-López B, Merkoçi A. Nanomaterials based biosensors for food analysis applications.
Trends Food Sci Tech 2011; 22: 625–639.
7. Stephen Inbaraj B, Chen BH. Nanomaterial-based sensors for detection of foodborne
bacterial pathogens and toxins as well as pork adulteration in meat products. J Food Drug
Anal 2016; 24: 15–28.
8. McGrath TF, Elliott CT, Fodey TL. Biosensors for the analysis of microbiological and
chemical contaminants in food. Anal Bioanal Chem 2012; 403: 75–92.
9. Sharma R, Ragavan KV, Thakur MS, Raghavarao KS. Recent advances in nanoparticle
based aptasensors for food contaminants. Biosens Bioelectron 2015; 74: 612–627.
10. Wang Y, Qu K, Tang L, Li Z, Moore E, Zeng X, Liu Y, Li J. Nanomaterials in carbohy-
drate biosensors. TrAC Trends Anal Chem 2014; 58: 54–70.
11. Cui L, Wu J, Ju H. Electrochemical sensing of heavy metal ions with inorganic, organic
and bio-materials. Biosens Bioelectron 2015; 63: 276–286.
12. Eltzov E, Guttel S, Low Yuen Kei A, Sinawang PD, Ionescu RE, Marks RS. Lateral flow
immunoassays — from paper strip to smartphone technology. Electroanalysis 2015; 27:
2116–2130.
13. Huang X, Aguilar ZP, Xu H, Lai W, Xiong Y. Membrane-based lateral flow immunochro-
matographic strip with nanoparticles as reporters for detection: A review. Biosens
Bioelectron 2016; 75: 166–180.
14. Syed MA. Advances in nanodiagnostic techniques for microbial agents. Biosens Bioelectron
2014; 51: 391–400.
15. Wang X, Niessner R, Tang D, Knopp D. Nanoparticle-based immunosensors and immu-
noassays for aflatoxins. Anal Chim Acta 2016; 912: 10–23.
16. Jiang C, Wu H, Song X, Ma X, Wang J, Tan M. Presence of photoluminescent carbon
dots in Nescafe(R) original instant coffee: Applications to bioimaging. Talanta 2014;
127: 68–74.
17. Gao Z, Wang L, Su R, Huang R, Qi W, He Z. A carbon dot-based “off-on” fluorescent
probe for highly selective and sensitive detection of phytic acid. Biosens Bioelectron 2015;
70: 232–238.
18. Wang B, Chen Y, Wu Y, Weng B, Liu Y, Lu Z, Li CM, Yu C. Aptamer induced assembly
of fluorescent nitrogen-doped carbon dots on gold nanoparticles for sensitive detection of
AFB1. Biosens Bioelectron 2016; 78: 23–30.
19. Lai IP-J, Harroun SG, Chen S-Y, Unnikrishnan B, Li Y-J, Huang C-C. Solid-state syn-
thesis of self-functional carbon quantum dots for detection of bacteria and tumor cells.
Sensor Actuat B Chem 2016; 228: 465–470.
20. Duan N, Gong W, Wang Z, Wu S. An aptasensor based on fluorescence resonance energy
transfer for multiplexed pathogenic bacteria determination. Anal Methods 2016; 8:
1390–1395.
21. Loo AH, Sofer Z, Bousa D, Ulbrich P, Bonanni A, Pumera M. Carboxylic carbon quan-
tum dots as a fluorescent sensing platform for DNA detection. ACS Appl Mater Interfaces
2016; 8: 1951–1957.
22. Xue M, Zhang L, Zhan Z, Zou M, Huang Y, Zhao S. Sulfur and nitrogen binary doped
carbon dots derived from ammonium thiocyanate for selective probing doxycycline in liv-
ing cells and multicolor cell imaging. Talanta 2016; 150: 324–330.
23. Liu J, Chen Y, Wang W, Feng J, Liang M, Ma S, Chen X. “Switch-on” fluorescent sensing
of ascorbic acid in food samples based on carbon quantum dots-MnO2 Probe. J Agr Food
Chem 2016; 64: 371–380.
24. Liu G, Chen Z, Jiang X, Feng D, Zhao J, Fan D, Wang W. In-situ hydrothermal synthesis
of molecularly imprinted polymers coated carbon dots for fluorescent detection of bisphe-
nol A. Sensor Actuat B-chem 2016; 228: 302–307.
25. Hou J, Dong G, Tian Z, Lu J, Wang Q, Ai S, Wang M. A sensitive fluorescent sensor for
selective determination of dichlorvos based on the recovered fluorescence of carbon dots-
Cu(II) system. Food Chem 2016; 202: 81–87.
26. Yue X, Zhu W, Ma S, Yu S, Zhang Y, Wang J, Wang Y, Zhang D, Wang J. Highly sensitive
and selective determination of tertiary butylhydroquinone in edible oils by competitive
reaction induced “on–off–on” fluorescent switch. J Agr Food Chem 2016; 64: 706–713.
27. Xu H, Yang X, Li G, Zhao C, Liao X. Green synthesis of fluorescent carbon dots for selec-
tive detection of tartrazine in food samples. J Agr Food Chem 2015; 63: 6707–6714.
28. Wang Z, Yu J, Gui R, Jin H, Xia Y. Carbon nanomaterials-based electrochemical aptasen-
sors. Biosens Bioelectron 2016; 79: 136–149.
29. Yan Z, Qu X, Niu Q, Tian C, Fan C, Ye B. A green synthesis of highly fluorescent nitro-
gen-doped graphene quantum dots for the highly sensitive and selective detection of
mercury(II) ions and biothiols. Anal Methods 2016; 8: 1565–1571.
30. Zhu S, Song Y, Zhao X, Shao J, Zhang J, Yang B. The photoluminescence mechanism in
carbon dots (graphene quantum dots, carbon nanodots, and polymer dots): Current state
and future perspective. Nano Res 2015; 8: 355–381.
31. Zhang Z, Zhang J, Chen N, Qu L. Graphene quantum dots: An emerging material for
energy-related applications and beyond. Energy Environ Sci 2012; 5: 8869–8890.
32. Shen J, Zhu Y, Yang X, Li C. Graphene quantum dots: Emergent nanolights for bioimag-
ing, sensors, catalysis and photovoltaic devices. Chem Commun 2012; 48: 3686–3699.
33. Guo X, Mei N. Assessment of the toxic potential of graphene family nanomaterials. J Food
Drug Anal 2014; 22: 105–115.
34. Zhao C, Liu Z, Xu W, Chen M, Weng S, Xu L, Cai Q. A glassy carbon electrode based
on graphene quantum dots (GQDs) for simultaneous detection of acetaminophen and
ascorbic acid. Anal Methods 2015; 7: 8877–8881.
35. Jian X, Liu X, Yang H, Guo M, Song X, Dai H, Liang Z. Graphene quantum dots modi-
fied glassy carbon electrode via electrostatic self-assembly strategy and its application.
Electrochim Acta 2016; 190: 455–462.
36. Shi J, Chan C, Pang Y, Ye W, Tian F, Lyu J, Zhang Y, Yang M. A fluorescence resonance
energy transfer (FRET) biosensor based on graphene quantum dots (GQDs) and gold
nanoparticles (AuNPs) for the detection of mecA gene sequence of Staphylococcus aureus.
Biosens Bioelectron 2015; 67: 595–600.
37. Zor E, Morales-Narvaez E, Zamora-Galvez A, Bingol H, Ersoz M, Merkoci A. Graphene
quantum dots-based photoluminescent sensor: A Multifunctional composite for pesticide
detection. ACS Appl Mater Interfaces 2015; 7: 20272–20279.
38. Wang L, Zheng J, Yang S, Wu C, Liu C, Xiao Y, Li Y, Qing Z, Yang R. Two-photon
sensing and imaging of endogenous biological cyanide in plant tissues using graphene
Chapter 6
97
98 S. Li et al.
with respect to in vivo or even in vitro uses. And the blinking characteristics
of semiconductor QDs make single-molecule tracking difficult. In addition,
because of a much larger size than a biomolecule, semiconductor QDs might
affect the dynamics and functions of the target molecules and create artificial
clusters while associating with multiple targets. It is worth mentioning that
GQDs, as a new emerging star in the field of fluorescent carbon nanomateri-
als, are also exciting in this regard.22,23 GQDs are advantageous compared to
fluorescent proteins, organic dyes, and semiconductor QDs due to their
chemical inertness, low cytotoxicity, good biocompatibility as well as superior
resistance to photobleaching and blinking.24,25 In particular, functional
groups such as carboxyl, amino, carbonyl, hydroxyl, and other groups on the
surface of GQDs result in their high hydrophilicity and readiness for further
functionalization with various organic, polymeric, or biological species,26
which expand the range and heighten the sensitivity and selectivity of bioim-
aging. By taking advantage of the unique properties of GQDs coupled with
other merits such as low cost and ease of synthesis, a large number of research
groups have investigated the potential applications of GQDs in cellular imag-
ing.27,28 Moreover, GQDs with up-conversion fluorescence or red fluores-
cence have been successfully synthesized by designing appropriate carbon
precursors or synthetic strategies,29,30 which further expands the applications
of GQDs in in vivo imaging. Following will focus on the recent advances of
GQDs for in vitro and in vivo imaging, respectively, and discuss issues and
prospects in such applications.
100 S. Li et al.
Figure 1. Cellular imaging and cellular toxicity of GQDs. (A)–(C) are washed cells imaged
under bright field, 405 nm and 488 nm excitations, respectively. (D) Effect of GQDs on
MG-63 cells viability. Reprinted with permission from Ref. [31]. Copyright 2011 Royal
Society of Chemistry.
weaken the cell activity significantly, suggesting that GQDs possess low toxic-
ity and could be used in bioimaging at high concentrations. Since then,
GQDs have been used to label a variety of cancer cells such as human cervical
carcinoma cells (HeLa),32–37 human lung cancer cells (A-549),38–41 murine
alveolar macrophage cells (MH-S),42 human hepatic cancer cells (Huh7),43
human breast cancer cells (MCF-7),44 and ovarian cancer cells (A2780).45 All
these studies suggested the low cytotoxicity and excellent biocompatibility of
GQDs, thus they could be used as an eco-friendly material for biolabeling
and bioimaging.
Although GQDs emitting blue to red fluorescence have shown potential
application in cancer cellular imaging, most researches only concentrated on
the uptake of bare GQDs by cancer cells without considering the selectivity
and function of the agents. Targeted cancer cellular imaging, as a fundamen-
tal research, promotes cancer detection and targeted therapeutic modalities
for effective cancer chemotherapy with improved safety. GQDs are generally
composed of sp2- and sp3-hybridized carbon nanostructures in the form of
conjugated carbon clusters functionalized with abundant periphery oxygen-
containing functional groups, imparting them with suitability for subsequent
functionalization with organic, polymeric, inorganic or biological species by
p–p conjugation or chemical modifications.46–53 Therefore, targeted cancer
cellular imaging based on functionalized GQDs has also been developed.
Fluorescent GQDs were prepared and conjugated with folic acid (FA) to
demonstrate highly selective and specific cancer cellular imaging by Wang
et al.53 The FA-conjugated GQDs (GQDs-FA) were incubated with three cell
types that express FA receptor (FR) at different levels to demonstrate that FA
retained its binding affinity to FR after conjugation with GQDs. As shown in
Figure 2, the fluorescence of GQDs-FA in HeLa cells is considerably stronger
than that in A549 and HEK293A (human embryonic kidney cell line) cells,
which expressed FA receptor (FR) at a low level, indicating that GQDs-FA
were internalized via FR-induced endocytosis. The GQDs-FA were then
employed as carriers of the anticancer drug doxorubicin (DOX) for targeted
cell delivery.
The cell nucleus is critical to various important cellular events, including
metabolism, heredity and reproduction.54 Nucleus-staining is the first step to
revealing the nucleus morphology, to investigating the nuclear function, and
to transferring the agent to the cell nucleus.55 In most of the cellular imaging
experiments, GQDs, like other nanomaterials, were located only in the cyto-
plasm. However, single-layer GQDs prepared by refluxing Vulcan CX-72
carbon black with concentrated nitric acid were found to be able to enter the
cell nucleus (Figure 3).56 This was the first time for GQDs to label the cell
nucleus, but the researchers did not come up with a convincing explanation
of GQDs entering the nucleus. Immediately following these findings, Wang
et al. also demonstrated the nucleus labeling property of polyethylenimine
functionalized GQDs (GQDs-PEI) synthesized by a new convenient solvo-
thermal method which combined the “top-down” cutting, reducing and
functionalizing GO sheets in one-step.57 They deduced that the nucleus per-
meability of GQDs-PEI may be attributed to the ultrasmall size of GQDs-
PEI and high concentration of positively charged nitrogen atoms of PEI,
which also exhibits an excellent membrane-disruption ability. Cell nucleus
imaging using GQDs offers a further application in the diagnosis of diseased
phenotypes and the targeted therapy of cancer.
102 S. Li et al.
Figure 2. Confocal laser scanning microscopy images of (A) HeLa, (B) A549 and (C)
HEK293A cells incubated with GQDs-FA (20.0 mg/mL) at 37oC for 10 min. Left column:
bright field images. Middle column: fluorescence image with Hoechst nuclear stained (blue).
Right column: GQDs fluorescence image (green). Reprinted with permission from Ref. [53].
Copyright 2014 Elsevier.
Figure 3. Confocal laser scanning microscope images of MCF-7 cells labeled with GQDs.
(A) Fluorescent image, (B) bright-field image, (C) merged fluorescent and bright-field image,
and (D) section analysis. Reprinted with permission from Ref. [56]. Copyright 2012 Royal
Society of Chemistry.
Although GQDs can penetrate into the cancer cells effortlessly for imaging in
various studies, stem cells labeling still poses a considerable challenge.
Zhang et al. firstly used stabilizer-free GQDs as a fluorescent labeling
agent in long-term stem cellular imaging.60 Bright yellow fluorescent GQDs
with fluorescence quantum yield (QY) of 14% have demonstrated direct and
easy penetration into three different kinds of stem cells, neurospheres cells
(NSCs), pancreas progenitor cells (PPCs) and cardiac progenitor cells (CPCs)
without affecting their viability, proliferation or differentiation capacity. As
shown in Figure 4, the morphology of stem cells could be clearly discerned
after incubated with GQDs. They further investigated the uptake mechanism
and biocompatibilty of such GQDs with human neural stem cells (hNSCs).61
Transmission electron microscopy (TEM) images confirmed that the GQDs
were indeed internalized by the hNSCs via the endocytosis mechanism and
104 S. Li et al.
Figure 4. Confocal fluorescence microscopy images of NSCs (A), PPCs (C) and CPCs (E)
with the yellow fluorescent GQDs incorporated at the excitation wavelength of 405 nm and
corresponding images under bright field (B, D, F). Reprinted with permission from Ref. [60].
Copyright 2012 Royal Society of Chemistry.
were located in the cytoplasm. And GQDs did not affect the self-renewal and
expression of the cell type-specific marker in hNSCs. Interestingly, the fluo-
rescent C60 nanoparticles which could easily penetrate into cancer cells were
not detected inside stem cells62 and CdS QDs were shown to be cytotoxic to
stem cells.63 Therefore, such GQDs have a distinct advantage in terms of their
penetrating power to stem cells.
Cancer stem cells (CSCs), as a subpopulation of stem-like cells within
tumors, exhibit the characteristics of both stem cells and cancer cells.64 They
have the capacity to self-renew inside a tumor and to generate heterogeneous
lineages of cancer cells that drive cancer progression, metastasis and drug
resistance.65,66 CSCs have been proposed as the cause of tumor relapse and
are the relatively new target of cancer therapies, and it is of vital importance
and urgent necessity to successfully label CSCs. Guo et al. synthesized
rhodamine B derivative-functionalized GQDs (RBD-GQDs) which can bind
to Fe3+ forming RBD-GQDs–Fe3+ with strong orange–red fluorescence of
43% QY.67 Such RBD-GQDs–Fe3+ were first demonstrated to be biomarkers
for pancreatic CSCs (Figure 5). Compared with other competing fluorescent
106 S. Li et al.
108 S. Li et al.
Figure 7. (A) The visualization of light penetration through the tissues according to its wave-
length. (B) Energy gap of π–π* transitions calculated based on DFT as a function of the num-
ber of fused aromatic rings. (C) Fluorescence spectra (c1) and high-resolution TEM image (c2)
of RF-GQDs.91 (A) Reprinted with permission from Ref. [83]. Copyright 2011 American
Cancer Society. (B) Reprinted with permission from Ref. [90]. Copyright 2010 John Wiley
and Sons. (C) Reprinted with permission from Ref. [91]. Copyright 2015 Royal Society of
Chemistry.
an sp2 cluster of more than 20 aromatic rings has an energy gap of around
2 eV (Figure 7(B)) which can convert into about 600 nm PL.90 Larger sp2
clusters have a lower energy gap and may emit at the red light region. Red
fluorescent GQDs (RF-GQDs) without any chemical modification were suc-
cessfully prepared via facile electrochemical exfoliation of graphite in K2S2O8
solution by Tan et al. (Figure 7(C)).91 The RF-GQDs were found to consist
of isolated sp2 domains with a diameter of about 3 nm, and the very active
SO4•- radicals produced from S2O82- acted as electrochemical ‘‘scissors’’ to
sharply cut the graphene sheets into small intact sp2 structures, known as the
Figure 8. (A) In vivo fluorescence imaging of mice injected GQDs subcutaneously (spot a)
and intramuscularly (spot b). The images were taken at various excitation wavelengths and
emission wavelengths indicated at the top of each image. Copyright 2013 Royal Society of
Chemistry. (B) Bright-field image (b1) and red-fluorescence image (b2) after subcutaneous
injection of GQDs in different areas. The excitation wavelength was 502–540 nm, and the
collected fluorescence channel was 695–775 nm. (A) Reprinted with permission from Ref.
[92]. (B) Reprinted with permission from Ref. [93]. Copyright 2014 Nature Publishing
Group.
110 S. Li et al.
References
1. Megason SG, Fraser SE. Cell 2007; 130: 784–795.
2. Mahmood U. IEEE Eng Med Biol Mag 2004; 23: 58–66.
3. Pomper MG, Hammoud DA. IEEE Eng Med Biol Mag 2004; 23: 28–37.
4. Hogemann D, Basilion JP. Eur J Nucl Med Mol Imaging 2002; 29: 400–408.
5. Hogemann D, Basilion JP, Weissleder R. Radiologe 2001; 41: 116–120.
6. Bremer C, Ntziachristos V, Mahmood U, Tung CH, Weissleder R. Radiologe 2001; 41:
131–137.
112 S. Li et al.
41. Zhang X, Wang S, Liu M, Yang B, Feng L, Ji Y, Tao L, Wei Y. Phys Chem Chem Phys 2013;
15: 19013–19018.
42. Wu X, Tian F, Wang W, Chen J, Wu M, Zhao JX, J Mater Chem C 2013; 1:
4676–4684.
43. Kumar V, Singh V, Umrao S, Parashar V, Abraham S, Singh AK, Nath G, Saxena PS,
Srivastava A. RSC Adv 2014; 4: 21101–21107.
44. Zhou L, Geng J, Liu B. Part Part Syst Char 2013; 30: 1086–1092.
45. Zhang C, Liu Y, Xiong XQ, Peng LH, Gan L, Chen CF, Xu HB. Org Lett 2012; 14:
5912–5915.
46. Qiu J, Zhang R, Li J, Sang Y, Tang W, Gil PR, Liu H. Int J Nanomed 2015; 10:
6709–6724.
47. Jing Y, Zhu Y, Yang X, Shen J, Li C. Langmuir 2011; 27: 1175–1180.
48. Chen T, Yu H, Yang NW, Wang MD, Ding CD, Fu JJ. J Mater Chem B 2014; 2:
4979–4982.
49. Deng L, Liu L, Zhu C, Li D, Dong S. Chem Commun 2013; 49: 2503–2505.
50. Xue Q, Huang H, Wang L, Chen Z, Wu M, Li Z, Pan D. Nanoscale 2013; 5:
12098–12103.
51. Qian Z, Ma J, Shan X, Shao L, Zhou J, Chen J, Feng H. RSC Adv 2013; 3:
14571–14579.
52. Nigam P, Waghmode S, Louis M, Wangnoo S, Chavan P, Sarkar D. J Mater Chem B
2014; 2: 3190–3195.
53. Wang XJ, Sun X, Lao J, He H, Cheng TT, Wang MQ, Wang SJ, Huang F. Colloid Surface
B 2014; 122: 638–644.
54. Li C, Liu Y, Wu Y, Sun Y, Li F. Biomaterials 2013; 34: 1223–1234.
55. Horobin RW, Stockert JC, Doubell FR. Histochem Cell Biol 2013; 139: 623–637.
56. Dong YQ, Chen CQ, Zheng XT, Gao LL, Cui ZM, Yang HB, Guo CX, Chi YW, Li CM.
J Mater Chem 2012; 22: 8764–8766.
57. Wang H, Wang X. RSC Adv 2015; 5: 75380–75385.
58. Prockop DJ. Science 2001; 293: 211–212.
59. Parker GC, Kristeva MA, Eisenberg LM, Rao MS, Williams MA, Sanberg PR, English D.
Stem Cells Dev 2005; 14: 463–469.
60. Zhang M, Bai LL, Shang WH, Xie WJ, Ma H, Fu YY, Fang DC, Sun H, Fan LZ, Han M,
Liu CM, Yang SH. J Mater Chem 2012; 22: 7461–7467.
61. Shang WH, Zhang XY, Zhang M, Fan ZT, Sun Y, Han M, Fan LZ. Nanoscale 2014; 6:
5799–5806.
62. E YF, Bai LL, Fan LZ, Han M, Zhang XY, Yang SH. J Mater Chem 2011; 21: 819.
63. CZ Wang, E YF, Fan LZ, Wang ZH, Liu HB, Li YL, Yang SH, Li YL. Adv Mater 2007;
19: 3677–3681.
64. Yu Z, Pestell TG, Lisanti MP, Pestell RG. Int J Biochem Cell Biol 2012; 44: 2144–2151.
65. Bao Q, Zhao Y, Renner A, Niess H, Seeliger H, Jauch K-W, Bruns CJ. Cancers 2010; 2:
1629–1641.
66. Reya T, Morrison SJ, Clarke MF, Weissman IL. Nature 2001; 414: 105–111.
67. Guo RH, Zhou SX, Li YC, Li XH, Fan LZ, Voelcker NH. ACS Appl Mater Interfaces
2015; 7: 23958–23966.
68. Zheng XT, Than A, Ananthanaraya A, Kim D-H, Chen P. ACS Nano 2013; 7:
6278–6286.
69. Biswal S, Resnick DL, Hoffman JM, Gambhir SS. Radiology 2007; 244: 651–671.
70. Li Y, Zhao Y, Cheng HH, Hu Y, Shi GQ, Dai LM, Qu LT. J Am Chem Soc 2012; 134:
15–18.
71. Guo BD, Liu Q, Chen ED, Zhu HW, Fang L, Gong JR. Nano Lett 2010; 10:
4975–4980.
72. Guo BD, Fang L, Zhang BH, Gong JR. Electron Lett 2011; 47: 663–664.
73. Guo BD, Fang L, Zhang BH, Gong JR. Insciences J 2011; 1: 80–89.
74. Li Q, Zhang S, Dai L, Li LS. J Am Chem Soc 2012; 134: 18932–18935.
75. Tetsuka H, Asahi R, Nagoya A, Okamoto K, Tajima I, Ohta R, Okamoto A. Adv Mater
2012; 24: 5333–5338.
76. Ponomarenko L, Schedin F, Katsnelson M, Yang R, Hill E, Novoselov K, Geim A. Science
2008; 320: 356–358.
77. Girit CO, Meyer JC, Erni R, Rossell MD, Kisielowski C, Yang L, Park CH, Crommie M,
Cohen ML, Louie SG. Science 2009; 323: 1705–1708.
78. Loh KP, Bao QL, Eda G, Chhowalla M. Nat Chem 2010; 2: 1015–1024.
79. Padilha, LA, Nootz G, Olszak PD, Webster S, Hagan DJ, Van Stryland EW, Levina L,
Sukhovatkin V, Brzozowski L, Sargent EH. Nano Lett 2011; 11: 1227–1231.
80. Luo ZT, Vora PM, Mele EJ, Johnson ATC, Kikkawa JM. Appl Phys Lett 2009; 94:
111909-1–3.
81. Trauzettel B, Bulaev DV, Loss D, Burkard G. Nat Phys 2007; 3: 192–196.
82. Collini E. Phys Chem Chem Phys 2012; 14: 3725–3736.
83. Agostinis P, Berg K, Cengel KA, Foster TH, Girotti AW, Gollnick SO, Hahn SM,
Hamblin MR, Juzeniene A, Kessel D, Korbelik M, Moan J, Mroz P, Nowis D, Piette J,
Wilson BC, Golab J. CA Cancer J Clin 2011; 61: 250–281.
84. Mattevi C, Eda G, Agnoli S, Miller S, Mkhoyan KA, Celik O, Mastrogiovanni D,
Granozzi G, Garfunkel E, Chhowalla M. Adv Funct Mater 2009; 19: 2577–2583.
85. Ishigami M, Chen JH, Cullen WG, Fuhrer MS, Williams ED. Nano Lett 2007; 7:
1643–1648.
86. Kudin KN, Ozbas B, Schniepp HC, Prud’homme RK, Aksay IA, Car R. Nano Lett 2007;
8: 36–41.
87. Gomez-Navarro C, Meyer JC, Sundaram RS, Chuvilin A, Kurasch S, Burghard M, Kern
K, Kaiser U. Nano Lett 2010; 10: 1144–1148.
88. Jung I, Field DA, Clark NJ, Zhu Y, Yang D, Piner RD, Stankovich S, Dikin DA, Geisler
H, Ventrice CA, Ruoff RS. J Phys Chem C 2009; 113: 18480–18486.
89. Wilson NR, Pandey PA, Beanland R, Young RJ, Kinloch IA, Gong L, Liu Z, Suenaga K,
Rourke JP, York SJ, Sloan J. ACS Nano 2009; 3: 2547–2556.
90. Eda G, Lin YY, Mattevi C, Yamaguchi H, Chen HA, Chen IS, Chen CW, Chhowalla M.
Adv Mater 2010; 22: 505–509.
91. Tan XY, Li YC, Li XH, Zhou SX, Fan LZ, Yang SH. Chem Commun 2015; 51:
2544–2546.
92. Ge J, Lan M, Zhou B, Liu W, Guo L, Wang H, Jia Q, Niu G, Huang X, Zhou H, Meng
X, Wang P, Lee CS, Zhang W, Han X. Nat Commun 2014; 5: 4596–4603.
93. Denk W, Strickler JH, Webb WW. Science 1990; 248: 73–76.
94. Helmchen F, Denk W. Nat Meth 2005; 2: 932–940.
Chapter 7
Graphene quantum dots (GQDs), sharing advantages and properties of both graphene
and semiconductor quantum dots (QDs) have been widely used for biomedical
application mainly in bioimaging and biosensing since they have small lateral size
and excellent electrochemical and optical properties, capability to bind with a variety
of aromatic biomolecules through a π–π stacking interaction and/or electrostatic
interaction, as well as fine water solubility, excellent biocompatibility and minimal
toxicity, which also made them the ideal materials for drug delivery. In this chapter,
we provide an overview of significant advances in GQDs-based carriers for delivery of
chemical and biological drugs.
1. Introduction
Nowadays, tremendous efforts have been contributed to the development of
targeted drug delivery systems. Among these systems, nanomaterials play
important roles, including improving the solubility of hydrophobic drugs,
achieving targeted delivery, controlling drug release and depressing toxic side
effects.1 An ideal nanomaterial for drug delivery needs to have excellent drug
loading capacity, active functional groups to conjugate drugs or targeting
antibodies, good biocompatibility, and low toxicity. Organic nanoparticles
like liposomes, micelles and inorganic nanoparticles like gold nanoparticles,
115
Figure 1. Intracellular distribution of DOX–GQD–FA and free DOX. HeLa cells were incu-
bated with DOX–GQD–FA (33.7 mg/mL) at 37°C for (A) 0.5 h, (B) 8 h, and (C) 24 h. Left
to right columns: bright field s, GQD fluorescence, DOX fluorescence, overlay of the corre-
sponding. The GQDs and DOX were excited under 488 nm laser irradiation, and fluorescence
signals were collected from 500 to 530 nm and 552 to 617 nm, respectively. Reprinted with
permission from Ref. [26]. Copyright (2014) Elsevier.
excitation was about 18.8%, which was higher than other GQDs reported in
the literature. More importantly, the surface-passivated PEG on GQDs can
not only enhance PL intensity but also load drug by hydrogen bonding, and
improved the solubility of GQDs. Moreover, they investigated the loading
and release behavior of GQDs–PEG. High specific surface area endowed
them with high loading capability (2.5 mg/mg) to carry drug. The highest
loading capacity was observed at the neutral condition, rather than acidic or
alkaline conditions (0.9 mg/mg at pH 5.5, 2.5 mg/mg at pH 7.4, and 1.1
mg/mg at pH 9). On the contrary, the release of DOX was different: 37% at
pH 7.4, 75%, and 51% at pH 5.5 and 9.0 after 48 h, more easily under acidic
and alkaline conditions.
In these three papers mentioned above, the loading and release behavior of
DOX were all pH dependent. And there are two main reasons for this, one is
attributed to the increased hydrophilicity and solubility of DOX at pH 5.0.29,30
Figure 2. Schematic illustration of f-GQDs for cell imaging and combined gene-targeting
agents delivery. Probe 1: the inhibitor probe of miRNA-21 and probe 2: Survivin antisense
oligodeoxynucleotide. Reprinted with permission from Ref. [18]. Copyright (2015) American
Chemical Society.
4. Discussion
Although until now not too many researches have been reported in the field
of GQDs based drug delivery systems, we can already see the special advan-
tages in them. Thanks to their excellent electrochemical and optical proper-
ties and large surface area, they can bind with a variety of aromatic
biomolecules, and may have the potential to monitor the whole process of
targeted drug delivery. Their small lateral size range enables GQDs to pene-
trate the nucleus by diffusion so as to enhance the toxicity of drugs to DNA
and gene delivery efficiency. Even though GQDs notably improve drug load-
ing efficacy by conjugation, they are not ideal for non-aromatic biomolecules,
and GQDs based composites are more promising in multifunctional drug
delivery. On the other hand, through the chemical binding, the drugs are
loaded on the surface, so they may not be as efficient in sustained and con-
trolled release as the 3D structure carriers. Also, there is still one problem we
need to pay attention to, which is the biosafety of GQDs. These carbon nano-
particles, like GQDs, may be non-toxicity to cells, but they are difficult to be
degraded and stay in the body for a long time. So, the long-term toxicity need
to be deeply studied for further usage.
References
1. Liang R, Wei M, Evans DG, Duan X. Inorganic nanomaterials for bioimaging, targeted
drug delivery and therapeutics. Chem Commun 2014; 50: 14071–14081.
2. Maurer N, Fenske DB, Cullis PR. Developments in liposomal drug delivery systems.
Expert Opin Biol Ther 2001; 1: 923–947.
3. Barenholz Y. Liposome application: problems and prospects. Curr Opin Colloid In 2001;
6: 66.
4. Miura Y, Takenaka T, Toh K, Wu S, Nishihara H, Kano MR, et al. Cyclic RGD-linked
polymeric micelles for targeted delivery of platinum anticancer drugs to glioblastoma
through the blood-brain tumor barrier. ACS Nano 2013; 7: 8583–8592.
5. Paciotti GFD, Kingston GI, Tamarkin L. Colloidal gold nanoparticles: A novel nanoparti-
cle platform for developing multifunctional tumor-targeted drug delivery vectors. Drug
Dev Res 2006; 67: 47–54.
6. Bacon M, Bradley SJ, Nann T. Graphene quantum dots. Part Syst Charact 2014; 31:
415–428.
7. Ding H, Wei J-S, Xiong H-M. Nitrogen and sulfur co-doped carbon dots with strong blue
luminescence. Nanoscale 2014; 6: 13817–13823.
8. Jiang F, Chen D, Li R, Wang Y, Zhang G, Li S, et al. Eco-friendly synthesis of size-
controllable amine-functionalized graphene quantum dots with antimycoplasma properties.
Nanoscale 2013; 5: 1137.
9. Li Y, Zhao Y, Cheng H-H, Hu Y, Shi G-Q, Dai L-M, et al. Nitrogen-doped graphene
quantum dots with oxygen-rich functional groups. J Am Chem Soc 2012; 134: 15–18.
10. Abdullah-Al-Nahain, Lee J-E, In I, Lee H, Lee KD, Jeong JH, et al. Target delivery and
cell imaging using hyaluronic acid functionalized graphene quantum dots. Mol Pharm
2013; 10: 736–3744.
11. Huang C-L, Huang C-C, Mai F-D, Yen C-L, Tzing S-H, Hsieh HT, et al. Application
of paramagnetic graphene quantum dots as a platform for simultaneous dual-modality
bioimaging and tumor targeted drug delivery. J Mater Chem B 2015; 3: 651–664.
12. Yuan X, Liu Z, Guo Z, Ji Y, Jin M, Wang X. Cellular distribution and cytotoxicity of
graphene quantum dots with different functional groups. Nanoscale Res Lett 2014; 9:
1–9.
13. Zheng XT, Than A, Ananthanaraya A, Kim D-H, Chen P. Graphene quantum dots as
universal fluorophores and their use in revealing regulated trafficking of insulin receptors
in adipocytes. ACS Nano 2013; 7: 6278–6286.
14. Chandra A, Deshpande S, Shinde DB, Pillai VK, Singh N. Mitigating the cytotoxicity of
graphene quantum dots and enhancing their applications in bioimaging and drug delivery.
ACS Macro Lett 2014; 3: 1064–1068.
15. Wang C, Wu C, Zhou X, Han T, Xin X, Wu J, et al. Enhancing cell nucleus accumulation
and DNA cleavage activity of anti-cancer drug via graphene quantum dots. Sci Rep 2013;
3: 2852–2860.
16. Wang XY, Lei R, Huang HD, Wang N, Yuan L, Xiao RY, et al. The permeability and
transport mechanism of graphene quantum dots (GQDs) across the biological barrier.
Nanoscale 2015; 7: 2034–2041.
17. Shang W, Zhang X, Zhang M, Fan Z, Sun Y, Han M, et al. The uptake mechanism and
biocompatibility of graphene quantum dots with human neural stem cells. Nanoscale
2014; 6: 5799–5806.
18. Dong H, Dai W, Ju H, Lu H, Wang S, Xu L, et al. Multifunctional poly(L-lactide)-
polyethylene glycol-grafted graphene quantum dots for intracellular MicroRNA imaging
and combined specific-gene-targeting agents delivery for improved therapeutics. ACS
Appl Mater Interfaces 2015; 7: 11015–11023.
19. Gottesman MM, Fojo T, Bates SE. Multidrug resistance in cancer: Role of ATP-
dependent transporters. Nat Rev Cancer 2002; 2: 48–58.
20. Gottesman MM. Mechanism of cancer drug resistance. Annu Rev Med 2002; 53:
615–627.
21. Some S, Gwon AR, Hwang E, Bahn GH, Yoon Y, Kim Y, et al. Cancer therapy using
ultrahigh hydrophobic drug-loaded graphene derivatives. Sci Rep 2014; 3: 6314–6323.
22. Sui X, Luo C, Wang C, Zhang F, Zhang J, Guo S. Graphene quantum dots enhance
anticancer activity of cisplatin via increasing its cellular and nuclear uptake. Nanomedicine:
NBM 2016; 12: 1997–2006.
23. Zhou X, Zhang Y, Wang C, Wu X, Yang Y, Zheng B, et al. Photo-Fenton reaction of
graphene oxide: A new strategy to prepare graphene quantum dots for DNA cleavage.
ACS Nano 2012; 6: 6592–6599.
24. Zheng B, Wang C, Xin X, Liu F, Zhou X, Zhang J, et al. Electron transfer from graphene
quantum dots to the copper complex enhances its nuclease activity. J Phys Chem C 2014;
118: 7637–7642.
25. Zheng B, Wang C, Wu C, Zhou X, Lin M, Wu X, et al. Nuclease activity and cytotoxicity
enhancement of the DNA intercalators via graphene oxide. J Phys Chem C 2012; 116:
15839–15846.
26. Wang X, Sun X, Lao J, He H, Cheng T, Wang M, et al. Multifunctional graphene quan-
tum dots for simultaneous targeted cellular imaging and drug delivery. Colloid Surface B
2014; 122: 638–644.
27. Chen H, Wang Z, Zong S, Chen P, Zhu D, Wu L, et al. A graphene quantum dot-based
FRET system for nuclear-targeted and real-time monitoring of drug delivery. Nanoscale
2015; 7: 15477–15486.
28. Wang Z, Xia J, Zhou C, Via B, Xia Y, Zhang F, et al. Synthesis of strongly green photo-
luminescent graphene quantum dots for drug carrier. Colloid Surface B 2013; 112:
192–196.
29. Liu Z, Robinson JT, Sun XM, Dai HJ. PEGylated nanographene oxide for delivery of
water-insoluble cancer drugs. J Am Chem Soc 2008; 130: 10876–10877.
30. Liu Z, Sun XM, Nakayama-Ratchford N, Dai HJ. Supramolecular chemistry on water-
soluble carbon nanotubes for drug loading and delivery. ACS Nano 2007; 1: 50–56.
31. Bhirde AA, Patel V, Gavard J, et al. Targeted killing of cancer cells in vivo and in vitro
with EGF-directed carbon nanotube-based drug delivery. ACS Nano 2009; 3: 307–316.
32. Yang XQ, Grailer JJ, Rowland IJ, et al. Multifunctional stable and pH-responsive polymer
vesicles formed by hetero functional triblock copolymer for targeted anticancer drug
delivery and ultrasensitive MR imaging. ACS Nano 2010; 4: 6805–6817.
33. Pan L, He Q, Liu J, Chen Y, Ma M, Zhang L, et al. Nuclear-targeted drug delivery of
TAT peptide-conjugated monodisperse mesoporous silica nanoparticles. J AmChem Soc
2012; 134: 5722–5725.
34. Alber F, Dokudovskaya S, Veenhoff LM, Zhang W, Kipper J, Devos D, et al. The molecu-
lar architecture of the nuclear pore complex. Nature 2007; 450: 695–701.
35. Habiba K, Encarnacion-Rosado J, Garcia-Pabon K, Villalobos-Santos JC, Makarov VI,
Avalos JA, et al. Improving cytotoxicity against cancer cells by chemo-photodynamic
combined modalities using silver-graphene quantum dots nanocomposites. Int J Nanomed
2016; 11: 107–119.
Chapter 8
Maoquan Chu
Graphene quantum dots (GQDs) are novel biomaterials with numerous applications
in biomedical fields. Because these nanomaterials may be suitable for clinical use,
their bio-safety should be investigated carefully. Previous studies have focused on the
in vitro cytotoxicity of GQDs, showing that GQDs were biocompatible materials
and exhibited low cytotoxicity to living cells. However, animal studies may be more
useful for determining whether these nanomaterials can be used in humans. Several
research groups have investigated the in vivo toxicity of GQDs in mice, rats, and
zebrafish embryos. In this chapter, toxicity studies of GQDs are described along with
suggestions for future work.
1. Introduction
In the previous chapters, the physical and chemical properties of graphene
quantum dots (GQDs) have been described, as well as their biomedical appli-
cations. GQDs may be excellent carriers for drug delivery.1–5 These nanoma-
terials have also been used as biosensors in biomedical detection assays,6–16
and as photosensitizers and photothermal agents for cancer photodynamic
and photothermal therapies.17,18
GQDs have the largest specific surface area of all nanomaterials as they are
single-atom-thick materials and only 1–10 nm in lateral dimension size.
127
128 M. Chu
tail veins for in vivo and ex vivo imaging studies. The tumors had been
grown from human breast cancer KB cells. GQD doses were 5 mg/kg and
10 mg/kg body weight. Tumor sites gradually emitted fluorescence 2 h post
injection; the brightest fluorescence was emitted 6–12 h post injection. Because
these GQDs have no red or near-infrared fluorescence (NIR), effects on major
organs such as lung and kidney could not be detected using this technique.
However, when the organs and tumors were isolated after sacrifice at 24 h post
injection, ex vivo imaging showed that the heart, liver, spleen, lung, kidney, and
tumor all emitted bright fluorescence in GQD-treated groups compared with
those in the untreated group. The tumor and major organ fluorescence in mice
treated with a 10 mg/kg dose was significantly brighter than that of mice
treated with a 5 mg/kg dose. These results indicate that GQDs could easily
distribute through blood and lymphatic capillaries and then arrive at various
organs. At 24 h post injection, the fluorescent signal at the tumor sites was
significantly decreased compared with fluorescence at 2–12 h post injection.
This finding implies that most GQDs may have been excreted from the body
by then, or that GQD fluorescence was not stable in mouse organs. They also
noted that fluorescence was brighter in tumors than in the major organs,
indicating that the GQDs may have passively targeted the tumors through the
enhanced permeability and retention (EPR) effect of tumors.
To improve the sensitivity of in vivo fluorescent imaging, Chong et al.20
conjugated GQDs with a NIR fluorescent dye and then monitored their dis-
tribution in tumor-bearing mice. The GQDs had been synthesized by an
oxidative-cutting method using graphene as a precursor. These GQDs had a
sheet-like (planar) structure, and measured 3–5 nm in size and 0.5–1 nm in
thickness. These GQDs contained a large number of oxygen atoms (the
molar ratio of carbon to oxygen was 16:7). They then modified the GQDs
with polyethylene glycol (PEG) and conjugated them with Cy7 (a commonly
used NIR fluorescent dye) via amide bond for in vivo biodistribution analysis.
When the 4T1 breast cancer tumor-bearing mice were intravenously injected
with the GQD-PEG–Cy7 conjugate (15 mg/kg), fluorescent imaging
showed that the GQDs mainly accumulated in kidneys and tumors. At 48 h
post injection, both kidneys and tumors showed only weak fluorescence. This
finding implies that GQDs injected intravenously could passively target
tumor tissue, and are then excreted through the renal pathway. Similar results
were observed when tumor-bearing mice were intraperitoneally injected with
15 mg/kg of GQD-PEG–Cy7. The passive tumor targeting and fast clear-
ance of GQDs may have been a result of their small size.
GQDs may also be used as drug carriers delivered orally or transdermally.
However, in vivo biodistribution patterns of GQDs in mice after oral gavage
130 M. Chu
or cutaneous penetration are not clear. Our research group30 recently demon-
strated that small (87.97 ± 30.83 nm) and large (472.08 ± 249.17 nm) rGO
nanomaterials accumulated mainly in the kidney of normal mice 1 and 15
days after gavage. Other organs, such as the stomach, lung, liver, and spleen,
also contained small amounts of rGO. Sixty days after the last gavage, most
of the rGO had been excreted renally. Although both rGO and GQDs are
graphene nanosheets, GQDs are significantly smaller than rGO. Therefore,
the in vivo biodistribution after oral gavage may be similar, but the excretion
rate of the GQDs should be faster. We found that 3-mercaptopropionic acid
capped-CdTe QDs several nanometers in diameter could penetrate mouse
skin and migrate into major organs.33 QD levels in spleen on days 1, 3, or 7
after skin staining were higher than those found in other organs, such as lung
and liver. QDs may have been excreted renally as the kidney contained
numerous cadmium atoms on days 3 and 7. GQDs are also several nanome-
ters in length, and their surfaces are also rich in carboxyl groups. Therefore,
GQDs may penetrate animal skin and distribute throughout the body, includ-
ing deposition in the spleen. These GQDs may be excreted in the urine.
132 M. Chu
in several parameters, such as ALT, BUN, and WBC, between the treated and
control groups.
134 M. Chu
6. Discussion
Recent studies have reported that GQDs exert little or no effect on mouse or
rat health. One reason may be their small size and unique carbon-based com-
position. Ultrasmall GQDs are easily excreted from the murine body. This
may be a common character of ultrasmall nanoparticles. Choi et al.36 reported
that semiconductor QDs with no more than 5.5 nm in final hydrodynamic
diameter were excreted efficiently in the urine of rats and mice.
As mentioned earlier, the toxicity of nanomaterials is related to their dose.
Wang et al.31 incubated zebrafish embryos with GQDs (0, 12.5, 25, 50, 100,
and 200 mg/mL) for 4–96 h post fertilization. Low concentrations of GQDs
were non-toxic to the zebrafish embryos. However, the hatching and heart
rates decreased, and mortality in the embryos increased in a concentration-
dependent manner. Various embryonic malformations, including pericardial
edema, vitelline cyst, bent spine, and bent tail, were observed when GQD
concentrations were increased to 200 mg/mL. In published studies, the high-
est dose of GQDs used in mice was 20 mg/kg every other day for 14 days.
The toxicity of GQDs in higher concentrations (>20 mg/kg) to mammals is
unknown, and should be investigated.
The toxicity of nanomaterials to animals is also related to their surface
potentials and the properties of ligand molecules used to modify
References
1. Lv O, Tao Y, Qin Y, Chen C, Pan Y, Deng L, Liu L, Kong Y. Highly fluorescent and
morphology-controllable graphene quantum dots-chitosan hybrid xerogels for in vivo imag-
ing and pH-sensitive drug carrier. Mater Sci Eng C Mater Biol Appl 2016; 67: 478–486.
2. Justin R, Tao K, Román S, Chen D, Xu Y, Geng X, Ross IM, Grant RT, Pearson A,
Zhou G, MacNeil S, Sun K, Chen B. Photoluminescent and superparamagnetic
reduced graphene oxide–iron oxide quantum dots for dual-modality imaging, drug
delivery and photothermal therapy. Carbon 2016; 97: 54–70.
3. Wang X, Sun X, Lao J, He H, Cheng T, Wang M, Wang S, Huang F. Multifunctional
graphene quantum dots for simultaneous targeted cellular imaging and drug delivery.
Colloid Surface B 2014; 122: 638–644.
4. Wang Z, Xia J, Zhou C, Via B, Xia Y, Zhang F, Li Y, Xia L, Tang J. Synthesis of strongly
green-photoluminescent graphene quantum dots for drug carrier. Colloid Surface B
2013; 112: 192–196.
5. Sui X, Luo C, Wang C, Zhang F, Zhang J, Guo S. Graphene quantum dots enhance
anticancer activity of cisplatin via increasing its cellular and nuclear uptake. Nanomed-
Nanotechnol 2016; 7: 1997–2006.
6. Wang L, Tricard S, Yue P, Zhao J, Fang J, Shen W. Polypyrrole and graphene quantum
dots @ Prussian Blue hybrid film on graphite felt electrodes: Application for amperometric
determination of L-cysteine. Biosens Bioelectron 2016; 77: 1112–1118.
7. Zhou X, Jiang Y, Li Z, Gu Z, Wang G. Improved activity and thermo-stability of the horse
radish peroxidase with graphene quantum dots and its application in fluorometric detec-
tion of hydrogen peroxide. Spectrochim Acta A 2016; 165: 106–113.
8. Hu T, Zhang L, Wen W, Zhang X, Wang S. Enzyme catalytic amplification of miRNA-155
detection with graphene quantum dots-based electrochemical biosensor. Biosens
Bioelectron 2016; 77: 451–456.
9. Vasilescu I, Eremia SAV, Kusko M, Radoi A, Vasile E, Radu G-L. Molybdenum disulphide
and graphene quantum dots as electrode modifiers for laccase biosensor. Biosens
Bioelectron 2016; 75: 232–237.
136 M. Chu
10. Poon C-Y, Li Q, Zhang J, Li Z, Dong C, Lee AW-M, Chan W-H, Li H-W. FRET-based
modified graphene quantum dots for direct trypsin quantification in urine. Anal Chim
Acta 2016; 917: 64–70.
11. Liang R-P, Qiu W-B, Zhao H-F, Xiang C-Y, Qiu J-D. Electrochemiluminescence reso-
nance energy transfer between graphene quantum dots and graphene oxide for sensitive
protein kinase activity and inhibitor sensing. Anal Chim Acta 2016; 904: 58–64.
12. Nirala NR, Abraham S, Kumar V, Bansal A, Srivastava A, Saxena PS. Colorimetric detec-
tion of cholesterol based on highly efficient peroxidase mimetic activity of graphene
quantum dots. Sensor Actuatr B-Chem 2015; 218: 42–50.
13. Shi J, Chan C, Pang Y, Ye W, Tian F, Lyu J, Zhang Y, Yang M. A fluorescence resonance
energy transfer (FRET) biosensor based on graphene quantum dots (GQDs) and gold
nanoparticles (AuNPs) for the detection of mecA gene sequence of Staphylococcus aureus.
Biosens Bioelectron 2015; 67: 595–600.
14. He Y, Wang X, Sun J, Jiao S, Chen H, Gao F, Wang L. Fluorescent blood glucose monitor
by hemin-functionalized graphene quantum dots based sensing system. Anal Chim
Acta 2014; 810: 71–78.
15. Razmi H, Mohammad-Rezaei R. Graphene quantum dots as a new substrate for immo-
bilization and direct electrochemistry of glucose oxidase: Application to sensitive glucose
determination. Biosens Bioelectron 2013; 41: 498–504.
16. Lu J, Yan M, Ge L, Ge S, Wang S, Yan J, Yu J. Electrochemiluminescence of blue-lumi-
nescent graphene quantum dots and its application in ultrasensitive aptasensor for adeno-
sine triphosphate detection. Biosens Bioelectron 2013; 47: 271–277.
17. Wo F, Xu R, Shao Y, Zhang Z, Chu M, Shi D, Liu S. Multimodal cancer theraputics via
magneto-mechanical force and near-infrared laser-triggered doxorubicin-loaded graphene
quantum dot-coated hollow magnetic nanospheres. Theranostics 2016; 6: 485–500.
18. Ge J, Lan M, Zhou B, Liu W, Guo L, Wang H, Jia Q, Niu G, Huang X, Zhou H,
Meng X, Wang P, Lee C-S, Zhang W, Han X. A graphene quantum dot photodynamic
therapy agent with high singlet oxygen generation. Nat Commun 2014; 5: 4596.
19. Zhang M, Bai L, Shang W, Xie W, Ma H, Fu Y, Fang D, Sun H, Fan L, Han M, Liu C,
Yang S. Facile synthesis of water-soluble, highly fluorescent graphene quantum dots as a
robust biological label for stem cells. J Mater Chem 2012; 22: 7461–7467.
20. Chong Y, Ma Y, Shen H, Tu X, Zhou X, Xu J, Dai J, Fan S, Zhang Z. The in vitro and
in vivo toxicity of graphene quantum dots. Biomaterials 2014; 35: 5041–5048.
21. Nurunnabi M, Khatun Z, Huh KM, Park SY, Lee DY, Cho KJ, Lee Y-k. In Vivo biodis-
tribution and toxicology of carboxylated graphene quantum dots. ACS Nano 2013; 7:
6858–6867.
22. Yuan X, Liu Z, Guo Z, Ji Y, Jin M, Wang X. Cellular distribution and cytotoxicity of
graphene quantum dots with different functional groups. Nanoscale Res Lett 2014;
9: 108.
23. Liu Y, Xu L-P, Dai W, Dong H, Wen Y, Zhang X. Graphene quantum dots for the inhibi-
tion of β amyloid aggregation. Nanoscale 2015; 7: 19060–19065.
24. Zhang XY, Yin JL, Peng C, Hu WQ, Zhu ZY, Li WX, et al. Distribution and biocompat-
ibility studies of graphene oxide in mice after intravenous administration. Carbon 2011;
49: 986–95.
25. Fu CH, Liu TL, Li LL, Liu HY, Liang QH, Meng XW. Effects of graphene oxide on the
development of offspring mice in lactation period. Biomaterials 2015; 40: 23–31.
26. Yang K, Gong H, Shi XZ, Wan JM, Zhang YJ, Liu Z. In vivo biodistribution and toxicol-
ogy of functionalized nano-graphene oxide in mice after oral and intraperitoneal adminis-
tration. Biomaterials 2013; 34: 2787–2795.
27. Yang K, Wan JM, Zhang SA, Zhang YJ, Lee ST, Liu Z. In vivo pharmacokinetics, long-
term biodistribution, and toxicology of PEGylated graphene in mice. ACS Nano 2011; 5:
516–22.
28. Xu S, Zhang Z, Chu M. Long-term toxicity of reduced graphene oxide nanosheets:
Effects on female mouse reproductive ability and offspring development. Biomaterials
2015; 54: 188–200.
29. Liang S, Xu S, Zhang D, He J, Chu M. Reproductive toxicity of nanoscale graphene oxide
in male mice. Nanotoxicology 2015; 19: 92–105.
30. Zhang D, Zhang Z, Liu Y, Chu M, Yang C, Li W, Shao Y, Yue Y, Xu R. The short- and
long-term effects of orally administered high-dose reduced graphene oxide nanosheets on
mouse behaviors. Biomaterials 2015; 68: 100–113.
31. Wang ZG, Zhou R, Jiang D, Song JE, Xu Q, Si J, Ping CY, Xin Z, Lu G, Zhen LJ, Hong
Z, Bin L. Toxicity of Graphene Quantum Dots in Zebrafish Embryo. Biomed Environ Sci
2015; 28: 341–351.
32. Huang CL, Huang CC, Mai FD, Yen CL, Tzing SH, Hsieh HT, Ling YC, Chang JY.
Application of paramagnetic graphene quantum dots as a platform for simultaneous dual-
modality bioimaging and tumor-targeted drug delivery. J Mater Chem B 2015; 3:
651–664.
33. Chu M, Wu Q, Wang J, Hou S, Miao Y, Peng J, Sun Y. In vitro and in vivo transdermal
delivery capacity of quantum dots through mouse skin. Nanotechnology 2007; 18: 455103
1–6.
34 Markovic ZM, Ristic BZ, Arsikin KM, Klisic DG, Harhaji-Trajkovic LM, Todorovic-
Markovic BM, Kepic DP, Kravic-Stevovic TK, Jovanovic SP, Milenkovic MM,
Milivojevic DD, Bumbasirevic VZ, Dramicanin MD, Trajkovic VS. Graphene quantum
dots as autophagy-inducing photodynamic agents. Biomaterials 2012; 33: 7084–7092.
35. Ristic BZ, Milenkovic MM, Dakic IR, Todorovic-Markovic BM, Milosavljevic MS,
Budimir MD, Paunovic VG, Dramicanin MD, Markovic ZM, Trajkovic VS. Photodynamic
antibacterial effect of graphene quantum dots. Biomaterials 2014; 35: 4428–4435.
36. Choi HS, Liu W, Misra P, Tanaka E, Zimmer JP, Ipe BI, Bawendi MG, Frangioni JV.
Renal clearance of quantum dots. Nat Biotechnol 2007; 25: 1165–1170.
37. Bozich JS, Lohse SE, Torelli MD, Murphy CJ, Hamers RJ, Klaper RD. Surface chemistry,
charge and ligand type impact the toxicity of gold nanoparticles to Daphnia magna.
Environ Sci Nano 2014; 1: 260–270.
Index
A I
amino-functionalization, 19–20, 29–31 in vivo biodistribution, 128–130
in vivo clearance, 128
B in vivo imaging, 99, 107, 109–110
bottom-up strategies, 7, 11
M
C medical and pharmaceutical analyses,
cancer cellular imaging, 99–101 57–58
chemotherapy drug delivery, 117 microwave-solvothermal method, 19,
chromatographic detection, 89 21, 28, 37
controllable synthesis, 13
O
D optical detection, 83
doping, 39, 44–45, 47, 52
P
E photoluminescence, 19
electrochemical detection, 89 phototoxicity, 133–134
plasmons, 39–44, 48–49
F
fluorescence imaging, 97–98, 107, S
109–110 sensor, 57–58, 60–61, 66, 71
food analysis, 77–80, 83–84 stem cellular imaging, 102–103, 110
G T
gene delivery, 121–122 top-down strategies, 7–8
graphene quantum dots, 1, 19–20, 39, toxicity, 127–128, 130, 132–135
57, 64, 77, 80, 97, 115–116, 127 tunable fluorescent properties, 4
139