Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

DOI: 10.1002/asia.

201700814 Focus Review

Sensor Design

Fluorescence and Sensing Applications of Graphene Oxide and


Graphene Quantum Dots: A Review
Peng Zheng and Nianqiang Wu*[a]

Chem. Asian J. 2017, 12, 2343 – 2353 2343 T 2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Focus Review

Abstract: Graphene oxide and graphene quantum dots are nance energy transfer (FRET) process. In this review, the
attractive fluorophores that are inexpensive, nontoxic, pho- origin of fluorescence and the mechanism of excitation
tostable, water-soluble, biocompatible, and environmentally wavelength-dependent fluorescence of graphene oxide and
friendly. They find extensive applications in fluorescent bio- graphene quantum dots are discussed. Sensor design strat-
sensors and chemosensors, in which they serve as either flu- egies based on graphene oxide and graphene quantum
orophores or quenchers. As fluorophores, they display tuna- dots are presented. The applications of these sensors in
ble photoluminescence emission and the “giant red-edge health care, the environment, agriculture, and food safety
effect”. As quenchers, they exhibit a remarkable quenching are highlighted.
efficiency through either electron transfer or Fçrster reso-

Introduction Fluorescence Principles of Graphene Oxide


Fluorescence Origin
Graphene-based nanomaterials are two-dimensional (2D)
atomic crystals composed of sp2-hybridized carbon atoms.[1] Electronic energy transitions are responsible for the fluores-
The family includes graphene, graphene oxide (GO), reduced cence of GO.[5] As shown in Figure 1, each ’’fingerprint’’ fluores-
graphene oxide (rGO), and graphene quantum dots (GQDs). cence band of functionalized GO comes from specific electron-
Ideally, graphene is completely composed of sp2-hybridized ic transitions between the antibonding and the bonding mo-
carbon atoms, which exhibit no fluorescence due to a zero op- lecular orbitals such as s*!n, p*!n, and p*!p.[5] GO general-
tical bandgap. Due to the introduction of functional groups, ly contains various oxygen-containing functional groups, such
GO, rGO, and GQDs consist of mixtures of sp2 and sp3 carbons as hydroxyl groups (COH), carboxyl groups (COOH), carbonyl
that open the optical bandgap and result in fluorescence.[2] GO (C=O), epoxy (C-O-C), and aromatic rings (C=C). As a result,
is rich in oxygen-containing functional groups. The chemical multiple fluorescence peaks, which correspond to different
reduction of partial oxygen-containing functional groups in GO electronic transitions, are usually excited simultaneously. When
leads to the formation of rGO.[3] GQDs are cut from a single to these fluorescence peaks overlap with each other, a single
several atomic layers of graphene sheets several nanometers in broad peak is observed, as shown in Figure 1.[8]
lateral size.[4] Typically, GQDs contain both sp2 and sp3 carbon Functional groups, dopants, lateral size, localized domains,
atoms, which are inherent from synthesis processes. GQDs and strain in GO can alter the electronic energy transitions and
have a size-dependent optical bandgap due to the quantum lead to changes in the fluorescence peak position, width, and
confinement.[5] GO and GQDs are inexpensive, nontoxic, photo- shape. For example, after GO is incubated in aqueous KOH or
stable, water-soluble, biocompatible, and environmentally HNO3, it becomes enriched with OH or COOH groups, respec-
friendly, and they exhibit excitation-wavelength-dependent tively. Consequently, the fluorescence peak position and inten-
fluorescence.[6] In addition, their 2D surface shows strong non- sity are tuned as shown in Figure 1.[8] Bear in mind that the
covalent interactions with adsorbed biomolecules through p–p bandgap of GO is created as a result of the mixture of sp2 and
interactions, electrostatic forces, or hydrogen bonding, which sp3 carbon atoms after the introduction of functional groups.
provide a chemically tunable platform for bioconjugation.[6–7] The electronic band structure can be tuned by varying the
Therefore, both GO and GQDs have found extensive applica- ratio of sp2/sp3 carbon atoms, which allows the fluorescence to
tions in fluorescent sensors in recent years. They can be em- be tuned from the visible light to the near-infrared (NIR) wave-
ployed not only as a tunable fluorophore but also as an effec- length range. The fluorescence of GQDs has the same origin as
tive fluorescence quencher. GO. Doping has been found to modulate the electronic struc-
We begin with a discussion of the physical principles of fluo- ture of GQDs and thus affect the fluorescence emission. For ex-
rescence of GO and GQDs. We then show how to design fluo- ample, in nitrogen-doped GQDs, the fluorescence is dominated
rescent sensors by utilizing the unique physicochemical prop- by the n!p* transition between the aromatic rings that con-
erties of GO and GQDs. Also, we highlight the applications of tain N and the conjugate structure of graphene.[9] An increase
GO- and GQD-based fluorescent sensors in a variety of fields. in the doping percentage of nitrogen in GQDs could redshift
the emission and improve photostability.[10] Due to quantum
confinement, changes in the size of GQDs can modulate the
electronic structure and thus shift the fluorescence from green,
to red, to NIR. In addition, changes in the localized domains, in
[a] P. Zheng, Prof. N. Wu which p electrons are confined in localized sp2 regions, can
Department of Mechanical and Aerospace Engineering also alter the fluorescence of GO.[10–11]
West Virginia University
PO Box 6106, Morgantown, WV 26506 (USA)
E-mail: nick.wu@mail.wvu.edu
The ORCID identification number(s) for the author(s) of this article can be
found under https://doi.org/10.1002/asia.201700814.

Chem. Asian J. 2017, 12, 2343 – 2353 www.chemasianj.org 2344 T 2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Focus Review

Figure 1. a) UV/Vis absorption of as-synthesized GO and GO treated with KOH and HNO3. b) Fluorescence spectra corresponding to a); inset: the different
electronic transitions. Reproduced from Ref. [8] with permission from The Royal Society of Chemistry.

Excitation-Wavelength-Dependent Fluorescence dipoles are realigned with one another and reduce the interac-
tion energy to achieve equilibrium. For a common fluorophore
Breakdown of Kasha’s rule
in a polar solvent, the solvation dynamics are orders of magni-
According to Kasha’s rule, the peak fluorescence wavelength tude faster than fluorescence. The solvation is usually complet-
for organic dyes and inorganic quantum dots is independent ed prior to fluorescence. Therefore, the final fluorescence only
of the excitation wavelength because fluorescence always undergoes a small redshift. However, if the solvation dynamics
starts at the band edge after rapid internal relaxation. Howev- are slowed down to the fluorescence timescale, fluorescence
er, Kasha’s rule does not hold for GO. Excitation-wavelength- occurs simultaneously as the energy of the excited fluorophore
dependent fluorescence occurs in GO suspended in polar sol- is reduced, which leads to a strong dependence of fluores-
vents, as shown in Figure 2.[8] When the excitation wavelength cence on the excitation. The giant red-edge effect takes place
was increased from l = 325 to 650 nm, the fluorescence band not only in GO but also in GQDs and carbon dots.
was redshifted to longer wavelengths.

Mr. Peng Zheng received his B.S. degree in


2011 from Central South University, China,
and his M.S. degree in 2014 from West Virgin-
ia University (WVU), USA. He is working
toward his Ph.D. degree at WVU in the group
of Prof. Nianqiang Wu. He strives to under-
stand the optical properties of nanostructures
and to develop biosensors.

Dr. Nianqiang (Nick) Wu is currently Professor


Figure 2. Excitation-wavelength-dependent fluorescence of GO. Reproduced of Materials Science at West Virginia Universi-
from Ref. [8] with permission from The Royal Society of Chemistry. ty, USA. He is a Fellow of Royal Society of
Chemistry (FRSC) and a Fellow of the Electro-
chemical Society (FECS). He serves on the
Board of Directors in the Electrochemical Soci-
Giant red-edge effect ety (ECS) and is Chair of Sensor Division. His
research interest lies in photocatalysts and
Excitation-wavelength-dependent fluorescence of GO is due to photoelectrochemical cells, batteries and su-
the “giant red-edge effect”.[12] As shown in Figure 3, the giant percapacitors, and biosensors and labs-on-
red-edge effect occurs in a polar solvent, such as water, in chips.
which the solvation introduces an additional relaxation step.
For an excited fluorophore in a polar solvent, the dipole is not
in equilibrium with the solvent dipole. During solvation, these

Chem. Asian J. 2017, 12, 2343 – 2353 www.chemasianj.org 2345 T 2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Focus Review

tion wavelength without changing any other factor, such as


chemical composition and size.

Influence of pH on GO fluorescence
In addition, the fluorescence of GO is dependent on the pH
value of the solvents, as shown in Figure 4.[11a] Under acidic
conditions, GO exhibits a broad fluorescence peak centered at
around l = 680 nm. As the pH gradually increased, the l =
680 nm peak reduced in intensity, and finally disappeared with
the emergence of two new peaks at around l = 500 nm under
basic conditions. This pH-dependent fluorescence is believed
to be caused by an excited-state proton transfer, as shown in
Figure 3. Under acidic conditions, the COOH group is largely
Figure 3. Fluorescence properties of GO. GO exhibits excitation-wavelength- deprotonated and exists in the form of ionic COO@ in the
dependent fluorescence (or giant red-edge effect) due to solvation. It also
displays an excited-state protonation from COOH groups. The measured
ground state. Upon excitation, the excited-state protonation
fluorescence of GO is a superposition of contributions from COH and COOH from ionic COO@ to COOH contributes to the broad fluores-
groups. Reprinted with permission from Ref. [12]. Copyright (2014) American cence peak at l = 668 nm. In comparison, the fluorescence
Chemical Society. emission under basic conditions is associated with the excited
COO@ moieties.
Excitation-wavelength-dependent fluorescence due to strain
Sensing Applications of GO and GQDs
Given the fast solvation dynamics ( & 10@12 s@1) of the polar sol-
vents used, there must be some important factors relevant to GO can be used as a fluorophore with tunable fluorescence
material itself that drastically slow down the dipole realign- emission by varying the chemical composition, size, and other
ment process to the timescale of fluorescence. In this regard, factors. GO can also be used as a highly efficient quencher
efforts have been made to study the effects of oxygen per- through either charge-transfer or resonance-energy-transfer
centage, doping percentage, disorder, and strain in GO, which processes owing to its super electrical conductivity and 2D
have been commonly hypothesized to be the underlying rea- planar structure. The dual roles of GO as both a fluorophore
sons for the giant red-edge effect.[11a, 13] Of the aforementioned and a quencher open up new opportunities for sensor design,
factors, strain has been found to correlate with the giant red- as shown in Figure 5. Therefore, graphene-based nanomaterials
edge effect.[14] When there is an out-of-plane strain, the plane are finding increasing applications in environment, agriculture,
of the GO sheet becomes curved, which generates a strong biomedicine, food safety, and homeland security. This section
local reorganization potential. The local potential prevents a outlines different design strategies for fluorescent biosensors
fast reorientation of the excited dipoles with the solvated di- and also highlights some important applications.
poles in the oxygen-containing functional groups nearby and
drastically slows down the dipole realignment process to the
GO as a fluorescent label
fluorescence timescale, which generates the giant red-edge
effect.[14] The redshift in the peak fluorescence wavelength in GO has been widely used as a fluorescent label, just like organ-
GO is also found to scale linearly with the increase in the exci- ic dyes and inorganic quantum dots.[15] However, the broad
tation wavelength.[8, 12] The peak fluorescence can be tuned emission peak limits the sensing performance when the fluo-
from visible to the NIR region by simply changing the excita- rescence spectral intensity is transduced as the signal. Instead,

Figure 4. pH-dependent photoluminescence of GO excited at l = 440 nm. a) The pH effect on photoluminescence. b) Plot of the photoluminescence intensi-
ties at l = 683, 506, and 479 nm with respect to the pH of the GO solutions. Reprinted with permission from Macmillan Publishers Ltd: [Scientific Reports]
(Ref. [11a]), copyright (2011).

Chem. Asian J. 2017, 12, 2343 – 2353 www.chemasianj.org 2346 T 2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Focus Review

Figure 5. Schematic illustration of sensor design based on GO. a) GO as a fluorescent label. b) GO as a FRET donor and its fluorescence can be quenched by a
quencher, such as a gold nanoparticle, through FRET. c) Charge transfer occurs between GO and a fluorophore in close proximity. d) GO acts as a quencher
and can quench the fluorescence of a nearby fluorophore through FRET.

GO is remarkably suitable for NIR biological imaging by using (GONs) were functionalized with transferrin (Trf) molecules to
two-photon excitation spectroscopy thanks to the giant red- target the cancer cells. The Trf-GO needs a much lower laser
edge effect.[12] Two-photon excitation is based on the simulta- power to generate sharp and well-resolved GON images. Be-
neous absorption of two photons that have only half the cause Trf can specifically target cancer cells, so does the Trf-
energy of a single photon traditionally used to excite the same GON complex. The process of killing cancer cells by using the
event.[16] The selection of a light source with a longer wave- Trf-GON complex is thus directly reflected in the GON fluores-
length, particularly NIR excitation, greatly helps to reduce pho- cence images. Liu et al. have also reported cellular and deep-
totoxicity, lower the background signal, and increase the light tissue imaging by using two-photon fluorescence spectroscopy
penetration depth in a biological sample. GO with excitation- based on GQDs.[20] Wang et al. took advantage of two-photon
wavelength-dependent fluorescence is well coupled with two- spectroscopy to image endogenous biological cyanide in plant
photon excitation spectroscopy because its excitation wave- tissues based on a GQD–gold nanoparticles complex.[21] Two-
length can be deliberately chosen to be fallen into one of the photon imaging with GO as a fluorescent label has been em-
biological windows to allow selective biological imaging.[17] ployed in biomedical fields to target liver tumor cells,[22] breast
The first and second NIR biological windows are located at l = tumor cells,[23] melanoma cells,[24] and also for food safety.[21]
650 to 950 and 1000 to 1350 nm, at which biological samples
are almost transparent to incident light.[18] As an example, GO
GO as a donor in charge transfer
exhibits different fluorescence when excited in the first and
second biological windows by using two-photon excitation The fluorescence of GO can be quenched by a charge-transfer
fluorescence spectroscopy, as shown in Figure 6a and b.[17] The process, in which GO acts as a charge donor.[25] Compared to
advantages of the use of GO in biological imaging under two- FRET, which is a long-range energy transfer mechanism, charge
photon excitation are further manifested by applications in transfer is a short-range energy transfer process. It requires the
cancer treatment, as shown in Figure 6c–f.[19] GO nanoparticles charge donor and the acceptor to be separated by within

Figure 6. Two-photon fluorescence imaging a) tunable excitation-wavelength-dependent two-photon imaging in the first and second biological transparency
windows. b) Two-photon fluorescence imaging of methicillin-resistant Staphylococcus aureus (MRSA) by using aptamer-modified GO, in which the excitation
wavelength for b1)–b4) were l = 1160, 880, 980, and 760 nm, respectively. Reprinted by permission from Macmillan Publishers Ltd: [Scientific Reports]
(Ref. [17]), copyright (2014). c)–f) A comparison of in vitro two-photon luminescence imaging of GO nanoparticles and molecular dye FITC. Reprinted from
Ref. [19] with permissions from Wiley.

Chem. Asian J. 2017, 12, 2343 – 2353 www.chemasianj.org 2347 T 2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Focus Review

Figure 7. Fluorescent sensors based on charge transfer. a) Dopamine detection based on charge transfer between GO and dopamine through p–p stacking.
Reprinted with permission from Ref. [15c]. Copyright (2011) American Chemical Society. b) Hg2 + detection, in which Hg2 + is captured through a thymine-
Hg2 + -thymine (T-Hg-T) interaction and quenches the fluorescence of GO. Reprinted from Ref. [25], Copyright (2013), with permission from Elsevier. c) Hg2 +
and I@ can both be detected on the same GO fluorescent biosensor platform. Reprinted with permission from Ref. [27]. Copyright (2015) American Chemical
Society.

10 a.[26] Such a short distance is often realized through chemi- GO as a donor in FRET
cal bonding or physical adsorption. As an example, dopamine
molecules have been detected in a GO fluorescent sensor In a FRET process, GO can serve as an energy donor in which
based on charge transfer (Figure 7a.[15c] The dopamine mole- its fluorescence gets quenched by an energy acceptor, such as
cule is chemically bonded to GO through the p–p interaction gold nanoparticles or organic dyes.[30] FRET is a nonradiative
between GO and the aromatic rings in the molecules. In the energy-transfer process based on the dipole–dipole interac-
excited GO–dopamine complex, the excited electrons in GO tion.[26] It requires the donor’s emission spectrum to overlap
are transferred to the dopamine molecule, which results in with the acceptor’s absorption spectrum. The FRET efficiency is
quenching of the GO fluorescence. In addition, Wu’s research inversely proportional to the sixth power of the separation dis-
group used the charge-transfer mechanism to detect metal tance between the donor and acceptor. Therefore, for FRET to
cations,[25] for which GO serves as a fluorophore and an elec- be efficient, the energy donor and the acceptor need to be
tron donor whereas the metal ions act as the electron acceptor brought close to each other, usually within about 6 nm (the
(Figure 7b). GO is initially functionalized with a single-stranded Fçrster distance) if organic dyes serve as both the energy
DNA (ssDNA) aptamer, which can capture Hg2 + and form a donors and acceptors. This can be realized through, for exam-
hairpin structure on GO due to the formation of a thymine- ple, antigen–antibody interactions or DNA hybridization.[30b, 31]
Hg2 + -thymine (T-Hg-T) complex. Once the GO-aptamer-Hg2 + Because GO has a broad fluorescence emission peak, the con-
complex is formed, the fluorescence of GO is quenched due to dition of spectrum overlap can be easily met by using noble-
charge transfer from GO to Hg2 + . A similar sensor platform has metal nanoparticles as the quencher because they have a large
been used for I@ detection, as shown in Figure 7c.[27] Because absorption cross-section with absorption band tunable from
the binding constant[28] for Hg2 + and I@ is > 1029, much larger the UV to the NIR region. For example, GO has been used as
than that for Hg2 + and thymine, which is about 106, HgI2 is an energy donor in an immuno-biosensor system for rotavius
preferentially formed and pulls the Hg2 + ions away from the detection (Figure 8).[30b] Initially, gold nanoparticles were func-
GO surface, which leads to recovery of the quenched fluores- tionalized with the detection antibody of the rotavirus, and
cence and allows I@ to be detected. In addition, the GO sensor the capture antibody was immobilized the GO surface. The
platform based on charge transfer has been successfully ap- gold nanoparticles and GO could be brought into proximity
plied for the detection of various other environmental and ag- through the antigen–antibody interaction after capture of the
ricultural pollutants, such as Au3 + , Fe3 + , Fe2 + , Co2 + , Ni2 + , Cd2 + rotavirus cell. Consequentially, the fluorescence of GO was
, and HCl.[15d,e, 29] quenched by the gold nanoparticles by the FRET process. GO

Chem. Asian J. 2017, 12, 2343 – 2353 www.chemasianj.org 2348 T 2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Focus Review

penetrate into cells. Also, the effective quenching distance


could be extended to around 30 nm if GO is used as an
energy acceptor. Figure 9 highlights several FRET detection
schemes.[29a, 33] For example, 1) the fluorophore-labeled single-
stranded aptamers were initially captured on GO through the
p–p stacking force between the aromatic nucleobases of apta-
mer and the aromatic rings in GO, as shown in Figure 9a. The
fluorescence of the fluorophore, such as organic dyes and
quantum dots, was quenched by GO in proximity through the
FRET process. In the presence of analytes, the aptamers
formed hairpin or G-quadruplex structures, which have a weak
interacting force with GO, so they were released from GO and
turned on the fluorescence. 2) In Figure 9b, the fluorophore-la-
beled double-stranded DNA molecules were initially captured
on the GO surface, with one strand of DNA covalently bonded
to GO and the other strand of DNA labeled with a fluorophore.
Fluorescence of the fluorophore was quenched if the fluoro-
phore was controlled to be close to the GO surface (typically
less than 6 nm). The addition of analytes caused DNA dehy-
Figure 8. GO as a FRET donor in a sensor for rotavirus detection, in which
the fluorescence of GO is quenched by gold nanoparticles. Reprinted from bridization, and the fluorophore-labeled DNA strand formed a
Ref. [30b] with permission from Wiley. hairpin structure and detached from the GO surface. 3) In Fig-
ure 9c, the free-standing fluorophore-labeled detection anti-
bodies initially displayed fluorescence. When analytes were
has also been exploited as an energy donor in FRET sensors for present, the analytes were sandwiched between the detection
drug detection, disease diagnosis, pharmaceutical screening, antibody and the capture antibody, which brought the fluoro-
and so on.[30a, 32] phore close to the GO surface and led to quenching of the
fluorescence. 4) In Figure 9d, the fluorophore-labeled antigens
were initially attached to the capture antibody on the GO sur-
GO as an acceptor in FRET
face, with the fluorescence quenched. After addition of ana-
In a FRET process, GO can also act as an energy acceptor, in lytes, the fluorophore-labeled antigens were displaced away
which it quenches the fluorescence of an energy donor, such from GO due to the stronger affinity between the analyte and
as quantum dots and organic dyes.[29a, 33] The use of GO as a the capture antibody, which led to recovery of fluorescence.
quencher has several advantages: 1) The 2D planar surface For example, microRNA (miRNA) is an important type of bio-
area renders a large number of bonding sites; 2) the oxygen- marker because it regulates the expression of diverse genes
containing functional groups enable further bioconjugation; and can be related to many diseases, such as cancers and dia-
3) it is water-soluble, non-toxic, biocompatible, and able to betes. GO, with its ability to penetrate through intracellular

Figure 9. Schematic illustration of sensor design strategies with GO as a quencher. a) An aptamer-based sensor. b) A DNA-based sensor. c) A sandwich-struc-
tured immunoassay. d) A competitive immunoassay.

Chem. Asian J. 2017, 12, 2343 – 2353 www.chemasianj.org 2349 T 2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Focus Review

Figure 10. GO as a quencher in sensors. a) miRNA detection based on nanosized graphene oxide (NGO) and peptide nucleic acid (PNA), in which the fluores-
cence of the dye label on the PNA probe was initially quenched by GO but later recovered after the probe detached from NGO and hybridized with a target
miRNA. Reprinted with permission from Ref. [33a]. Copyright (2013) American Chemical Society. b) Detection of the ovarian cancer biomarker CA-125, based
on CRET. Reproduced from Ref. [36] with permission of The Royal Society of Chemistry.

membranes, has been used to build FRET sensors for the in been reported that the strong quenching capability of GO and
vitro detection of miRNA, as shown in Figure 10a.[33a] In this GQDs can be used to quench chemiluminescence through
sensor, a dye-labeled peptide nucleic acid (PNA) was initially chemiluminescence resonance energy transfer (CRET), as
quenched after being adsorbed onto GO. The GO-PNA-dye shown in Figure 10b, in which the cancer biomarker CA-125
complex was then delivered into the cytoplasm, in which the was detected based on CRET.[36] GQDs were used as the
target miRNA hybridized with the dye-labeled PNA and de- quenchers (energy acceptors) and assembled on a glass sub-
tached from GO, which led to fluorescence recovery. In addi- strate, then further functionalized with the capture antibody
tion, GO as a FRET quencher has also been used for in vitro for CA-125. The detection antibody for CA-125 was labeled
and in vivo molecular probing based on two-photon excitation with the horseradish peroxidase (HRP) enzyme. In the presence
spectroscopy.[34] By replacing the capture and detection re- of the CA-125 antigen, an antibody-antigen-antibody complex
agents, the GO sensor platform has been extended for the de- was formed with HRP and brought in proximity to GQDs. As a
tection of proteins, DNA, drugs, heavy metals, and so on.[33b, 35] result, the HRP enzyme catalyzed H2O2 to produce reactive
In addition to fluorescence, chemiluminescence provides an oxygen species (ROS), which further oxidized luminol to gener-
alternative for optical biosensing. Chemiluminescence occurs ate chemiluminescence. If HRP is brought close to GQDs in the
as a result of chemical reactions rather than direct photoexcita- presence of the CA-125 antigen, the produced chemilumines-
tion, which could lead to a low signal-to-noise ratio. It has cence is quenched by nearby GQDs. This “turn-off” sensor ena-

Chem. Asian J. 2017, 12, 2343 – 2353 www.chemasianj.org 2350 T 2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Focus Review

bled the CA-125 cancer biomarker to be detected sensitively. pared with most inorganic semiconductor quantum dots, GO
In addition, CRET sensors with GO or GQDs as quenchers have and GQDs are nontoxic and biocompatible. The strong excita-
also been used for the detection of other protein biomarkers, tion-wavelength-dependent fluorescence of GO and GQDs are
DNA, and environmental pollutants.[36–37] unique to fluorescent sensing and imaging, and can be ex-
tended to other photoluminescent devices. However, GO and
GQDs exhibit broad fluorescence peaks in the presence of a
GQDs in fluorescent biosensors
large variety of oxygen-containing moieties, as compared with
Although GQDs shares many common excellent properties the sharp fluorescence peak of organic dyes and inorganic
with GO, such as easy functionalization, photostability, water- quantum dots. If a narrow fluorescence peak is needed for
solubility, biocompatibility, and nontoxicity, GQDs have their some applications, it is necessary to purify and functionalize
own unique properties. For example, GQDs have a much small- GO.
er steric effect due to their small size and can easily penetrate Interestingly, GO is suitable for NIR biological imaging by
through intracellular membranes. This unique property makes using two-photon excitation fluorescence spectroscopy due to
GQDs a candidate material in drug delivery and in vivo and in a large two-photon absorption cross-section. Conversely, GO
vitro bioimaging.[38] For example, Lannazzo et al. employed can be purified to eliminate various oxygen-containing moiet-
GQDs for cancer-targeted drug delivery.[38c] Because GQDs also ies and further functionalized to tailor the fluorescence peak
have an excitation-wavelength-dependent emission, it can dis- position and width. In the future, further studies are needed to
play different colors in bioimaging under different excitations tune the fluorescence of purified and functionalized GO.
to meet the needs of various types of biosensing, as shown in GO and GQDs can be employed not only as fluorophores
Figure 11a.[38e] The ability of GQDs to penetrate through intra- but also as quenchers, which provides plenty of room for the
cellular membranes also allows in-depth bioimaging by using development of new fluorescent biosensors and bioimaging
NIR light, which enormously extends the scope of fluorescent systems. It is envisioned that the applications of GO and GQDs
bioimaging, as shown in Figure 11b.[38d] in in vitro and in vivo sensing and imaging will continue to
expand.
Multiplex detection is significant for practical applications of
Conclusions and Perspectives
biosensors because it will enormously improve the specificity
There are many superior properties of GO and GQDs. Com- and lower the cost of detection.[39] The dual roles of GO and
pared with organic dyes, GO and GQDs are photostable. Com- GQDs as a fluorophore and a quencher make them highly

Figure 11. Bioimaging with GQDs by using the giant red-edge effect. a) Fluorescence images of A549 cells cultured with GQDs under different excitations
(shown on the upper right of each image). Reprinted by permission from Macmillan Publishers Ltd: [Light: Science & Applications] (Ref. [38e]), copyright
(2015). b) Fluorescence images with GQDs at different depths in tissue. Reprinted with permission from Ref. [38d]. Copyright (2017) American Chemical Soci-
ety.

Chem. Asian J. 2017, 12, 2343 – 2353 www.chemasianj.org 2351 T 2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Focus Review

promising. They can transduce an excitation-wavelength-de- [10] S. K. Das, C. M. Luk, W. E. Martin, L. B. Tang, D. Y. Kim, S. P. Lau, C. I. Ri-
pendent fluorescence signal and simultaneously quench the chards, J. Phys. Chem. C 2015, 119, 17988 – 17994.
[11] a) C. Galande, A. D. Mohite, A. V. Naumov, W. Gao, L. J. Ci, A. Ajayan, H.
fluorescence from external fluorophores, which has strong im- Gao, A. Srivastava, R. B. Weisman, P. M. Ajayan, Sci. Rep. 2011, 1, 85;
plications in the detection of multiple events on a single bio- b) X. M. Sun, Z. Liu, K. Welsher, J. T. Robinson, A. Goodwin, S. Zaric, H. J.
sensor. Dai, Nano Res. 2008, 1, 203 – 212; c) Z. T. Luo, P. M. Vora, E. J. Mele,
Furthermore, the 2D planar structure of GO makes it ideal to A. T. C. Johnson, J. M. Kikkawa, Appl. Phys. Lett. 2009, 94, 111909; d) G.
Eda, Y. Y. Lin, C. Mattevi, H. Yamaguchi, H. A. Chen, I. S. Chen, C. W.
couple with lab-on-chip devices.[40] The integration of all these
Chen, M. Chhowalla, Adv. Mater. 2010, 22, 505 – 509; e) C. W. Chen, J.
functionalities into a single chip would greatly promote point- Robertson, J. Non-Cryst. Solids 1998, 227, 602 – 606.
of-care and field-deployable applications of biosensing devices [12] S. K. Cushing, M. Li, F. Q. Huang, N. Q. Wu, ACS Nano 2014, 8, 1002 –
in the future. It is expected that GO and GQDs, along with 1013.
[13] a) W. N. Nan, Y. A. Niu, H. Y. Qin, F. Cui, Y. Yang, R. C. Lai, W. Z. Lin, X. G.
other materials in the family of graphene-based materials, will
Peng, J. Am. Chem. Soc. 2012, 134, 19685 – 19693; b) Z. S. Qian, J. Zhou,
find increasing applications in healthcare, environmental pro- J. R. Chen, C. Wang, C. C. Chen, H. Feng, J. Mater. Chem. 2011, 21,
tection, agriculture, food safety, homeland security, and so on. 17635 – 17637.
[14] S. K. Cushing, W. Q. Ding, G. Chen, C. Wang, F. Yang, F. Q. Huang, N. Q.
Wu, Nanoscale 2017, 9, 2240 – 2245.
Acknowledgements [15] a) H. S. Kushwaha, R. Sao, R. Vaish, J. Appl. Phys. 2014, 116, 034701;
b) Y. X. Lu, H. Kong, F. Wen, S. C. Zhang, X. R. Zhang, Chem. Commun.
2013, 49, 81 – 83; c) J. L. Chen, X. P. Yan, K. Meng, S. F. Wang, Anal.
Research reported in this publication was supported by the Chem. 2011, 83, 8787 – 8793; d) H. J. Zhu, H. D. Xu, Y. H. Yan, K. Zhang, T.
National Institute of Neurological Disorders and Stroke of the Yu, H. Jiang, S. H. Wang, Sens. Actuators B 2014, 202, 667 – 673; e) L. He,
National Institutes of Health under award number J. N. Li, J. H. Xin, Biosens. Bioelectron. 2015, 70, 69 – 73; f) D. L. Meng, S. J.
Yang, D. M. Sun, Y. Zeng, J. H. Sun, Y. Li, S. K. Yan, Y. Huang, C. W. Bielaw-
R15NS087515. The content is solely the responsibility of the
ski, J. X. Geng, Chem. Sci. 2014, 5, 3130 – 3134.
authors and does not necessarily represent the official views of [16] P. T. C. So, C. Y. Dong, B. R. Masters, K. M. Berland, Annu. Rev. Biomed.
the National Institutes of Health. This work was also partially Eng. 2000, 2, 399 – 429.
supported by an NSF grant (CBET-1336205). [17] A. Pramanik, Z. Fan, S. R. Chavva, S. S. Sinha, P. C. Ray, Sci. Rep. 2014, 4,
6090.
[18] M. F. Tsai, S. H. G. Chang, F. Y. Cheng, V. Shanmugam, Y. S. Cheng, C. H.
Su, C. S. Yeh, ACS Nano 2013, 7, 5330 – 5342.
Conflict of interest [19] J. L. Li, H. C. Bao, X. L. Hou, L. Sun, X. G. Wang, M. Gu, Angew. Chem. Int.
Ed. 2012, 51, 1830 – 1834; Angew. Chem. 2012, 124, 1866 – 1870.
The authors declare no conflict of interest. [20] Q. Liu, B. D. Guo, Z. Y. Rao, B. H. Zhang, J. R. Gong, Nano Lett. 2013, 13,
2436 – 2441.
[21] L. L. Wang, J. Zheng, S. Yang, C. C. Wu, C. H. Liu, Y. Xiao, Y. H. Li, Z. H.
Keywords: biosensors · fluorescence · graphene · quantum Qing, R. H. Yang, ACS Appl. Mater. Interfaces 2015, 7, 19509 – 19515.
dots · sensors [22] Y. L. Shi, A. Pramanik, C. Tchounwou, F. Pedraza, R. A. Crouch, S. R.
Chavva, A. Vangara, S. S. Sinha, S. Jones, D. Sardar, C. Hawker, P. C. Ray,
ACS Appl. Mater. Interfaces 2015, 7, 10935 – 10943.
[1] a) K. S. Novoselov, V. I. Fal’ko, L. Colombo, P. R. Gellert, M. G. Schwab, K.
[23] A. Pramanik, S. R. Chavva, Z. Fan, S. S. Sinha, B. Priya, V. Nellore, P. C.
Kim, Nature 2012, 490, 192 – 200; b) A. K. Geim, K. S. Novoselov, Nat.
Ray, J. Phys. Chem. Lett. 2014, 5, 2150 – 2154.
Mater. 2007, 6, 183 – 191; c) A. H. Castro Neto, F. Guinea, N. M. R. Peres,
[24] C. Tchounwou, S. S. Sinha, B. P. V. Nellore, A. Pramanik, R. Kanchanapally,
K. S. Novoselov, A. K. Geim, Rev. Mod. Phys. 2009, 81, 109 – 162.
S. Jones, S. R. Chavva, P. C. Ray, ACS Appl. Mater. Interfaces 2015, 7,
[2] a) H. M. Huang, Z. B. Li, J. C. She, W. L. Wang, J. Appl. Phys. 2012, 111,
054317; b) A. Mathkar, D. Tozier, P. Cox, P. J. Ong, C. Galande, K. Balak- 20649 – 20656.
rishnan, A. L. M. Reddy, P. M. Ajayan, J. Phys. Chem. Lett. 2012, 3, 986 – [25] M. Li, X. J. Zhou, W. Q. Ding, S. W. Guo, N. Q. Wu, Biosens. Bioelectron.
991. 2013, 41, 889 – 893.
[3] a) H. A. Becerril, J. Mao, Z. Liu, R. M. Stoltenberg, Z. Bao, Y. Chen, ACS [26] J. R. Lakowicz, Principles of Fluorescence Spectroscopy, 3rd ed., Springer
Nano 2008, 2, 463 – 470; b) S. Stankovich, D. A. Dikin, R. D. Piner, K. A. US, 2006.
Kohlhaas, A. Kleinhammes, Y. Jia, Y. Wu, S. T. Nguyen, R. S. Ruoff, Carbon [27] D. Dinda, B. K. Shaw, S. K. Saha, ACS Appl. Mater. Interfaces 2015, 7,
2007, 45, 1558 – 1565; c) G. Eda, G. Fanchini, M. Chhowalla, Nat. Nano- 14743 – 14749.
technol. 2008, 3, 270 – 274. [28] H. Torigoe, A. Ono, T. Kozasa, Chem. Eur. J. 2010, 16, 13218 – 13225.
[4] a) D. Y. Pan, J. C. Zhang, Z. Li, M. H. Wu, Adv. Mater. 2010, 22, 734 – 738; [29] a) A. Kundu, R. K. Layek, A. Kuila, A. K. Nandi, ACS Appl. Mater. Interfaces
b) J. H. Shen, Y. H. Zhu, X. L. Yang, C. Z. Li, Chem. Commun. 2012, 48, 2012, 4, 5576 – 5582; b) D. Y. Wang, D. W. Wang, H. A. Chen, T. R. Chen,
3686 – 3699. S. S. Li, Y. C. Yeh, T. R. Kuo, J. H. Liao, Y. C. Chang, W. T. Chen, S. H. Wu,
[5] S. J. Zhu, J. H. Zhang, C. Y. Qiao, S. J. Tang, Y. F. Li, W. J. Yuan, B. Li, L. C. C. Hu, C. W. Chen, C. C. Chen, Carbon 2015, 82, 24 – 30.
Tian, F. Liu, R. Hu, H. N. Gao, H. T. Wei, H. Zhang, H. C. Sun, B. Yang, [30] a) S. Y. Kwak, J. K. Yang, S. J. Jeon, H. I. Kim, J. Yim, H. Kang, S. Kyeong,
Chem. Commun. 2011, 47, 6858 – 6860. Y. S. Lee, J. H. Kim, Adv. Funct. Mater. 2014, 24, 5119 – 5128; b) J. H. Jung,
[6] a) Y. W. Zhu, S. Murali, W. W. Cai, X. S. Li, J. W. Suk, J. R. Potts, R. S. Ruoff, D. S. Cheon, F. Liu, K. B. Lee, T. S. Seo, Angew. Chem. Int. Ed. 2010, 49,
Adv. Mater. 2010, 22, 3906 – 3924; b) D. R. Dreyer, S. Park, C. W. Bielawski, 5708 – 5711; Angew. Chem. 2010, 122, 5844 – 5847.
R. S. Ruoff, Chem. Soc. Rev. 2010, 39, 228 – 240; c) G. Eda, M. Chhowalla, [31] Y. He, B. N. Jiao, Talanta 2017, 163, 140 – 145.
Adv. Mater. 2010, 22, 2392 – 2415. [32] Y. Shi, H. C. Dai, Y. J. Sun, J. T. Hu, P. J. Ni, Z. Li, Analyst 2013, 138, 7152 –
[7] K. P. Loh, Q. L. Bao, G. Eda, M. Chhowalla, Nat. Chem. 2010, 2, 1015 – 7156.
1024. [33] a) S. R. Ryoo, J. Lee, J. Yeo, H. K. Na, Y. K. Kim, H. Jang, J. H. Lee, S. W.
[8] M. Li, S. K. Cushing, X. J. Zhou, S. W. Guo, N. Q. Wu, J. Mater. Chem. Han, Y. Lee, V. N. Kim, D. H. Min, ACS Nano 2013, 7, 5882 – 5891; b) X. Q.
2012, 22, 23374 – 23379. Liu, R. Aizen, R. Freeman, O. Yehezkeli, I. Willner, ACS Nano 2012, 6,
[9] J. Sun, S. W. Yang, Z. Y. Wang, H. Shen, T. Xu, L. T. Sun, H. Li, W. W. Chen, 3553 – 3563.
X. Y. Jiang, G. Q. Ding, Z. H. Kang, X. M. Xie, M. H. Jiang, Part. Part. Syst. [34] M. Yi, S. Yang, Z. Y. Peng, C. H. Liu, J. S. Li, W. W. Zhong, R. H. Yang, W. H.
Char. 2015, 32, 434 – 440. Tan, Anal. Chem. 2014, 86, 3548 – 3554.

Chem. Asian J. 2017, 12, 2343 – 2353 www.chemasianj.org 2352 T 2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Focus Review
[35] a) Y. F. Liu, M. Luo, X. Xiang, C. H. Chen, X. H. Ji, L. Chen, Z. K. He, Chem. [38] a) M. L. Chen, Y. J. He, X. W. Chen, J. H. Wang, Bioconjugate Chem. 2013,
Commun. 2014, 50, 2679 – 2681; b) A. Huang, W. W. Li, S. Shi, T. M. Yao, 24, 387 – 397; b) P. Nigam, S. Waghmode, M. Louis, S. Wangnoo, P.
Sci. Rep. 2017, 7, 40772; c) L. Lin, Y. Liu, X. Zhao, J. H. Li, Anal. Chem. Chavan, D. Sarkar, J. Mater. Chem. B 2014, 2, 3190 – 3195; c) D. Iannazzo,
2011, 83, 8396 – 8402; d) A. Shirai, K. Nakashima, K. Sueyoshi, T. Endo, H. A. Pistone, M. Salamo, S. Galvagno, R. Romeo, S. V. Giofre, C. Branca, G.
Hisamoto, Analyst 2017, 142, 472 – 477; e) X. X. Sun, J. Fan, Y. P. Zhang, Visalli, A. Di Pietro, Int. J. Pharm. 2017, 518, 185 – 192; d) L. L. Feng, Y. X.
H. L. Chen, Y. Q. Zhao, J. X. Xiao, Biosens. Bioelectron. 2016, 79, 15 – 21; Wu, D. L. Zhang, X. X. Hu, J. Zhang, P. Wang, Z. L. Song, X. B. Zhang,
f) L. H. Tang, D. Y. Li, J. H. Li, Chem. Commun. 2013, 49, 9971 – 9973; g) B. W. H. Tan, Anal. Chem. 2017, 89, 4077 – 4084; e) D. Qu, M. Zheng, J. Li,
Tan, H. M. Zhao, L. Du, X. R. Gan, X. Quan, Biosens. Bioelectron. 2016, 83, Z. G. Xie, Z. C. Sun, Light Sci. Appl. 2015, 4, e364.
267 – 273; h) X. Weng, S. Neethirajan, Biosens. Bioelectron. 2016, 85, [39] P. Zuo, X. J. Li, D. C. Dominguez, B. C. Ye, Lab Chip 2013, 13, 3921 – 3928.
649 – 656; i) M. Li, X. J. Zhou, S. W. Guo, N. Q. Wu, Biosens. Bioelectron. [40] a) Y. V. Stebunov, O. A. Aftenieva, A. V. Arsenin, V. S. Volkov, ACS Appl.
2013, 43, 69 – 74. Mater. Interfaces 2015, 7, 21727 – 21734; b) C. K. Chua, M. Pumera, Elec-
[36] I. Al-Ogaidi, H. L. Gou, Z. P. Aguilar, S. W. Guo, A. K. Melconian, A. K. A. troanalysis 2013, 25, 945 – 950; c) G. Perry, V. Thomy, M. R. Das, Y. Coffini-
Al-Kazaz, F. K. Meng, N. Q. Wu, Chem. Commun. 2014, 50, 1344 – 1346. er, R. Boukherroub, Lab Chip 2012, 12, 1601 – 1604.
[37] a) C. Chen, B. X. Li, J. Mater. Chem. B 2013, 1, 2476 – 2481; b) M. Iranifam,
P. R. Babakalak, A. Imani-Nabiyyi, M. M. Abolghasemi, A. Khataee, Anal.
Methods 2016, 8, 3496 – 3502; c) X. Y. Liu, Z. L. Han, F. Li, L. F. Gao, G. L.
Manuscript received: June 1, 2017
Liang, H. Cui, ACS Appl. Mater. Interfaces 2015, 7, 18283 – 18291; d) Y.
Revised manuscript received: July 23, 2017
He, H. Cui, J. Mater. Chem. 2012, 22, 9086 – 9091; e) Y. He, H. Cui, J. Phys.
Chem. C 2012, 116, 12953 – 12957; f) Y. P. Dong, Y. Zhou, J. Wang, J. J. Accepted manuscript online: July 25, 2017
Zhu, Anal. Chem. 2016, 88, 5469 – 5475. Version of record online: August 30, 2017

Chem. Asian J. 2017, 12, 2343 – 2353 www.chemasianj.org 2353 T 2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

You might also like