SlideShare a Scribd company logo
1 of 13
Download to read offline
Biotechnology
Journal
DOI 10.1002/biot.201300001 Biotechnol. J. 2013, 8, 1280–1291
1280 © 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
1 Introduction
Cell metabolism boasts a complicated network. Exploring
the tremendous amount of information contained within
this network is of great significance to related studies. To
be able to visualize and monitor cell activities in vivo with
high spatial and temporal resolution is attractive because
traditional analysis methods destroy living cells to extract
metabolites of interest. Since the discovery of green fluo-
rescent protein (GFP), scientists have used site-directed
and random mutagenesis approaches to develop fluores-
cent protein (FP) mutants, and have created a family of
proteins that almost span the fluorescence spectrum. This
family has enlightened efforts to construct genetically
encoded fluorescent biosensors based on protein/protein
interactions and intramolecular conformational changes
targeted to living cells or tissues. Recently, another sort of
genetically encoded biosensor, the RNA-based biosensor,
has been developed for tracking small-molecule metabo-
lites such as adenosine, ADP, S-adenosyl methionine
(SAM), guanine, and GTP, and shows more advantages
than FP-based biosensors. Here, we describe different
modes of constructing fluorescent biosensors and review
recent advances in the use of genetically encoded fluo-
rescent biosensors for tracing intracellular metabolism.
2 Fluorescent proteins: Gifts from the ocean
When a bright, greenish protein in jellyfish extracts was
observed by Shimomura in the 1960s [1], the glow of FPs
Review
Imaging and tracing of intracellular metabolites utilizing
genetically encoded fluorescent biosensors
Chang Zhang*, Zi-Han Wei* and Bang-Ce Ye
Laboratory of Biosystems and Microanalysis, State Key Laboratory of Bioreactor Engineering, East China University of Science
and Technology, Shanghai, China
Intracellular metabolites play a crucial role in characterizing and regulating corresponding cellular
activities. Tracking intracellular metabolites in real time by traditional means was difficult until the
powerful toolkit of genetically encoded biosensors was developed. Over the past few decades, iter-
ative improvements of these biosensors have been made, resulting in the effective monitoring of
metabolites. In this review, we introduce and discuss the recent advances in the use of genetical-
ly encoded biosensors for tracking some key metabolites, such as ATP, cAMP, cGMP, NADH, reac-
tive oxygen species, sugar, carbon monoxide, and nitric oxide. A brief phylogeny of fluorescent pro-
teins and several typical construction modes for genetically encoded biosensors are also
described. We also discuss the development of novel RNA-based sensors, which are genetically
encoded biosensors active at the transcriptional level.
Keywords: Genetically encoded fluorescent biosensors · In vivo imaging · Metabolite
Correspondence: Prof. Bang-Ce Ye, Lab of Biosystems and Microanalysis,
State Key Laboratory of Bioreactor Engineering, East China University of
Science and Technology, Meilong RD 130, Shanghai 200237, China
E-mail: bcye@ecust.edu.cn
Abbreviations: 2OG, 2-Oxogluatarate; BRET, bioluminescence resonance
energy transfer; cAMP, cyclic adenosine monophosphate; cGMP, cyclic
guanosine 5’-monophosphate; CFP, cyan fluorescent protein; cpFP, circu-
lar permutation fluorescent protein; DFHBI, 3,5-difluoro-4-hydroxybenzyli-
dene imidazolinone; FLIP, fluorescence indicator protein; FP, fluorescent
protein; FRET, fluorescence resonance energy transfer; GFP, green fluores-
cent protein; GPCR, G-protein coupled receptor; PBP, periplasmic binding
protein; PKA, cAMP-dependent protein kinase; PKG, cGMP-dependent pro-
tein kinase; OHP, organic hydroperoxide; RFP, red fluorescent protein;
ROS, reactive oxygen species; SAM, S-adenosyl methionine; YFP, yellow
fluorescent protein
Received 31 MAR 2013
Revised 02 AUG 2013
Accepted 26 AUG 2013
Supporting information
available online
* These authors contributed equally to this work.
Color online: refer to online PDF file for figures in color.
1281
Biotechnol. J. 2013, 8, 1280–1291
www.biotecvisions.com
© 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
provided the ability to light up the invisible metabolism in
a living life form. In 1992, Prasher and colleagues [2] used
a cDNA library from Aequorea victoria to clone the gene
encoding the aequorin companion protein, GFP. They
demonstrated GFP’s promising future, a gift for cellular
biologists: genetically encoded GFP could be expressed in
a variety of biological systems without additional factors
or external enzyme components. Stable chromophore for-
mation within a cylinder structure makes GFP a powerful
cellular reporter [3]. A hexapeptide sequence starting at
amino acid 64 (number indicates the position in the intact
peptide sequence) determines the light absorption prop-
erties of GFP. The adjacent Ser65-Tyr66-Gly67 sequence
forms the chromophore, and fluorescence results from the
oxidization of the Tyr66 a–b carbon bond [4]. The crystal
structure for GFP shows the cyclic tripeptide chro-
mophore buried in the center of an 11-stranded β-barrel
cylinder structure [5]. In further studies of the cylinder
structure, mutagenesis of the amino acid residues sur-
rounding the chromophore significantly affected the
spectral properties of the protein. A series of FPs, with
emission wavelengths ranging from the blue to yellow
regions of the visible spectrum, and “enhanced” (E) FPs,
with increased protein folding efficiency and maturation
at physiological temperature (Supporting information,
Table S1), were produced [6].
Over the past 15 years, Anthozoa-based FPs spanning
the visible spectrum have been characterized and opti-
mized for imaging applications [6]. DsRed-based variants
and the mFruit (“m” for monomer) series are two repre-
sentative FP families with longer wavelength emissions
[7]. Members of the mFruit family, including mHoneydew,
mBanana, mOrange, mTangerine, mStrawberry, mCherry,
mApple, mPlum, and dTomato (“d” for dimer) derive from
monomeric red FP (RFP) 1, a mutant of DsRed (33 amino
acid substitutions) [8]. They have been useful in multi-
color imaging experiments. EosFP, another type of pho-
toactivatable FP, from the reef coral Lobophyllia hem-
prichii can be photoconverted from green to red fluores-
cence by near-ultraviolet light [9, 10]. Mutagenesis was
used to develop a monomeric protein named mEosFP. An
improved form, mIrisFP, enables pulse-chase experiments
with superb resolution and provides an excellent fusion
marker for imaging in living cells.
Figure 1. Design methods for construct-
ing single fluorophore sensors. The units
of a dimeric binding domain can be
directly attached to the original N and
C termini of an FP (A) or to novel N and
C termini of a cpFP (C). A monomeric
binding domain can be inserted into an
FP (B) or a cpFP (D). The binding or
effective domain undergoes a structural
change upon substrate (red star) bind-
ing or in response to other influencing
factors, such as ROS, leading to a
change in the fluorescent signal of the
sensor.
Biotechnology
Journal
Biotechnol. J. 2013, 8, 1280–1291
www.biotechnology-journal.com
1282 © 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
3 Design modes for genetically encoded
fluorescent biosensors
3.1 Single fluorophore sensors
Proteins can tolerate circular permutation, which pro-
duces novel N- and C-termini from different portions of
the protein, while maintaining a stable structure. Tsien
and coworkers [11], researching circular permutation and
receptor insertion within GFPs, showed that GFP is sur-
prisingly robust for circular permutation and insertion,
and offered a new strategy for creating genetically encod-
ed indicators [12–14]. Circular permutation FP (cpFP) is
especially attractive because the binding domain of inter-
est undergoes a weak structural change upon substrate
binding, and the signal change is amplified when the
domain is placed in the sensitive region of the chro-
mophore. Depending on the structure of the receptor,
such as the distance between N and C termini, oligomer-
ization, etc., researchers can choose a proper design
method for constructing cpFP-based sensors (Fig. 1). Giv-
en the structural similarity between GFP and its mutants,
many active cpFP have been produced using yellow FP
(YFP) [15–17], Venus [18], and RFP [19, 20]. Recently,
cpYFP-based sensors for NADH were developed by
inserting cpYFP into dimeric NADH binding domain Rex
as pictured in Fig. 1C.
In addition to circular permutation, GFP family mem-
bers are also amenable to protein fragmentation [21]. When
the fragments of an FP are brought into proximity with
each other, they can reunite to form a functional protein
(Fig.  2). This method, termed bimolecular fluorescence
complementation, has been exploited to detect protein-
protein interactions in living cells and plants [22–25].
A great advantage of this strategy is its high sensitivity,
since GFP fragments in the separated state only display a
modest background fluorescence until they are brought
closer by protein-protein interactions [26].
3.2 Fluorescence resonance energy transfer-based
reporter systems
Fluorescence resonance energy transfer (FRET) is used in
the design of indicators for optically imaging biochemical
and physiological functions in living cells [27, 28].
A genetically encoded FRET biosensor consists of a
recognition module, which specifically binds a ligand,
sandwiched between two variants of GFP – typically cyan
FP (CFP) and YFP [29, 30]. The efficiency of fluorescence
energy transfer between the two fluorophores is highly
dependent on their distance and orientation [31–33]. Tra-
ditional FRET sensors can be divided into two classes, as
described below.
Bimolecular probes are designed based on protein
interaction and are suitable for studying the dissociation
or association of a protein upon ligand binding [34, 35].
Donor and acceptor FPs are fused to interacting proteins
separately to form a pair of FRET sensors. When the inter-
action of proteins X and Y brings the FRET sensors clos-
er, the intermolecular FRET signal efficiency between
donor and acceptor FPs is elevated (Fig. 3A). This mech-
anism has been used to construct probes for measuring
the cytosolic Ca2+ concentration and protein interactions
within the same cell, using Fura-2 with super-enhanced
CFP and YFP as a FRET pair [36]. Single-chain probes are
designed based on a protein conformational change that
occurs upon ligand binding or in response to another
effector. Kolossov et al. [37] inserted redox-sensitive link-
ers between CFP and YFP to engineer single-chain FRET
sensors for reactive oxygen species (ROS) detection.
FRET-based sensors are also suitable for detecting lig-
ands of interest in vivo [38]. The selected binding domain
is sandwiched between FP pairs, usually CFP and YFP.
A conformational change in the domain upon ligand bind-
ing likely leads to a change in the distance or orientation
between the CFP and YFP chromophores, resulting in a
FRET efficiency change (Fig. 3B). Optimization is impor-
tant to obtain sensors with obvious FRET ratio changes or
Figure 2. When the interacting proteins
X and Y bind to each other, the seg-
ments of a fluorescent protein are
brought into proximity to form a func-
tional protein with recovered fluorescent
signal.
1283
Biotechnol. J. 2013, 8, 1280–1291
www.biotecvisions.com
© 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
a dynamic range consistent with concentration level of
the target molecules of interest in cell metabolism [31, 32].
Traditional bioluminescence resonance energy trans-
fer (BRET)-based assays have also been used for detect-
ing protein interactions [39–41]. In a BRET-based sensor,
luciferase is usually used as the BRET donor and an FP as
the BRET acceptor. However, using BRET to monitor pro-
tein complexes at the subcellular level has been limited by
the amount of light emission intrinsic to the luciferase
donor [42]. Recently, Dragulescu-Andrasi et al. [43]
demonstrated the use of red light-emitting BRET systems
for investigating protein interactions in deep tissue and in
small animal tumor models, which may help to eliminate
this drawback. Binkowski et al. [44] employed a circularly
permuted form of firefly luciferase to construct a lumines-
cent biosensor for intracellular cyclic adenosine mono-
phosphate (cAMP) detection. They studied signal trans-
duction mediated by G-protein coupled receptors
(GPCRs), and showed that luciferase is suitable for the
construction of genetically encoded biosensors as cpFPs.
4 Biosensors for tracing metabolism
4.1 Visualization of ATP levels inside cells
ATP is the major energy currency in cellular processes.
Real-time monitoring of ATP levels inside individual living
cells would help elucidate the compartmentation of ATP
and its role in modulating ion channels and signaling cas-
cades [45]. Imamura et al. [46] generated a series of FRET-
based indicators for ATP, and showed that the ATP levels
differed between the mitochondrial matrix and cyto-
plasm/nucleus in HeLa cells. Additional research demon-
strated that the ATP-generating pathway changed in
response to nutritional changes in the environment. The
probes used in these studies show high selectivity for
ATP over other nucleotides. By modulating the affinity
through substitution of different effector domains, or
mutation of residues at the binding interface, the probes
can measure ATP levels ranging from 2 μM to 8 mM. On
attaching a proper signal sequence, these sensors could
monitor any intracellular compartment of interest and
facilitate biological research in other fields.
Since the absolute levels of ATP, ADP, and AMP can
fluctuate, the ratio of [ATP]/[ADP] is thought to be a more
reliable indicator of energy status in cells. Berg et al. [47]
presented a novel cpFP-based fluorescent probe for meas-
uring the cellular ATP/ADP ratio. They measured the
energy level by competitive binding of ATP and ADP in a
cell without perturbing the cellular energy balance, in
contrast to sensors based on luciferase [48]. The regula-
tion of the metabolic machinery in cancer is complicated.
A better understanding of differential cancer cell metab-
olism would greatly benefit therapeutics. Zadran et al. [49]
monitored cytosolic ATP in tumor cells with a sensor that
could also be used in mitochondria, nuclei, and the endo-
plasmic reticulum. These efforts will improve our under-
standing of cellular and subcellular energy variations in
normal and abnormal cells [50, 51].
4.2 Fluorescent indicators of cAMP
cAMP is a second messenger of many GPCRs and regu-
lates cAMP-dependent protein kinase (PKA) and
exchange proteins activated by cAMP (Epacs) to mediate
cellular functions. Spatial and temporal readouts of cAMP
levels are vital to the comprehension of network regula-
tion in various signaling cascades. DiPilato et al. [52] con-
structed fluorescent indicators to report intracellular
Figure 3. Schematic presentation of
(A) a bimolecular FRET probe and (B) a
single-chain FRET probe. (A) When the
interacting proteins X and Y bind to each
other, the fluorescent pair is brought
into proximity and a FRET phenomenon
appears. (B) When the binding domain
in a single-chain FRET probe undergoes
conformational changes upon ligand
binding, the distance or orientation
between the fluorescent pair changes
and then alters the FRET efficiency.
Biotechnology
Journal
Biotechnol. J. 2013, 8, 1280–1291
www.biotechnology-journal.com
1284 © 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
cAMP dynamics and Epac activation. They revealed that
the cAMP response at the membrane was faster than that
in the cytoplasm and mitochondria. In later work that has
drawn much attention, they directed an improved sensor
to membrane rafts and caveolae and successfully investi-
gated the role of membrane raft integrity via cholesterol
depletion. This promising indicator could provide insight
into differences between membrane microdomains in
healthy and diseased states with good spatiotemporal
resolution [53]. Nikolaev et al. [54] exploited the cAMP-
binding domains of Epac and PKA to construct FRET-
based indicators without cooperativity, catalytic proper-
ties, of interactions with other proteins. The investigators
used these indicators to trace intracellular cAMP, and
found that cAMP signals traveled very fast throughout
hippocampal neurons and peritoneal macrophages. They
also used cAMP sensors to directly monitor the spatial
and temporal distribution of cAMP in adult cardiomy-
ocytes, and distinguished the cAMP signal diffusion
mediated by different adrenergic receptors.
Recently, Fan and colleagues [44, 55] optimized a nov-
el biosensor based on PKA and a circularly permuted form
of luciferase, and demonstrated its suitability for high-
throughput screening applications. Mazina et al. [56] con-
structed a simple and robust system for ligand screening
in a variety of mammalian cell lines utilizing viral infec-
tion. This system is applicable to high-throughput screen-
ing as a fast and dependable platform for drug discovery.
4.3 Visualization of intracellular cGMP signaling
The second messenger cyclic guanosine 5’-monophos-
phate (cGMP) participates in a variety of physiological
processes in mammals, especially in the nervous and vas-
cular systems [57–59]. These processes include mediation
of smooth muscle relaxation and modulation of synaptic
plasticity. cGMP functions by regulating effectors such as
cGMP-specific phosphodiesterases [60], cGMP-depend-
ent protein kinases (PKGs), and cyclic nucleotide-activat-
ed ion channels [61]. Understanding of the cGMP signal-
ing pathway has triggered great interest. Sato et al. [62]
described fluorescent indicators for cGMP in single living
cells that contained PKG I fused to FPs. Modified PKG I
(some amino acid residues were deleted in a dimerization
domain) was found to increase affinity upon cGMP bind-
ing and to respond reversibly to cGMP in nitric oxide
stimulated HEK293 cells. Honda et al. [63] increased the
selectivity for cGMP and eliminated the constitutive
kinase activity of the binding domain to diminish inter-
ference from the sensor. Using transfected Purkinje neu-
rons, they also confirmed that cGMP increases in cells
under conditions that induce synaptic plasticity.
To investigate the dynamics of cGMP in living cells,
Russwurm et al. [64] systematically studied the design of
FRET-based cGMP indicators and created indicators with
excellent specificity and rapid kinetics. The binding
affinities of these indicators ranged from 500 nM to 6 μM,
and radioimmunoassays were applied to validate their
performance. Nausch et al. [65] reported the design of a
cpFP-based cGMP biosensor to assess the dynamics of
intracellular cGMP synthesis, showing that such sensors
could act as innovative tools in elucidating the cGMP
dynamics. All indicators using different fragments of PKG
I displayed good selectivity, with high dynamic ranges
and apparent KDs.
4.4 Sensors for intracellular NADH detection
Reduced NADH and its oxidized form, NAD+, are the most
important cofactors in electron transfer metabolism. They
participate in multiple biological and pathological
processes, such as ischemia [66], energy metabolism [67],
tumor cell migration [68], and epilepsy [69]. Zhao et al. [70]
recently developed cpYFP-based sensors for NADH and
monitored the dynamic changes in NADH levels in the
organelles of mammalian cells. The NADH-binding
domain was mutated to expand the dynamic range, to
achieve high sensitivity with minimal perturbation, and
to target the sensor to subcellular organelles. When this
probe was used to trace the transport of exogenous NADH
across the plasma membrane in different cell lines,
researchers found that the inhibitor of P2X7R did not play
a role in the transport of NADH across the plasma mem-
brane. Zhao et al. [70] also investigated the differences in
NADH concentration in response to environmental
changes in different subcellular compartments. Mito-
chondria tended to maintain NADH at a stable level. To
evaluate the cellular NADH-NAD+ redox state, Hung et al.
[71] used a circularly permuted GFP T-Sapphire-based
sensor to monitor NADH responses to Phosphatidylinosi-
tide 3-kinase pathway inhibition. The fluorescence signal
of the sensor was calibrated with exogenous lactate and
pyruvate to measure cytosolic NADH:NAD+ ratios in vivo
and in vitro.
4.5 Optical measurement of amino acids
Glutamate plays a vital role in amino acid metabolism
[72], regulating not only nitrogen circulation together with
glutamine and 2-oxoglutarate [73], but also affecting sig-
nal transduction [74, 75]. Glutamate is the predominant
excitatory neurotransmitter in the mammalian brain.
Okumoto and colleagues [76, 77] developed a FRET-based
indicator for the detection of glutamate release from neu-
rons and glutamate imaging in brain slices. Site-directed
mutagenesis of the binding pocket was used to generate
mutants with binding affinities covering physiological
glutamate concentrations. Hires et al. [78] also quantita-
tively measured synaptic glutamate release with centi-
second temporal and dendritic spine-sized spatial resolu-
tion. With systematic optimization of linkers and gluta-
mate binding affinities, the improved sensor exhibits a
1285
Biotechnol. J. 2013, 8, 1280–1291
www.biotecvisions.com
© 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
6.2-fold increase in response magnitude over that of the
original. Thus, the sensor may be useful as a calibration
tool in tracing the propagation of glutamate, mapping the
functional connectivity of the brain, or in drug screening
for cerebral ischemia. Gruenwald et al. [79] studied anoth-
er critical amino acid, glutamine, and estimated the glut-
amine concentration in cells using sensors with different
affinities. Site-directed mutagenesis was performed to
create sensors with different affinities to facilitate physi-
ological glutamine analysis. These sensors can be easily
used to confirm the properties of glutamine transporters.
Many efforts have been made to enrich the family of sen-
sors for other amino acids. Okada et al. [80] used bacteri-
al periplasmic binding proteins (PBPs) to construct FRET-
based sensors, taking advantage of the structure of PBPs
to expand the dynamic range of the sensors by circular
permutation of the PBP module.
4.6 Fluorescent sensors for sugar detection
Sugars involved in key metabolic reactions need to be
monitored in real time to determine their subcellular dis-
tribution [81]. PBPs were utilized to develop the substrate-
binding element of sensors by linking an FP. These were
defined as fluorescence indicator protein (FLIP) family
sensors (also FRET-based sensors) because their hinge-
bend movement leads to highly responsive FRET. From-
mer and colleagues [32, 82–85] used PBPs to design a
series of FLIP sensors, and monitored the cytosolic distri-
bution of glucose, galactose, maltose, ribose, arabinose,
and sucrose in vivo. Glucose/galactose-binding protein
for glucose and galactose detection [82], maltose-binding
protein (for maltose detection [83], ribose-binding protein
for ribose detection [84], sucrose-binding protein for
sucrose detection [85], and L-arabinose-binding protein
for arabinose detection have been used as PBPs to devel-
op FLIP biosensors in which the bacterial PBPs are
flanked with GFP variants [32]. These FLIP sensors have
been improved and optimized to yield large dynamic
ranges when binding to target sugar molecules [83].
Effectively monitoring the cytosolic and intracellular dis-
tribution and concentration of sugars in real time will
shed light on their flux during metabolism. For example,
the FLIPGlu detection experiments showed that the con-
centration of cytosolic glucose fluctuates by several
orders of magnitude depending on the external glucose
supply [82]. FLIPmal, on the other hand, paved a way for
a better understanding of the transport processes within
and between cells. The determination of sucrose (as an
energy and carbon skeleton supplier in non-photosyn-
thetic organs of plants) and maltose concentrations by
this sensor in organelles such as chloroplasts character-
ized the actual function of transporters from correspon-
ding compartments [85]. Since the cytosolic and intercel-
lular sugar concentration and distribution directly relate
to carbohydrate metabolism, FLIP sensors additionally
have potential applications for fermentation processes in
the food, pharmaceutical, and biofuel industries. Sensors
for arabinose and maltose were used to determine the
concentrations of intracellular ligands in bacterial cul-
tures, making it possible to calculate accumulation rates
after the addition of the metabolite [86]. The use of PBP-
based sensors in metabolite flux research and bioprocess
visualization has progressed, but novel and optimized
sensors for other important sugar metabolites remain to
be developed.
4.7 Fluorescent sensors for redox state
Changes in ROS and, by extension, in cellular or intercel-
lular redox equilibrium are important for regulating a wide
range of physiological and pathological functions. Oster-
gaard and colleagues developed a redox-sensitive GFP
with two surface-exposed cysteines close to the chro-
mophore for real-time visualization of the molecule’s own
redox state [87–89]. Subsequently, a high oxidative
response sensor based on a redox-sensitive mutant of YFP
(rxYFP) was developed [90]. In other cases, sensors were
designed by flanking a polypeptide, acting as redox-sen-
sitive linker or redox-sensitive switch, with CFP/YFP,
leading to a FRET signal change in response to a different
redox environment [37]. These sensors depend on ROS-
induced intramolecular formation of disulfide bonds,
which lead to reversible conformational changes and
enable the imaging of ROS-induced oxidoreductase reac-
tion processes and signal transduction. HyPer, developed
by Belousov et al. [91], is an example of an ROS (mainly
H2O2) response domain-based sensor. HyPer consists of a
sensing domain, derived from Escherichia coli OxyR, with
two redox-active cysteines and cpYFP. The single-site
mutation A406V in HyPer (HyPer-2, A233V in OxyR)
expands the dynamic range of the probe up to
6-fold, thereby improving the detection of H2O2 levels in
the cytosol and peroxisome of tobacco and Arabidopsis
[92]. Sakai and colleagues [93, 94] engineered a novel
FRET-based redox sensor, named Redoxfluor, by fusing
the GFP variants Cerulean and Citrine to the N terminus
and C terminus, respectively, of the cysteine-rich domain
(I601 to N650) of sensory Yap1. This multifunctional sen-
sor has been used to detect peroxisomal and cytosolic
redox states in wild-type and mutant cells, screen for
drugs that modulate abnormal cytosol redox state in CHO
cells, and investigate redox state maintenance via cyto-
plasmic thioredoxin in the yeast Saccharomyces cerevisi-
ae. Similarly, Chen’s group designed a highly selective
and sensitive FRET-based sensor for organic hydroperox-
ides (OHPs), which constantly generate cellular stress, by
combining the responsive domain of the transcriptional
regulatory protein OhrR with an environmentally sensi-
tive FP, a modified Venus [15]. Moreover, ROS or redox
state probes could be designed based on redox regulato-
ry systems from different species to monitor the response
Biotechnology
Journal
Biotechnol. J. 2013, 8, 1280–1291
www.biotechnology-journal.com
1286 © 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
processes of redox systems per se, and further define the
mechanisms of these regulatory networks. More conven-
ient and effective probes still need to be mined for com-
plex mechanism research.
4.8 Other significant fluorescent sensors
Traditionally, carbon monoxide (CO) is viewed as a toxi-
cant or a pollutant. However, a recent study revealed that
CO, like NO, functions as an essential second messenger
[95]. Wang et al. [17] constructed the fluorescent probe
COSer for CO by fusing cpYFP to a dimeric CO-sensitive
heme protein from Rhodospirillum rubrum, CooA, as a CO
recognition site. However, further studies are needed to
improve the properties of CO probes used in biological
systems and other settings, such as industrial facilities.
Sato et al. [96] have reported a novel cell-based indicator
to visualize NO release from living cells, made by com-
bining endogenously expressed guanylate cyclase with a
FRET-based cGMP indicator. The indicator was used to
visualize NO diffusion from single vascular endothelial
cells. It showed good sensitivity, selectivity, and potential
for deciphering NO dynamics in biological systems.
2-Oxogluatarate (2OG), a metabolite from the highly
conserved Krebs cycle, not only plays a critical role in
metabolism, but also acts as a signaling molecule in a
Table 1. A list of fluorescent biosensors for studying metabolism in vivo
Name Applied for Based on Sensory domain Donor/receptor Ref
ATeams ATP Single-chain FRET FoF1-ATP synthase MseCFP/ cpmVenus [46]
EAF Modified FoF1 synthase GFP/YFP [49]
Perceval [ATP]/[ADP] cpFP GlnK1 cpmVenus [47]
ICUE cAMP Single-chain FRET Epac1 ECFP/Citrine [52]
HCN2-camps HCN2 channel EYFP/ECFP [54]
22F BRET PKA /RIIβB Luciferase [44]
RII-CFP/C-YFP Bimolecular FRET PKA CFP/YFP [102]
cGMP Single-chain FRET cNMP-BD/GAF EYFP/ECFP [64]
FlincGs cpFP PKG I cpEGFP [65]
CGY Single-chain FRET PKG I ECFP/EYFP [62]
Cygnet-2.1 cNMP-BD ECFP/Citrine [63, 103, 104]
Frex/ FrexH NADH cpFP Rex cpYFP [70]
Peredox [NADH]/[NAD+] cpFP Rex cpT-Sapphire [71]
GluSnFR Glu Single-chain FRET Gltl ECFP/Citrine [78]
FLIPQ-TV3.0 Gln GlnH mTFP1/Venus [79]
OGsor 2OG GAF ECFP/EYFP [97]
Arg QBP ECFP/Citrine [105]
Laconic Lactate LldR mTFP/Venus [98]
FLIPsuc-4mu Sucrose Sugar-binding protein ECFP/EYFP [85]
FILP-HisJ His HisJ ECFP/Venus [80]
FLIPGlu Glucose/galactose GGBP CFP/YFP [82]
FLIPmal Maltose MBP ECFP/EYFP [83]
FLIPrib Ribose RBP ECFP/EYFP [84]
FLIPara Arabinose L-Arabinose-binding protein ECFP/EYFP [86]
roGFP1/roGFP2 ROS roFP Surface-exposed cysteine roGFP [89]
HyPer H2O2 cpFP OxyR cpYFP [91]
Hyper-2 Mutant OxyR [92]
CY-RL5 Redox states Single-chain FRET RL5 ECFP/EYFP [37]
Redoxfluor Yap1 c-CRD Cerulean/Citrine [93, 94]
OHSer OHPs cpFP Xc-OhrR cpVenus [15]
CLPY BAI-2 Single-chain FRET LuxP CFP/YFP [106, 107]
FLIP-CIT Citrate CitA Venus/CFP [29]
COSer CO cpFP CooA cpYFP [95]
FRET-MT NO Single-chain FRET hMTIIa ECFP/EYFP [108]
Piccell Cell-based indicator sGC CFP/YFP [96]
RNA-based ADP RNA-based ADP binding aptamer DFHBI [99, 100]
Adenosine fluorescence Ade binding aptamer
Guanine enlargement Gua binding aptamer
GTP GTP binding aptamer
SAM SAM binding aptamer
1287
Biotechnol. J. 2013, 8, 1280–1291
www.biotecvisions.com
© 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
selecting the targeted binding domains for signaling mol-
ecules, especially in small molecules, is difficult. Recent-
ly, fluorescent RNA has been generated as a novel genet-
ically encoded biosensor that mimics GFP. It is increas-
ingly used in various applications [99]. GFP contains a
three-residue fluorophore formation that produces the
green fluorescence, whereas native RNA lacks a fluo-
rophore element. Therefore, to confer a GFP-like ability to
RNA, Paige et al. [100] engineered a fluorophore-binding
RNA aptamer, termed Spinach, containing 3,5-difluoro-
4-hydroxybenzylidene imidazolinone (DFHBI), which
resembles the fluorophore in GFP, thereby yielding a dif-
ferent type of genetically encoded biosensor with GFP-
like brightness, photostability, and cellular compatibility.
Moreover, RNA-fluorophore complexes with emission
wavelengths spanning the entire visible spectrum have
been created using different optimized aptamers. These
RNA-based probes can be used widely to detect a variety
of small molecules, such as adenosine, ADP, SAM, gua-
nine, and GTP. DFHBI fluorescence is activated when the
targeted small molecule binds to a specific aptamer con-
nected to Spinach via an optimized stem (Fig. 4). RNA-
based sensors, unlike FP-based FRET sensors, produce
approximately 20-fold increases in fluorescence upon
metabolite binding. Thus, they are promising tools for
imaging various metabolites in vivo.
6 Conclusions
The discovery and artificial evolution of FPs has been a
prominent contribution to biology [101], making it possi-
ble to directly observe biochemical processes inside liv-
ing cells and organisms [6]. With developments in FPs and
variety of organisms. Environmental inorganic nitrogen is
reduced to ammonium by microorganisms, which boast
the metabolic pathways via conversion of 2OG to gluta-
mate and glutamine. Recently, Zhang et al. [97] developed
FRET-based biosensors for tracking 2OG in real-time; the
dynamic ranges of the sensors appeared identical to the
physiological range observed in E. coli. Citrate is another
important intermediate in catabolic pathways involving
glycolysis and lipid metabolism. Ewald et al. [29] have
described FRET-based citrate sensors. In an optimization
process, different peptide linkers were screened to
achieve a maximal change ratio, and residues in the cit-
rate-binding pocket were modified to construct sensors
with the proper affinity for the application. When
expressed in E. coli, the indicator showed that cells could
respond to a carbon source even after long-term starva-
tion. Lactate also plays significant roles in healthy tissues.
As an abnormal lactate level is associated with inflamma-
tion, hypoxia, ischemia, neurodegeneration, and cancer,
tracing lactate levels in cells has diagnostic and thera-
peutic applications. San Martin et al. [98] recently devel-
oped a lactate sensor that discriminates lactate flux in dif-
ferent cells and found that T98G glioma cells have a three
to fivefold higher rate of lactate production than normal
astrocytes, which is consistent with tumor metabolism.
All genetically encoded fluorescent biosensors construct-
ed to trace intracellular metabolites are listed in Table 1.
5 RNA-based fluorescent sensors
FP-based biosensors are powerful tools for visualizing cel-
lular activities, but the fusion of FPs may potentially inter-
fere with the function of sensory domains. Moreover,
Figure 4. RNA sensor structure and
imaging of SAM in E. coli. (A) The sensor
consists of Spinach (black), a transducer
(orange), and a target-binding aptamer
(blue). DFHBI (green) fluorescence is
activated when the sensor binds to the
target metabolite (purple). (B) Emission
spectra of the SAM sensor. (C) Imaging
shows distinct SAM accumulation pat-
terns. Some cells exhibit higher than
average (arrow) or slow (arrowhead)
increases in SAM, and one cell shows
increased and then decreased SAM
levels (double arrow).
Biotechnology
Journal
Biotechnol. J. 2013, 8, 1280–1291
www.biotechnology-journal.com
1288 © 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
design methods, genetically encoded fluorescent probes
have enabled scientists to decode the mechanisms of
intracellular signal transduction pathways and gather a
large amount of biological information from cellular sys-
tems. Genetically encoded fluorescent indicators present
a popular way to study metabolic processes, take advan-
tage of FRET technology, and allow noninvasive, spa-
tiotemporal monitoring of signaling molecules in vivo.
Future work will expand the range of detection ligands to
include molecules such as 2OG, ADP, NAD, and others,
and will emphasize the optimization of existing biosen-
sors [33]. Moreover, RNA-based fluorescent sensors will
provide a new platform for exploration in cells.
This study was supported by the China NSF (21276079,
21335003), SRFDP (no. 20120074110009), the Key Grant
Project (no. 313019) of the Chinese Ministry of Education,
and the Fundamental Research Funds for the Central Uni-
versities.
The authors declare no commercial or financial conflict of
interest.
7 References
[1] Morise, H., Shimomura, O., Johnson, F. H., Winant, J., Intermolecu-
lar energy transfer in the bioluminescent system of Aequorea. Bio-
chemistry 1974, 13, 2656–2662.
[2] Prasher, D. C., Eckenrode, V. K., Ward, W. W., Prendergast, F. G.,
Cormier, M. J., Primary structure of the Aequorea victoria green-flu-
orescent protein. Gene 1992, 111, 229–233.
[3] Tsien, R. Y., The green fluorescent protein. Annu. Rev. Biochem.
1998, 67, 509–544.
[4] Cody, C. W., Prasher, D. C., Westler, W. M., Prendergast, F. G., Ward,
W. W., Chemical structure of the hexapeptide chromophore of the
Aequorea green-fluorescent protein. Biochemistry 1993, 32, 1212–
1218.
[5] Ormo, M., Cubitt, A. B., Kallio, K., Gross, L. A. et al., Crystal struc-
ture of the Aequorea victoria green fluorescent protein. Science
1996, 273, 1392–1395.
[6] Day, R. N., Davidson, M. W., The Fluorescent protein palette: Tools
for cellular imaging. Chem. Soc. Rev. 2009, 38, 2887–2921
[7] Baird, G. S., Zacharias, D. A., Tsien, R. Y., Biochemistry, mutagene-
sis, and oligomerization of DsRed, a red fluorescent protein from
coral. Proc. Natl. Acad. Sci. USA 2000, 97, 11984–11989.
[8] Wang, L., Jackson, W. C., Steinbach, P. A., Tsien, R. Y., Evolution of
new nonantibody proteins via iterative somatic hypermutation.
Proc. Natl. Acad. Sci. USA 2004, 101, 16745–16749.
[9] Wiedenmann, J., Ivanchenko, S., Oswald, F., Schmitt, F. et al., EosFP,
a fluorescent marker protein with UV-inducible green-to-red fluo-
rescence conversion. Proc. Natl. Acad. Sci. USA 2004, 101, 15905–
15910.
[10] Adam, V., Moeyaert, B., David, C. C., Mizuno, H. et al., Rational
design of photoconvertible and biphotochromic fluorescent proteins
for advanced microscopy applications. Chem. Biol. 2011, 18, 1241–
1251.
[11] Baird, G. S., Zacharias, D. A., Tsien, R. Y., Circular permutation and
receptor insertion within green fluorescent proteins. Proc. Natl.
Acad. Sci. USA 1999, 96, 11241–11246.
[12] Kawai, Y., Sato, M., Umezawa, Y., Single color fluorescent indicators
of protein phosphorylation for multicolor imaging of intracellular sig-
nal flow dynamics. Anal. Chem. 2004, 76, 6144–6149.
[13] Mizuno, T., Murao, K., Tanabe, Y., Oda, M., Tanaka, T., Metal-ion-
dependent GFP emission in vivo by combining a circularly permu-
tated green fluorescent protein with an engineered metal-ion-bind-
ing coiled-coil. J. Am. Chem. Soc. 2007, 129, 11378–11383.
[14] Leder, L., Stark, W., Freuler, F., Marsh, M. et al., The structure of Ca2+
sensor Case16 reveals the mechanism of reaction to low Ca2+ con-
centrations. Sensors (Basel) 2010, 10, 8143–8160.
Bang-Ce Ye is Professor of biology and
chemistry in the East China University
of Science and Technology in China,
and Director of the Laboratory of
Biosystems and Microanalysis in State
Key Laboratory of Bioreactor Engineer-
ing. He obtained his PhD degree in
1998. His research now focuses on
analytical biotechnology and systems
biotechnology.
Chang Zhang received his BSc in Life
Science in 2009 from the East China
University of Science and Technology
in China. He is pursuing a PhD in bio-
chemistry and molecular biology in the
Laboratory of Biosystems and Micro-
analysis group at the same institution.
His research is centered on tracking
intracellular metabolites in real time by
developing genetically encoded biosensors. His present work focuses
on the study of metabolites such as 2-oxogluatarate in tumor cells to
provide a better understanding of the regulatory pathways and signal-
ing networks for cancer.
Zihan Wei majored in biotechnology at
Jiangsu University (B.S. 2007) before
pursuing graduate work in the Labora-
tory of Biosystems and Microanalysis,
State Key Laboratory of Bioreactor Engi-
neering at the East China University of
Science and Technology. Under the
direction of Prof. Bang-Ce Ye, Zihan
is currently engineering genetically
encoded FRET-based biosensors for detecting metabolites such as
ROS and 2-oxoglutarate in Saccharopolyspora erythraea and characteriz-
ing its related metabolic pathways.
1289
Biotechnol. J. 2013, 8, 1280–1291
www.biotecvisions.com
© 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
[15] Zhao, B. S., Liang, Y., Song, Y., Zheng, C. et al., A highly selective
fluorescent probe for visualization of organic hydroperoxides in liv-
ing cells. J. Am. Chem. Soc. 2010, 132, 17065–17067.
[16] Schwarzlander, M., Logan, D. C., Fricker, M. D., Sweetlove, L. J., The
circularly permuted yellow fluorescent protein cpYFP that has been
used as a superoxide probe is highly responsive to pH but not super-
oxide in mitochondria: Implications for the existence of superoxide
‘flashes’. Biochem. J. 2011, 437, 381–387.
[17] Wang, J., Karpus, J., Zhao, B. S., Luo, Z. et al., A selective fluorescent
probe for carbon monoxide imaging in living cells. Angew. Chem.
Int. Ed. Engl. 2012, 51, 9652–9656.
[18] Liu, S., He, J., Jin, H., Yang, F. et al., Enhanced dynamic range in a
genetically encoded Ca2+ sensor. Biochem. Biophys. Res. Commun.
2011, 412, 155–159.
[19] Carlson, H. J., Cotton, D. W., Campbell, R. E., Circularly permuted
monomeric red fluorescent proteins with new termini in the beta-
sheet. Protein Sci. 2010, 19, 1490–1499.
[20] Shui, B., Wang, Q., Lee, F., Byrnes, L. J. et al., Circular permutation
of red fluorescent proteins. PLoS One 2011, 6, e20505.
[21] Magliery, T. J., Wilson, C. G., Pan, W., Mishler, D. et al., Detecting
protein-protein interactions with a green fluorescent protein frag-
ment reassembly trap: Scope and mechanism. J. Am. Chem. Soc.
2005, 127, 146–157.
[22] Ohad, N., Yalovsky, S., Utilizing bimolecular fluorescence comple-
mentation (BiFC) to assay protein-protein interaction in plants.
Methods Mol. Biol. 2010, 655, 347–358.
[23] Tian, G., Lu, Q., Zhang, L., Kohalmi, S. E., Cui, Y., Detection of pro-
tein interactions in plant using a gateway compatible bimolecular
fluorescence complementation (BiFC) system. J. Vis. Exp. 2011, (55).
pii: 3473.
[24] Shyu, Y. J., Suarez, C. D., Hu, C. D., Visualization of ternary com-
plexes in living cells by using a BiFC-based FRET assay. Nat. Protoc.
2008, 3, 1693–1702.
[25] Pfeiffer, D., Jendrossek, D., Development of a transferable bimolecu-
lar fluorescence complementation (BiFC) system for the investiga-
tion of interactions between poly(3-hydroxybutyrate) (PHB) granule-
associated proteins in Gram-negative bacteria. Appl. Environ.
Microbiol. 2013, 79, 2989–2999.
[26] Kodama, Y., Hu, C. D., Bimolecular fluorescence complementation
(BiFC) analysis of protein-protein interaction: How to calculate sig-
nal-to-noise ratio. Methods Cell Biol. 2013, 113, 107–121.
[27] Miranda, J. G., Weaver, A. L., Qin, Y., Park, J. G. et al., New alter-
nately colored FRET sensors for simultaneous monitoring of Zn(2)(+)
in multiple cellular locations. PLoS One 2012, 7, e49371.
[28] Salonikidis, P. S., Niebert, M., Ullrich, T., Bao, G. et al., An ion-insen-
sitive cAMP biosensor for long term quantitative ratiometric fluo-
rescence resonance energy transfer (FRET) measurements under
variable physiological conditions. J. Biol. Chem. 2011, 286, 23419–
23431.
[29] Ewald, J. C., Reich, S., Baumann, S., Frommer, W. B., Zamboni, N.,
Engineering genetically encoded nanosensors for real-time in vivo
measurements of citrate concentrations. PLoS One 2011, 6, e28245.
[30] de Lorimier, R. M., Smith, J. J., Dwyer, M. A., Looger, L. L. et al., Con-
struction of a fluorescent biosensor family. Protein Sci. 2002, 11,
2655–2675.
[31] Piljic, A., de Diego, I., Wilmanns, M., Schultz, C., Rapid development
of genetically encoded FRET reporters. ACS Chem. Biol. 2011, 6,
685–691.
[32] Deuschle, K., Okumoto, S., Fehr, M., Looger, L. L. et al., Construction
and optimization of a family of genetically encoded metabolite sen-
sors by semirational protein engineering. Protein Sci. 2005, 14,
2304–2314.
[33] Nagai, T., Yamada, S., Tominaga, T., Ichikawa, M., Miyawaki, A.,
Expanded dynamic range of fluorescent indicators for Ca(2+) by cir-
cularly permuted yellow fluorescent proteins. Proc. Natl. Acad. Sci.
USA 2004, 101, 10554–10559.
[34] Zaccolo, M., De Giorgi, F., Cho, C. Y., Feng, L. et al., A genetically
encoded, fluorescent indicator for cyclic AMP in living cells. Nat.
Cell Biol. 2000, 2, 25–29.
[35] van Dongen, E. M., Dekkers, L. M., Spijker, K., Meijer, E. W. et al.,
Ratiometric fluorescent sensor proteins with subnanomolar affinity
for Zn(II) based on copper chaperone domains. J. Am. Chem. Soc.
2006, 128, 10754–10762.
[36] Mori, M. X., Imai, Y., Itsuki, K., Inoue, R., Quantitative measurement
of Ca(2+)-dependent calmodulin-target binding by Fura-2 and CFP
and YFP FRET imaging in living cells. Biochemistry 2011, 50,
4685–4696.
[37] Kolossov, V. L., Spring, B. Q., Sokolowski, A., Conour, J. E. et al., Engi-
neering redox-sensitive linkers for genetically encoded FRET-based
biosensors. Exp. Biol. Med. (Maywood) 2008, 233, 238–248.
[38] John, S. A., Ottolia, M., Weiss, J. N., Ribalet, B., Dynamic modulation
of intracellular glucose imaged in single cells using a FRET-based
glucose nanosensor. Pflugers Arch. 2008, 456, 307–322.
[39] Achour, L., Kamal, M., Jockers, R., Marullo, S., Using quantitative
BRET to assess G protein-coupled receptor homo- and heterodimer-
ization. Methods Mol. Biol. 2011, 756, 183–200.
[40] Kocan, M., Pfleger, K. D., Study of GPCR-protein interactions by
BRET. Methods Mol. Biol. 2011, 746, 357–371.
[41] Xie, Q., Soutto, M., Xu, X., Zhang, Y., Johnson, C. H., Biolumines-
cence resonance energy transfer (BRET) imaging in plant seedlings
and mammalian cells. Methods Mol. Biol. 2011, 680, 3–28.
[42] Newman, R. H., Fosbrink, M. D., Zhang, J., Genetically encodable
fluorescent biosensors for tracking signaling dynamics in living
cells. Chem. Rev. 2011, 111, 3614–3666.
[43] Dragulescu-Andrasi, A., Chan, C. T., De, A., Massoud, T. F., Gamb-
hir, S. S., Bioluminescence resonance energy transfer (BRET) imag-
ing of protein-protein interactions within deep tissues of living sub-
jects. Proc. Natl. Acad. Sci. USA 2011, 108, 12060–12065.
[44] Binkowski, B. F., Butler, B. L., Stecha, P. F., Eggers, C. T. et al., A lumi-
nescent biosensor with increased dynamic range for intracellular
cAMP. ACS Chem. Biol. 2011, 6, 1193–1197.
[45] Pellegatti, P., Raffaghello, L., Bianchi, G., Piccardi, F. et al., Increased
level of extracellular ATP at tumor sites: In vivo imaging with plas-
ma membrane luciferase. PLoS One 2008, 3, e2599.
[46] Imamura, H., Nhat, K. P., Togawa, H., Saito, K. et al., Visualization of
ATP levels inside single living cells with fluorescence resonance
energy transfer-based genetically encoded indicators. Proc. Natl.
Acad. Sci. USA 2009, 106, 15651–15656.
[47] Berg, J., Hung, Y. P., Yellen, G., A genetically encoded fluorescent
reporter of ATP:ADP ratio. Nat. Methods 2009, 6, 161–166.
[48] Bell, C. J., Manfredi, G., Griffiths, E. J., Rutter, G. A., Luciferase
expression for ATP imaging: Application to cardiac myocytes. Meth-
ods Cell Biol. 2007, 80, 341–352.
[49] Zadran, S., Sanchez, D., Zadran, H., Amighi, A. et al., Enhanced-
acceptor fluorescence-based single cell ATP biosensor monitors
ATP in heterogeneous cancer populations in real time. Biotechnol.
Lett. 2013, 35, 175–180.
[50] Parks, S. K., Mazure, N. M., Counillon, L., Pouyssegur, J., Hypoxia
promotes tumor cell survival in acidic conditions by preserving ATP
levels. J. Cell Physiol. 2013.
[51] Ledur, P. F., Villodre, E. S., Paulus, R., Cruz, L. A. et al., Extracellular
ATP reduces tumor sphere growth and cancer stem cell population
in glioblastoma cells. Purinergic Signal. 2012, 8, 39–48.
[52] DiPilato, L. M., Cheng, X., Zhang, J., Fluorescent indicators of cAMP
and Epac activation reveal differential dynamics of cAMP signaling
within discrete subcellular compartments. Proc. Natl. Acad. Sci.
USA 2004, 101, 16513–16518.
Biotechnology
Journal
Biotechnol. J. 2013, 8, 1280–1291
www.biotechnology-journal.com
1290 © 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
[53] DiPilato, L. M., Zhang, J., The role of membrane microdomains in
shaping beta2-adrenergic receptor-mediated cAMP dynamics. Mol.
Biosyst. 2009, 5, 832–837.
[54] Nikolaev, V. O., Bunemann, M., Schmitteckert, E., Lohse, M. J.,
Engelhardt, S., Cyclic AMP imaging in adult cardiac myocytes
reveals far-reaching beta1-adrenergic but locally confined beta2-
adrenergic receptor-mediated signaling. Circ. Res. 2006, 99, 1084–
1091.
[55] Fan, F., Binkowski, B. F., Butler, B. L., Stecha, P. F. et al., Novel genet-
ically encoded biosensors using firefly luciferase. ACS Chem. Biol.
2008, 3, 346–351.
[56] Mazina, O., Reinart-Okugbeni, R., Kopanchuk, S., Rinken, A., Bac-
Mam system for FRET-based cAMP sensor expression in studies of
melanocortin MC1 receptor activation. J. Biomol. Screen. 2012, 17,
1096–1101.
[57] Rybalkin, S. D., Yan, C., Bornfeldt, K. E., Beavo, J. A., Cyclic GMP
phosphodiesterases and regulation of smooth muscle function. Circ.
Res. 2003, 93, 280–291.
[58] Munzel, T., Feil, R., Mulsch, A., Lohmann, S. M. et al., Physiology and
pathophysiology of vascular signaling controlled by guanosine
3’,5’-cyclic monophosphate-dependent protein kinase [corrected].
Circulation 2003, 108, 2172–2183.
[59] Feil, R., Lohmann, S. M., de Jonge, H., Walter, U., Hofmann, F., Cyclic
GMP-dependent protein kinases and the cardiovascular system:
Insights from genetically modified mice. Circ. Res. 2003, 93,
907–916.
[60] Conti, M., Beavo, J., Biochemistry and physiology of cyclic
nucleotide phosphodiesterases: Essential components in cyclic
nucleotide signaling. Annu. Rev. Biochem. 2007, 76, 481–511.
[61] Hofmann, F., Biel, M., Kaupp, U. B., International Union of Pharma-
cology. LI. Nomenclature and structure-function relationships of
cyclic nucleotide-regulated channels. Pharmacol. Rev. 2005, 57,
455–462.
[62] Sato, M., Hida, N., Ozawa, T., Umezawa, Y., Fluorescent indicators
for cyclic GMP based on cyclic GMP-dependent protein kinase Ial-
pha and green fluorescent proteins. Anal. Chem. 2000, 72, 5918–
5924.
[63] Honda, A., Adams, S. R., Sawyer, C. L., Lev-Ram, V. et al., Spa-
tiotemporal dynamics of guanosine 3’,5’-cyclic monophosphate
revealed by a genetically encoded, fluorescent indicator. Proc. Natl.
Acad. Sci. USA 2001, 98, 2437–2442.
[64] Russwurm, M., Mullershausen, F., Friebe, A., Jager, R. et al., Design
of fluorescence resonance energy transfer (FRET)-based cGMP indi-
cators: A systematic approach. Biochem. J. 2007, 407, 69–77.
[65] Nausch, L. W., Ledoux, J., Bonev, A. D., Nelson, M. T., Dostmann,
W. R., Differential patterning of cGMP in vascular smooth muscle
cells revealed by single GFP-linked biosensors. Proc. Natl. Acad. Sci.
USA 2008, 105, 365–370.
[66] Kannurpatti, S. S., Joshi, N. B., Energy metabolism and NAD-NADH
redox state in brain slices in response to glutamate exposure and
ischemia. Metab. Brain Dis. 1999, 14, 33–43.
[67] Li, Y., Dash, R. K., Kim, J., Saidel, G. M., Cabrera, M. E., Role of
NADH/NAD+ transport activity and glycogen store on skeletal mus-
cle energy metabolism during exercise: In silico studies. Am. J.
Physiol. Cell Physiol. 2009, 296, C25–46.
[68] Zhang, Q., Wang, S. Y., Nottke, A. C., Rocheleau, J. V. et al., Redox
sensor CtBP mediates hypoxia-induced tumor cell migration. Proc.
Natl. Acad. Sci. USA 2006, 103, 9029–9033.
[69] Garriga-Canut, M., Schoenike, B., Qazi, R., Bergendahl, K. et al.,
2-Deoxy-D-glucose reduces epilepsy progression by NRSF-CtBP-
dependent metabolic regulation of chromatin structure. Nat. Neu-
rosci. 2006, 9, 1382–1387.
[70] Zhao, Y., Jin, J., Hu, Q., Zhou, H. M. et al., Genetically encoded fluo-
rescent sensors for intracellular NADH detection. Cell Metab. 2011,
14, 555–566.
[71] Hung, Y. P., Albeck, J. G., Tantama, M., Yellen, G., Imaging cytosolic
NADH-NAD(+) redox state with a genetically encoded fluorescent
biosensor. Cell Metab. 2011, 14, 545–554.
[72] Forde, B. G., Lea, P. J., Glutamate in plants: Metabolism, regulation,
and signalling. J. Exp. Bot. 2007, 58, 2339–2358.
[73] Paczek, V., Dubois, F., Sangwan, R., Morot-Gaudry, J. F. et al., Cellu-
lar and subcellular localisation of glutamine synthetase and gluta-
mate dehydrogenase in grapes gives new insights on the regulation
of carbon and nitrogen metabolism. Planta 2002, 216, 245–254.
[74] Plaitakis, A., Zaganas, I., Regulation of human glutamate dehydro-
genases: Implications for glutamate, ammonia and energy metabo-
lism in brain. J. Neurosci. Res. 2001, 66, 899–908.
[75] Peng, S., Zhang, Y., Zhang, J., Wang, H., Ren, B., Glutamate recep-
tors and signal transduction in learning and memory. Mol. Biol. Rep.
2011, 38, 453–460.
[76] Dulla, C., Tani, H., Okumoto, S., Frommer, W. B. et al., Imaging of glu-
tamate in brain slices using FRET sensors. J. Neurosci. Methods
2008, 168, 306–319.
[77] Okumoto, S., Looger, L. L., Micheva, K. D., Reimer, R. J. et al., Detec-
tion of glutamate release from neurons by genetically encoded sur-
face-displayed FRET nanosensors. Proc. Natl. Acad. Sci. USA 2005,
102, 8740–8745.
[78] Hires, S. A., Zhu, Y., Tsien, R. Y., Optical measurement of synaptic
glutamate spillover and reuptake by linker optimized glutamate-sen-
sitive fluorescent reporters. Proc. Natl. Acad. Sci. USA 2008, 105,
4411–4416.
[79] Gruenwald, K., Holland, J. T., Stromberg, V., Ahmad, A. et al., Visu-
alization of glutamine transporter activities in living cells using
genetically encoded glutamine sensors. PLoS One 2012, 7, e38591.
[80] Okada, S., Ota, K., Ito, T., Circular permutation of ligand-binding
module improves dynamic range of genetically encoded FRET-
based nanosensor. Protein Sci. 2009, 18, 2518–2527.
[81] Ha, J. S., Song, J. J., Lee, Y. M., Kim, S. J. et al., Design and applica-
tion of highly responsive fluorescence resonance energy transfer
biosensors for detection of sugar in living Saccharomyces cerevisi-
ae cells. Appl. Environ. Microbiol. 2007, 73, 7408–7414.
[82] Fehr, M., Lalonde, S., Lager, I., Wolff, M. W., Frommer, W. B., In vivo
imaging of the dynamics of glucose uptake in the cytosol of COS-7
cells by fluorescent nanosensors. J. Biol. Chem. 2003, 278, 19127–
19133.
[83] Fehr, M., Frommer, W. B., Lalonde, S., Visualization of maltose
uptake in living yeast cells by fluorescent nanosensors. Proc. Natl.
Acad. Sci. USA 2002, 99, 9846–9851.
[84] Lager, I., Fehr, M., Frommer, W. B., Lalonde, S., Development of a flu-
orescent nanosensor for ribose. FEBS Lett. 2003, 553, 85–89.
[85] Lager, I., Looger, L. L., Hilpert, M., Lalonde, S., Frommer, W. B., Con-
version of a putative Agrobacterium sugar-binding protein into a
FRET sensor with high selectivity for sucrose. J. Biol. Chem. 2006,
281, 30875–30883.
[86] Kaper, T., Lager, I., Looger, L. L., Chermak, D., Frommer, W. B., Fluo-
rescence resonance energy transfer sensors for quantitative moni-
toring of pentose and disaccharide accumulation in bacteria.
Biotechnol. Biofuels 2008, 1, 11.
[87] Ostergaard, H., Henriksen, A., Hansen, F. G., Winther, J. R., Shed-
ding light on disulfide bond formation: Engineering a redox switch
in green fluorescent protein. EMBO J. 2001, 20, 5853–5862.
[88] Dooley, C. T., Dore, T. M., Hanson, G. T., Jackson, W. C. et al., Imag-
ing dynamic redox changes in mammalian cells with green fluores-
cent protein indicators. J. Biol. Chem. 2004, 279, 22284–22293.
[89] Hanson, G. T., Aggeler, R., Oglesbee, D., Cannon, M. et al., Investi-
gating mitochondrial redox potential with redox-sensitive green flu-
orescent protein indicators. J. Biol. Chem. 2004, 279, 13044–13053.
[90] Maulucci, G., Labate, V., Mele, M., Panieri, E. et al., High-resolution
imaging of redox signaling in live cells through an oxidation-sensi-
tive yellow fluorescent protein. Sci. Signal. 2008, 1, pl3.
1291
Biotechnol. J. 2013, 8, 1280–1291
www.biotecvisions.com
© 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
[91] Belousov, V. V., Fradkov, A. F., Lukyanov, K. A., Staroverov, D. B.
et al., Genetically encoded fluorescent indicator for intracellular
hydrogen peroxide. Nat. Methods 2006, 3, 281–286.
[92] Markvicheva, K. N., Bilan, D. S., Mishina, N. M., Gorokhovatsky, A.
Y. et al., A genetically encoded sensor for H2O2 with expanded
dynamic range. Bioorg. Med. Chem. 2011, 19, 1079–1084.
[93] Yano, T., Oku, M., Akeyama, N., Itoyama, A. et al., A novel fluores-
cent sensor protein for visualization of redox states in the cyto-
plasm and in peroxisomes. Mol. Cell. Biol. 2010, 30, 3758–3766.
[94] Oku, M., Hoseki, J., Ichiki, Y., Sakai, Y., A fluorescence resonance
energy transfer (FRET)-based redox sensor reveals physiological
role of thioredoxin in the yeast Saccharomyces cerevisiae. FEBS
Lett. 2013, 587, 793–798.
[95] Yuan, L., Lin, W., Tan, L., Zheng, K., Huang, W., Lighting up carbon
monoxide: Fluorescent probes for monitoring CO in living cells.
Angew. Chem. Int. Ed. Engl. 2013, 52, 1628–1630.
[96] Sato, M., Nakajima, T., Goto, M., Umezawa, Y., Cell-based indica-
tor to visualize picomolar dynamics of nitric oxide release from liv-
ing cells. Anal. Chem. 2006, 78, 8175–8182.
[97] Zhang, C., Wei, Z., Ye, B. C., Quantitative monitoring of 2-oxoglu-
tarate in Escherichia coli cells by a fluorescence resonance energy
transfer-based biosensor. Appl. Microbiol. Biotechnol. 2013, 97,
8307–8316.
[98] San Martin, A., Ceballo, S., Ruminot, I., Lerchundi, R. et al.,
A genetically encoded FRET lactate sensor and its use to detect the
Warburg effect in single cancer cells. PLoS One 2013, 8, e57712.
[99] Paige, J. S., Wu, K. Y., Jaffrey, S. R., RNA mimics of green fluores-
cent protein. Science 2011, 333, 642–646.
[100] Paige, J. S., Nguyen-Duc, T., Song, W., Jaffrey, S. R., Fluorescence
imaging of cellular metabolites with RNA. Science 2012, 335, 1194.
[101] Tsien, R. Y., Constructing and exploiting the fluorescent protein
paintbox (Nobel Lecture). Angew. Chem. Int. Ed. Engl. 2009, 48,
5612–5626.
[102] Zaccolo, M., Pozzan, T., Discrete microdomains with high concen-
tration of cAMP in stimulated rat neonatal cardiac myocytes. Sci-
ence 2002, 295, 1711–1715.
[103] Honda, A., Sawyer, C. L., Cawley, S. M., Dostmann, W. R., Cygnets:
In vivo characterization of novel cGMP indicators and in vivo imag-
ing of intracellular cGMP. Methods Mol. Biol. 2005, 307, 27–43.
[104] Honda, A., Moosmeier, M. A., Dostmann, W. R., Membrane-per-
meable cygnets: Rapid cellular internalization of fluorescent
cGMP-indicators. Front. Biosci. 2005, 10, 1290–1301.
[105] Bogner, M., Ludewig, U., Visualization of arginine influx into plant
cells using a specific FRET-sensor. J. Fluoresc. 2007, 17, 350–360.
[106] Rajamani, S., Zhu, J., Pei, D., Sayre, R., A LuxP-FRET-based
reporter for the detection and quantification of AI-2 bacterial quo-
rum-sensing signal compounds. Biochemistry 2007, 46, 3990–3997.
[107] Rajamani, S., Sayre, R. T., A sensitive fluorescence reporter for
monitoring quorum sensing regulated protease production in Vib-
rio harveyi. J. Microbiol. Methods 2011, 84, 189–193.
[108] Pearce, L. L., Gandley, R. E., Han, W., Wasserloos, K. et al., Role of
metallothionein in nitric oxide signaling as revealed by a green flu-
orescent fusion protein. Proc. Natl. Acad. Sci. USA 2000, 97, 477–
482.
© 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.biotechnology-journal.com
Editorial: Biotechnology Journal in Asia –
the first AFOB special issue
Tai Hyun Park and George G.Q. Chen
http://dx.doi.org/10.1002/biot.201300415
Review
Organ-on-a-chip technology and microfluidic whole-body
models for pharmacokinetic drug toxicity screening
Jong Bum Lee and Jong Hwan Sung
http://dx.doi.org/10.1002/biot.201300086
Review
Microfluidic-integrated biosensors: Prospects for point-of-care
diagnostics
Suveen Kumar, Saurabh Kumar, Md. Azahar Ali,
Pinki Anand,Ved Varun Agrawal, Renu John, Sagar Maji
and Bansi D. Malhotra
http://dx.doi.org/10.1002/biot.201200386
Review
Imaging and tracing of intracellular metabolites utilizing
genetically encoded fluorescent biosensors
Chang Zhang, Zi-Han Wei and Bang-Ce Ye
http://dx.doi.org/10.1002/biot.201300001
Review
Applications of cell-free protein synthesis in synthetic biology:
Interfacing bio-machinery with synthetic environments
Kyung-Ho Lee and Dong-Myung Kim
http://dx.doi.org/10.1002/biot.201200385
Review
Current developments in high-throughput analysis
for microalgae cellular contents
Tsung-Hua Lee, Jo-Shu Chang and Hsiang-Yu Wang
http://dx.doi.org/10.1002/biot.201200391
Research Article
Heterologous prime-boost immunization regimens using
adenovirus vector and virus-like particles induce broadly
neutralizing antibodies against H5N1 avian influenza viruses
Shih-Chang Lin, Wen-Chun Liu,Yu-Fen Lin,
Yu-Hsuan Huang, Jin-Hwang Liu and Suh-Chin Wu
http://dx.doi.org/10.1002/biot.201300116
Research Article
Repetitive Arg-Gly-Asp peptide as a cell-stimulating agent
on electrospun poly(ε-caprolactone) scaffold for tissue
engineering
Pacharaporn Chaisri, Artit Chingsungnoen
and Sineenat Siri
http://dx.doi.org/10.1002/biot.201300191
Research Article
Global gene expression analysis of Saccharomyces cerevisiae
grown under redox potential-controlled very-high-gravity
conditions
Chen-Guang Liu,Yen-Han Lin and Feng-Wu Bai
http://dx.doi.org/10.1002/biot.201300127
Research Article
Automated formation of multicomponent-encapuslating
vesosomes using continuous flow microcentrifugation
Huisoo Jang, Peichi C. Hu, Sungho Jung, Won Young Kim,
Sun Min Kim, Noah Malmstadt and Tae-Joon Jeon
http://dx.doi.org/10.1002/biot.201200388
Research Article
Size and CT density of iodine-containing ethosomal vesicles
obtained by membrane extrusion: Potential for use as CT
contrast agents
Bomin Na, Byoung Wook Choi and Bumsang Kim
http://dx.doi.org/10.1002/biot.201300110
Research Article
Site-targeted non-viral gene delivery by direct DNA injection
into the pancreatic parenchyma and subsequent in vivo
electroporation in mice
Masahiro Sato, Emi Inada, Issei Saitoh, Masato Ohtsuka,
Shingo Nakamura,Takayuki Sakurai and
Satoshi Watanabe
http://dx.doi.org/10.1002/biot.201300169
Biotechnology Journal – list of articles published in the November 2013 issue.
This issue is the first official special issue of Biotechnology Journal with the Asian Federation of
Biotechnology (AFOB). In the cover, we use the grain as a symbol and analogy for biotechnology,
as both the grain and biotechnology are common and uniting factors that link the member
countries of the AFOB. Gran image credit: © Myimagine – Fotolia.com.
Systems & Synthetic Biology ·
Nanobiotech · Medicine
ISSN 1860-6768 · BJIOAM 8 (11) 1245–1364 (2013) · Vol. 8 · November 2013
11/2013
Microfluidics
Biosensors
Organ-on-chip
www.biotechnology-journal.com

More Related Content

What's hot

Cupid Peptides presentation wjr
Cupid Peptides presentation wjrCupid Peptides presentation wjr
Cupid Peptides presentation wjrCupid Peptides
 
Biotemplates from plant viruses and their uses
Biotemplates from plant viruses and their usesBiotemplates from plant viruses and their uses
Biotemplates from plant viruses and their usesBASAVARAJ TAKKALAKI
 
Bioinformatics, comparative genemics and proteomics
Bioinformatics, comparative genemics and proteomicsBioinformatics, comparative genemics and proteomics
Bioinformatics, comparative genemics and proteomicsjuancarlosrise
 
Beyond Metagenomics- Integration Of Complementary Approaches For The Study Of...
Beyond Metagenomics- Integration Of Complementary Approaches For The Study Of...Beyond Metagenomics- Integration Of Complementary Approaches For The Study Of...
Beyond Metagenomics- Integration Of Complementary Approaches For The Study Of...guest5368597
 
Shivaputra
ShivaputraShivaputra
ShivaputraSHVA5965
 
Transcriptomics and metabolomics
Transcriptomics and metabolomicsTranscriptomics and metabolomics
Transcriptomics and metabolomicsSukhjinder Singh
 
Techniques in proteomics
Techniques in proteomicsTechniques in proteomics
Techniques in proteomicsN Poorin
 
Na f activates map ks and induces apoptosis in odontoblast-like
Na f activates map ks and induces apoptosis in odontoblast-likeNa f activates map ks and induces apoptosis in odontoblast-like
Na f activates map ks and induces apoptosis in odontoblast-likeGanesh Murthi
 
Expanding the Genetic Code
Expanding the Genetic CodeExpanding the Genetic Code
Expanding the Genetic CodeAbinNazar1
 
Comparative genomics and proteomics
Comparative genomics and proteomicsComparative genomics and proteomics
Comparative genomics and proteomicsNikhil Aggarwal
 
Mechanisms for the implementation of genetic information
Mechanisms for the implementation of genetic informationMechanisms for the implementation of genetic information
Mechanisms for the implementation of genetic informationHimanshu Baba
 
Functional proteomics, methods and tools
Functional proteomics, methods and toolsFunctional proteomics, methods and tools
Functional proteomics, methods and toolsKAUSHAL SAHU
 
GBS: Genotyping by sequencing
GBS: Genotyping by sequencingGBS: Genotyping by sequencing
GBS: Genotyping by sequencingsampath perumal
 
Whole genome sequencing of Bacillus subtilis a gram positive organism
Whole genome sequencing of Bacillus subtilis a gram positive organismWhole genome sequencing of Bacillus subtilis a gram positive organism
Whole genome sequencing of Bacillus subtilis a gram positive organismAshajyothi Mushineni
 
SAGE (Serial analysis of Gene Expression)
SAGE (Serial analysis of Gene Expression)SAGE (Serial analysis of Gene Expression)
SAGE (Serial analysis of Gene Expression)talhakhat
 

What's hot (20)

Cupid Peptides presentation wjr
Cupid Peptides presentation wjrCupid Peptides presentation wjr
Cupid Peptides presentation wjr
 
Biotemplates from plant viruses and their uses
Biotemplates from plant viruses and their usesBiotemplates from plant viruses and their uses
Biotemplates from plant viruses and their uses
 
Bioinformatics, comparative genemics and proteomics
Bioinformatics, comparative genemics and proteomicsBioinformatics, comparative genemics and proteomics
Bioinformatics, comparative genemics and proteomics
 
Beyond Metagenomics- Integration Of Complementary Approaches For The Study Of...
Beyond Metagenomics- Integration Of Complementary Approaches For The Study Of...Beyond Metagenomics- Integration Of Complementary Approaches For The Study Of...
Beyond Metagenomics- Integration Of Complementary Approaches For The Study Of...
 
Shivaputra
ShivaputraShivaputra
Shivaputra
 
Transcriptomics and metabolomics
Transcriptomics and metabolomicsTranscriptomics and metabolomics
Transcriptomics and metabolomics
 
Molekulare biologie 2
Molekulare biologie 2Molekulare biologie 2
Molekulare biologie 2
 
Techniques in proteomics
Techniques in proteomicsTechniques in proteomics
Techniques in proteomics
 
proteomics
 proteomics proteomics
proteomics
 
Na f activates map ks and induces apoptosis in odontoblast-like
Na f activates map ks and induces apoptosis in odontoblast-likeNa f activates map ks and induces apoptosis in odontoblast-like
Na f activates map ks and induces apoptosis in odontoblast-like
 
Heinrich et al., 2010
Heinrich et al., 2010Heinrich et al., 2010
Heinrich et al., 2010
 
Expanding the Genetic Code
Expanding the Genetic CodeExpanding the Genetic Code
Expanding the Genetic Code
 
Comparative genomics and proteomics
Comparative genomics and proteomicsComparative genomics and proteomics
Comparative genomics and proteomics
 
Mechanisms for the implementation of genetic information
Mechanisms for the implementation of genetic informationMechanisms for the implementation of genetic information
Mechanisms for the implementation of genetic information
 
CRISPR REPORT
CRISPR REPORTCRISPR REPORT
CRISPR REPORT
 
Functional proteomics, methods and tools
Functional proteomics, methods and toolsFunctional proteomics, methods and tools
Functional proteomics, methods and tools
 
GBS: Genotyping by sequencing
GBS: Genotyping by sequencingGBS: Genotyping by sequencing
GBS: Genotyping by sequencing
 
Gene expression
Gene expressionGene expression
Gene expression
 
Whole genome sequencing of Bacillus subtilis a gram positive organism
Whole genome sequencing of Bacillus subtilis a gram positive organismWhole genome sequencing of Bacillus subtilis a gram positive organism
Whole genome sequencing of Bacillus subtilis a gram positive organism
 
SAGE (Serial analysis of Gene Expression)
SAGE (Serial analysis of Gene Expression)SAGE (Serial analysis of Gene Expression)
SAGE (Serial analysis of Gene Expression)
 

Similar to REV

mRNA Technology How To Achieve Targeted Delivery to Organs, Cells, Tumors.pdf
mRNA Technology How To Achieve Targeted Delivery to Organs, Cells, Tumors.pdfmRNA Technology How To Achieve Targeted Delivery to Organs, Cells, Tumors.pdf
mRNA Technology How To Achieve Targeted Delivery to Organs, Cells, Tumors.pdfDoriaFang
 
Biosensors and Bioelectr
Biosensors and Bioelectr Biosensors and Bioelectr
Biosensors and Bioelectr Charles Zhang
 
Aplicación de Transferencia de Energía por Resonancia de Bioluminiscencia (BR...
Aplicación de Transferencia de Energía por Resonancia de Bioluminiscencia (BR...Aplicación de Transferencia de Energía por Resonancia de Bioluminiscencia (BR...
Aplicación de Transferencia de Energía por Resonancia de Bioluminiscencia (BR...LeidyCorrea16
 
Green Fluorescent Protein notes.ppt
Green Fluorescent Protein notes.pptGreen Fluorescent Protein notes.ppt
Green Fluorescent Protein notes.pptKimEliakim1
 
Fluorescent proteins in current biology
Fluorescent proteins in current biologyFluorescent proteins in current biology
Fluorescent proteins in current biologySSA KPI
 
Appl Microbiol Biotechnol
Appl Microbiol Biotechnol Appl Microbiol Biotechnol
Appl Microbiol Biotechnol Charles Zhang
 
The Fabrication And Modification Of T Cuas With Cellulose...
The Fabrication And Modification Of T Cuas With Cellulose...The Fabrication And Modification Of T Cuas With Cellulose...
The Fabrication And Modification Of T Cuas With Cellulose...Christy Hunt
 
CellAura Technologies Fluorescent Ligand User Group Programme
CellAura Technologies Fluorescent Ligand User Group ProgrammeCellAura Technologies Fluorescent Ligand User Group Programme
CellAura Technologies Fluorescent Ligand User Group Programmerichardmiddleton
 
Green Fluorescent Protein kinase in a cell
Green Fluorescent Protein kinase in a cellGreen Fluorescent Protein kinase in a cell
Green Fluorescent Protein kinase in a cellKimEliakim1
 
Green Fluorescent Protein.ppt
Green Fluorescent Protein.pptGreen Fluorescent Protein.ppt
Green Fluorescent Protein.pptKimEliakim1
 
Genetic Dna And Bioinformatics ( Accession No. Xp Essay
Genetic Dna And Bioinformatics ( Accession No. Xp EssayGenetic Dna And Bioinformatics ( Accession No. Xp Essay
Genetic Dna And Bioinformatics ( Accession No. Xp EssayJessica Deakin
 
Antimicrobial Activity Of Human Prion Protein Is Mediated By Its N Terminal R...
Antimicrobial Activity Of Human Prion Protein Is Mediated By Its N Terminal R...Antimicrobial Activity Of Human Prion Protein Is Mediated By Its N Terminal R...
Antimicrobial Activity Of Human Prion Protein Is Mediated By Its N Terminal R...Alfonso Enrique Islas Rodríguez
 
Environmental Factor - July 2014_ Intramural papers of the month
Environmental Factor - July 2014_ Intramural papers of the monthEnvironmental Factor - July 2014_ Intramural papers of the month
Environmental Factor - July 2014_ Intramural papers of the monthXunhai 郑训海
 
Target identification and validation in drug discovery
Target identification and validation in drug discoveryTarget identification and validation in drug discovery
Target identification and validation in drug discoverySpringer
 
IRJET- Subcellular Localization of Transmembrane E-cadherin-GFP Fusion Pr...
IRJET-  	  Subcellular Localization of Transmembrane E-cadherin-GFP Fusion Pr...IRJET-  	  Subcellular Localization of Transmembrane E-cadherin-GFP Fusion Pr...
IRJET- Subcellular Localization of Transmembrane E-cadherin-GFP Fusion Pr...IRJET Journal
 
JBEI highlights September 2019
JBEI highlights September 2019JBEI highlights September 2019
JBEI highlights September 2019LeahFreemanSloan
 

Similar to REV (20)

mRNA Technology How To Achieve Targeted Delivery to Organs, Cells, Tumors.pdf
mRNA Technology How To Achieve Targeted Delivery to Organs, Cells, Tumors.pdfmRNA Technology How To Achieve Targeted Delivery to Organs, Cells, Tumors.pdf
mRNA Technology How To Achieve Targeted Delivery to Organs, Cells, Tumors.pdf
 
Biosensors and Bioelectr
Biosensors and Bioelectr Biosensors and Bioelectr
Biosensors and Bioelectr
 
Aplicación de Transferencia de Energía por Resonancia de Bioluminiscencia (BR...
Aplicación de Transferencia de Energía por Resonancia de Bioluminiscencia (BR...Aplicación de Transferencia de Energía por Resonancia de Bioluminiscencia (BR...
Aplicación de Transferencia de Energía por Resonancia de Bioluminiscencia (BR...
 
Green Fluorescent Protein notes.ppt
Green Fluorescent Protein notes.pptGreen Fluorescent Protein notes.ppt
Green Fluorescent Protein notes.ppt
 
Presentation1
Presentation1Presentation1
Presentation1
 
Fluorescent proteins in current biology
Fluorescent proteins in current biologyFluorescent proteins in current biology
Fluorescent proteins in current biology
 
Appl Microbiol Biotechnol
Appl Microbiol Biotechnol Appl Microbiol Biotechnol
Appl Microbiol Biotechnol
 
The Fabrication And Modification Of T Cuas With Cellulose...
The Fabrication And Modification Of T Cuas With Cellulose...The Fabrication And Modification Of T Cuas With Cellulose...
The Fabrication And Modification Of T Cuas With Cellulose...
 
CellAura Technologies Fluorescent Ligand User Group Programme
CellAura Technologies Fluorescent Ligand User Group ProgrammeCellAura Technologies Fluorescent Ligand User Group Programme
CellAura Technologies Fluorescent Ligand User Group Programme
 
Green Fluorescent Protein kinase in a cell
Green Fluorescent Protein kinase in a cellGreen Fluorescent Protein kinase in a cell
Green Fluorescent Protein kinase in a cell
 
Green Fluorescent Protein.ppt
Green Fluorescent Protein.pptGreen Fluorescent Protein.ppt
Green Fluorescent Protein.ppt
 
FINAL BIOCHEM PAPER
FINAL BIOCHEM PAPERFINAL BIOCHEM PAPER
FINAL BIOCHEM PAPER
 
Genetic Dna And Bioinformatics ( Accession No. Xp Essay
Genetic Dna And Bioinformatics ( Accession No. Xp EssayGenetic Dna And Bioinformatics ( Accession No. Xp Essay
Genetic Dna And Bioinformatics ( Accession No. Xp Essay
 
Pglo Essay
Pglo EssayPglo Essay
Pglo Essay
 
Antimicrobial Activity Of Human Prion Protein Is Mediated By Its N Terminal R...
Antimicrobial Activity Of Human Prion Protein Is Mediated By Its N Terminal R...Antimicrobial Activity Of Human Prion Protein Is Mediated By Its N Terminal R...
Antimicrobial Activity Of Human Prion Protein Is Mediated By Its N Terminal R...
 
Environmental Factor - July 2014_ Intramural papers of the month
Environmental Factor - July 2014_ Intramural papers of the monthEnvironmental Factor - July 2014_ Intramural papers of the month
Environmental Factor - July 2014_ Intramural papers of the month
 
Target identification and validation in drug discovery
Target identification and validation in drug discoveryTarget identification and validation in drug discovery
Target identification and validation in drug discovery
 
IRJET- Subcellular Localization of Transmembrane E-cadherin-GFP Fusion Pr...
IRJET-  	  Subcellular Localization of Transmembrane E-cadherin-GFP Fusion Pr...IRJET-  	  Subcellular Localization of Transmembrane E-cadherin-GFP Fusion Pr...
IRJET- Subcellular Localization of Transmembrane E-cadherin-GFP Fusion Pr...
 
Grindberg - PNAS
Grindberg - PNASGrindberg - PNAS
Grindberg - PNAS
 
JBEI highlights September 2019
JBEI highlights September 2019JBEI highlights September 2019
JBEI highlights September 2019
 

REV

  • 1. Biotechnology Journal DOI 10.1002/biot.201300001 Biotechnol. J. 2013, 8, 1280–1291 1280 © 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 1 Introduction Cell metabolism boasts a complicated network. Exploring the tremendous amount of information contained within this network is of great significance to related studies. To be able to visualize and monitor cell activities in vivo with high spatial and temporal resolution is attractive because traditional analysis methods destroy living cells to extract metabolites of interest. Since the discovery of green fluo- rescent protein (GFP), scientists have used site-directed and random mutagenesis approaches to develop fluores- cent protein (FP) mutants, and have created a family of proteins that almost span the fluorescence spectrum. This family has enlightened efforts to construct genetically encoded fluorescent biosensors based on protein/protein interactions and intramolecular conformational changes targeted to living cells or tissues. Recently, another sort of genetically encoded biosensor, the RNA-based biosensor, has been developed for tracking small-molecule metabo- lites such as adenosine, ADP, S-adenosyl methionine (SAM), guanine, and GTP, and shows more advantages than FP-based biosensors. Here, we describe different modes of constructing fluorescent biosensors and review recent advances in the use of genetically encoded fluo- rescent biosensors for tracing intracellular metabolism. 2 Fluorescent proteins: Gifts from the ocean When a bright, greenish protein in jellyfish extracts was observed by Shimomura in the 1960s [1], the glow of FPs Review Imaging and tracing of intracellular metabolites utilizing genetically encoded fluorescent biosensors Chang Zhang*, Zi-Han Wei* and Bang-Ce Ye Laboratory of Biosystems and Microanalysis, State Key Laboratory of Bioreactor Engineering, East China University of Science and Technology, Shanghai, China Intracellular metabolites play a crucial role in characterizing and regulating corresponding cellular activities. Tracking intracellular metabolites in real time by traditional means was difficult until the powerful toolkit of genetically encoded biosensors was developed. Over the past few decades, iter- ative improvements of these biosensors have been made, resulting in the effective monitoring of metabolites. In this review, we introduce and discuss the recent advances in the use of genetical- ly encoded biosensors for tracking some key metabolites, such as ATP, cAMP, cGMP, NADH, reac- tive oxygen species, sugar, carbon monoxide, and nitric oxide. A brief phylogeny of fluorescent pro- teins and several typical construction modes for genetically encoded biosensors are also described. We also discuss the development of novel RNA-based sensors, which are genetically encoded biosensors active at the transcriptional level. Keywords: Genetically encoded fluorescent biosensors · In vivo imaging · Metabolite Correspondence: Prof. Bang-Ce Ye, Lab of Biosystems and Microanalysis, State Key Laboratory of Bioreactor Engineering, East China University of Science and Technology, Meilong RD 130, Shanghai 200237, China E-mail: bcye@ecust.edu.cn Abbreviations: 2OG, 2-Oxogluatarate; BRET, bioluminescence resonance energy transfer; cAMP, cyclic adenosine monophosphate; cGMP, cyclic guanosine 5’-monophosphate; CFP, cyan fluorescent protein; cpFP, circu- lar permutation fluorescent protein; DFHBI, 3,5-difluoro-4-hydroxybenzyli- dene imidazolinone; FLIP, fluorescence indicator protein; FP, fluorescent protein; FRET, fluorescence resonance energy transfer; GFP, green fluores- cent protein; GPCR, G-protein coupled receptor; PBP, periplasmic binding protein; PKA, cAMP-dependent protein kinase; PKG, cGMP-dependent pro- tein kinase; OHP, organic hydroperoxide; RFP, red fluorescent protein; ROS, reactive oxygen species; SAM, S-adenosyl methionine; YFP, yellow fluorescent protein Received 31 MAR 2013 Revised 02 AUG 2013 Accepted 26 AUG 2013 Supporting information available online * These authors contributed equally to this work. Color online: refer to online PDF file for figures in color.
  • 2. 1281 Biotechnol. J. 2013, 8, 1280–1291 www.biotecvisions.com © 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim provided the ability to light up the invisible metabolism in a living life form. In 1992, Prasher and colleagues [2] used a cDNA library from Aequorea victoria to clone the gene encoding the aequorin companion protein, GFP. They demonstrated GFP’s promising future, a gift for cellular biologists: genetically encoded GFP could be expressed in a variety of biological systems without additional factors or external enzyme components. Stable chromophore for- mation within a cylinder structure makes GFP a powerful cellular reporter [3]. A hexapeptide sequence starting at amino acid 64 (number indicates the position in the intact peptide sequence) determines the light absorption prop- erties of GFP. The adjacent Ser65-Tyr66-Gly67 sequence forms the chromophore, and fluorescence results from the oxidization of the Tyr66 a–b carbon bond [4]. The crystal structure for GFP shows the cyclic tripeptide chro- mophore buried in the center of an 11-stranded β-barrel cylinder structure [5]. In further studies of the cylinder structure, mutagenesis of the amino acid residues sur- rounding the chromophore significantly affected the spectral properties of the protein. A series of FPs, with emission wavelengths ranging from the blue to yellow regions of the visible spectrum, and “enhanced” (E) FPs, with increased protein folding efficiency and maturation at physiological temperature (Supporting information, Table S1), were produced [6]. Over the past 15 years, Anthozoa-based FPs spanning the visible spectrum have been characterized and opti- mized for imaging applications [6]. DsRed-based variants and the mFruit (“m” for monomer) series are two repre- sentative FP families with longer wavelength emissions [7]. Members of the mFruit family, including mHoneydew, mBanana, mOrange, mTangerine, mStrawberry, mCherry, mApple, mPlum, and dTomato (“d” for dimer) derive from monomeric red FP (RFP) 1, a mutant of DsRed (33 amino acid substitutions) [8]. They have been useful in multi- color imaging experiments. EosFP, another type of pho- toactivatable FP, from the reef coral Lobophyllia hem- prichii can be photoconverted from green to red fluores- cence by near-ultraviolet light [9, 10]. Mutagenesis was used to develop a monomeric protein named mEosFP. An improved form, mIrisFP, enables pulse-chase experiments with superb resolution and provides an excellent fusion marker for imaging in living cells. Figure 1. Design methods for construct- ing single fluorophore sensors. The units of a dimeric binding domain can be directly attached to the original N and C termini of an FP (A) or to novel N and C termini of a cpFP (C). A monomeric binding domain can be inserted into an FP (B) or a cpFP (D). The binding or effective domain undergoes a structural change upon substrate (red star) bind- ing or in response to other influencing factors, such as ROS, leading to a change in the fluorescent signal of the sensor.
  • 3. Biotechnology Journal Biotechnol. J. 2013, 8, 1280–1291 www.biotechnology-journal.com 1282 © 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 3 Design modes for genetically encoded fluorescent biosensors 3.1 Single fluorophore sensors Proteins can tolerate circular permutation, which pro- duces novel N- and C-termini from different portions of the protein, while maintaining a stable structure. Tsien and coworkers [11], researching circular permutation and receptor insertion within GFPs, showed that GFP is sur- prisingly robust for circular permutation and insertion, and offered a new strategy for creating genetically encod- ed indicators [12–14]. Circular permutation FP (cpFP) is especially attractive because the binding domain of inter- est undergoes a weak structural change upon substrate binding, and the signal change is amplified when the domain is placed in the sensitive region of the chro- mophore. Depending on the structure of the receptor, such as the distance between N and C termini, oligomer- ization, etc., researchers can choose a proper design method for constructing cpFP-based sensors (Fig. 1). Giv- en the structural similarity between GFP and its mutants, many active cpFP have been produced using yellow FP (YFP) [15–17], Venus [18], and RFP [19, 20]. Recently, cpYFP-based sensors for NADH were developed by inserting cpYFP into dimeric NADH binding domain Rex as pictured in Fig. 1C. In addition to circular permutation, GFP family mem- bers are also amenable to protein fragmentation [21]. When the fragments of an FP are brought into proximity with each other, they can reunite to form a functional protein (Fig.  2). This method, termed bimolecular fluorescence complementation, has been exploited to detect protein- protein interactions in living cells and plants [22–25]. A great advantage of this strategy is its high sensitivity, since GFP fragments in the separated state only display a modest background fluorescence until they are brought closer by protein-protein interactions [26]. 3.2 Fluorescence resonance energy transfer-based reporter systems Fluorescence resonance energy transfer (FRET) is used in the design of indicators for optically imaging biochemical and physiological functions in living cells [27, 28]. A genetically encoded FRET biosensor consists of a recognition module, which specifically binds a ligand, sandwiched between two variants of GFP – typically cyan FP (CFP) and YFP [29, 30]. The efficiency of fluorescence energy transfer between the two fluorophores is highly dependent on their distance and orientation [31–33]. Tra- ditional FRET sensors can be divided into two classes, as described below. Bimolecular probes are designed based on protein interaction and are suitable for studying the dissociation or association of a protein upon ligand binding [34, 35]. Donor and acceptor FPs are fused to interacting proteins separately to form a pair of FRET sensors. When the inter- action of proteins X and Y brings the FRET sensors clos- er, the intermolecular FRET signal efficiency between donor and acceptor FPs is elevated (Fig. 3A). This mech- anism has been used to construct probes for measuring the cytosolic Ca2+ concentration and protein interactions within the same cell, using Fura-2 with super-enhanced CFP and YFP as a FRET pair [36]. Single-chain probes are designed based on a protein conformational change that occurs upon ligand binding or in response to another effector. Kolossov et al. [37] inserted redox-sensitive link- ers between CFP and YFP to engineer single-chain FRET sensors for reactive oxygen species (ROS) detection. FRET-based sensors are also suitable for detecting lig- ands of interest in vivo [38]. The selected binding domain is sandwiched between FP pairs, usually CFP and YFP. A conformational change in the domain upon ligand bind- ing likely leads to a change in the distance or orientation between the CFP and YFP chromophores, resulting in a FRET efficiency change (Fig. 3B). Optimization is impor- tant to obtain sensors with obvious FRET ratio changes or Figure 2. When the interacting proteins X and Y bind to each other, the seg- ments of a fluorescent protein are brought into proximity to form a func- tional protein with recovered fluorescent signal.
  • 4. 1283 Biotechnol. J. 2013, 8, 1280–1291 www.biotecvisions.com © 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim a dynamic range consistent with concentration level of the target molecules of interest in cell metabolism [31, 32]. Traditional bioluminescence resonance energy trans- fer (BRET)-based assays have also been used for detect- ing protein interactions [39–41]. In a BRET-based sensor, luciferase is usually used as the BRET donor and an FP as the BRET acceptor. However, using BRET to monitor pro- tein complexes at the subcellular level has been limited by the amount of light emission intrinsic to the luciferase donor [42]. Recently, Dragulescu-Andrasi et al. [43] demonstrated the use of red light-emitting BRET systems for investigating protein interactions in deep tissue and in small animal tumor models, which may help to eliminate this drawback. Binkowski et al. [44] employed a circularly permuted form of firefly luciferase to construct a lumines- cent biosensor for intracellular cyclic adenosine mono- phosphate (cAMP) detection. They studied signal trans- duction mediated by G-protein coupled receptors (GPCRs), and showed that luciferase is suitable for the construction of genetically encoded biosensors as cpFPs. 4 Biosensors for tracing metabolism 4.1 Visualization of ATP levels inside cells ATP is the major energy currency in cellular processes. Real-time monitoring of ATP levels inside individual living cells would help elucidate the compartmentation of ATP and its role in modulating ion channels and signaling cas- cades [45]. Imamura et al. [46] generated a series of FRET- based indicators for ATP, and showed that the ATP levels differed between the mitochondrial matrix and cyto- plasm/nucleus in HeLa cells. Additional research demon- strated that the ATP-generating pathway changed in response to nutritional changes in the environment. The probes used in these studies show high selectivity for ATP over other nucleotides. By modulating the affinity through substitution of different effector domains, or mutation of residues at the binding interface, the probes can measure ATP levels ranging from 2 μM to 8 mM. On attaching a proper signal sequence, these sensors could monitor any intracellular compartment of interest and facilitate biological research in other fields. Since the absolute levels of ATP, ADP, and AMP can fluctuate, the ratio of [ATP]/[ADP] is thought to be a more reliable indicator of energy status in cells. Berg et al. [47] presented a novel cpFP-based fluorescent probe for meas- uring the cellular ATP/ADP ratio. They measured the energy level by competitive binding of ATP and ADP in a cell without perturbing the cellular energy balance, in contrast to sensors based on luciferase [48]. The regula- tion of the metabolic machinery in cancer is complicated. A better understanding of differential cancer cell metab- olism would greatly benefit therapeutics. Zadran et al. [49] monitored cytosolic ATP in tumor cells with a sensor that could also be used in mitochondria, nuclei, and the endo- plasmic reticulum. These efforts will improve our under- standing of cellular and subcellular energy variations in normal and abnormal cells [50, 51]. 4.2 Fluorescent indicators of cAMP cAMP is a second messenger of many GPCRs and regu- lates cAMP-dependent protein kinase (PKA) and exchange proteins activated by cAMP (Epacs) to mediate cellular functions. Spatial and temporal readouts of cAMP levels are vital to the comprehension of network regula- tion in various signaling cascades. DiPilato et al. [52] con- structed fluorescent indicators to report intracellular Figure 3. Schematic presentation of (A) a bimolecular FRET probe and (B) a single-chain FRET probe. (A) When the interacting proteins X and Y bind to each other, the fluorescent pair is brought into proximity and a FRET phenomenon appears. (B) When the binding domain in a single-chain FRET probe undergoes conformational changes upon ligand binding, the distance or orientation between the fluorescent pair changes and then alters the FRET efficiency.
  • 5. Biotechnology Journal Biotechnol. J. 2013, 8, 1280–1291 www.biotechnology-journal.com 1284 © 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim cAMP dynamics and Epac activation. They revealed that the cAMP response at the membrane was faster than that in the cytoplasm and mitochondria. In later work that has drawn much attention, they directed an improved sensor to membrane rafts and caveolae and successfully investi- gated the role of membrane raft integrity via cholesterol depletion. This promising indicator could provide insight into differences between membrane microdomains in healthy and diseased states with good spatiotemporal resolution [53]. Nikolaev et al. [54] exploited the cAMP- binding domains of Epac and PKA to construct FRET- based indicators without cooperativity, catalytic proper- ties, of interactions with other proteins. The investigators used these indicators to trace intracellular cAMP, and found that cAMP signals traveled very fast throughout hippocampal neurons and peritoneal macrophages. They also used cAMP sensors to directly monitor the spatial and temporal distribution of cAMP in adult cardiomy- ocytes, and distinguished the cAMP signal diffusion mediated by different adrenergic receptors. Recently, Fan and colleagues [44, 55] optimized a nov- el biosensor based on PKA and a circularly permuted form of luciferase, and demonstrated its suitability for high- throughput screening applications. Mazina et al. [56] con- structed a simple and robust system for ligand screening in a variety of mammalian cell lines utilizing viral infec- tion. This system is applicable to high-throughput screen- ing as a fast and dependable platform for drug discovery. 4.3 Visualization of intracellular cGMP signaling The second messenger cyclic guanosine 5’-monophos- phate (cGMP) participates in a variety of physiological processes in mammals, especially in the nervous and vas- cular systems [57–59]. These processes include mediation of smooth muscle relaxation and modulation of synaptic plasticity. cGMP functions by regulating effectors such as cGMP-specific phosphodiesterases [60], cGMP-depend- ent protein kinases (PKGs), and cyclic nucleotide-activat- ed ion channels [61]. Understanding of the cGMP signal- ing pathway has triggered great interest. Sato et al. [62] described fluorescent indicators for cGMP in single living cells that contained PKG I fused to FPs. Modified PKG I (some amino acid residues were deleted in a dimerization domain) was found to increase affinity upon cGMP bind- ing and to respond reversibly to cGMP in nitric oxide stimulated HEK293 cells. Honda et al. [63] increased the selectivity for cGMP and eliminated the constitutive kinase activity of the binding domain to diminish inter- ference from the sensor. Using transfected Purkinje neu- rons, they also confirmed that cGMP increases in cells under conditions that induce synaptic plasticity. To investigate the dynamics of cGMP in living cells, Russwurm et al. [64] systematically studied the design of FRET-based cGMP indicators and created indicators with excellent specificity and rapid kinetics. The binding affinities of these indicators ranged from 500 nM to 6 μM, and radioimmunoassays were applied to validate their performance. Nausch et al. [65] reported the design of a cpFP-based cGMP biosensor to assess the dynamics of intracellular cGMP synthesis, showing that such sensors could act as innovative tools in elucidating the cGMP dynamics. All indicators using different fragments of PKG I displayed good selectivity, with high dynamic ranges and apparent KDs. 4.4 Sensors for intracellular NADH detection Reduced NADH and its oxidized form, NAD+, are the most important cofactors in electron transfer metabolism. They participate in multiple biological and pathological processes, such as ischemia [66], energy metabolism [67], tumor cell migration [68], and epilepsy [69]. Zhao et al. [70] recently developed cpYFP-based sensors for NADH and monitored the dynamic changes in NADH levels in the organelles of mammalian cells. The NADH-binding domain was mutated to expand the dynamic range, to achieve high sensitivity with minimal perturbation, and to target the sensor to subcellular organelles. When this probe was used to trace the transport of exogenous NADH across the plasma membrane in different cell lines, researchers found that the inhibitor of P2X7R did not play a role in the transport of NADH across the plasma mem- brane. Zhao et al. [70] also investigated the differences in NADH concentration in response to environmental changes in different subcellular compartments. Mito- chondria tended to maintain NADH at a stable level. To evaluate the cellular NADH-NAD+ redox state, Hung et al. [71] used a circularly permuted GFP T-Sapphire-based sensor to monitor NADH responses to Phosphatidylinosi- tide 3-kinase pathway inhibition. The fluorescence signal of the sensor was calibrated with exogenous lactate and pyruvate to measure cytosolic NADH:NAD+ ratios in vivo and in vitro. 4.5 Optical measurement of amino acids Glutamate plays a vital role in amino acid metabolism [72], regulating not only nitrogen circulation together with glutamine and 2-oxoglutarate [73], but also affecting sig- nal transduction [74, 75]. Glutamate is the predominant excitatory neurotransmitter in the mammalian brain. Okumoto and colleagues [76, 77] developed a FRET-based indicator for the detection of glutamate release from neu- rons and glutamate imaging in brain slices. Site-directed mutagenesis of the binding pocket was used to generate mutants with binding affinities covering physiological glutamate concentrations. Hires et al. [78] also quantita- tively measured synaptic glutamate release with centi- second temporal and dendritic spine-sized spatial resolu- tion. With systematic optimization of linkers and gluta- mate binding affinities, the improved sensor exhibits a
  • 6. 1285 Biotechnol. J. 2013, 8, 1280–1291 www.biotecvisions.com © 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 6.2-fold increase in response magnitude over that of the original. Thus, the sensor may be useful as a calibration tool in tracing the propagation of glutamate, mapping the functional connectivity of the brain, or in drug screening for cerebral ischemia. Gruenwald et al. [79] studied anoth- er critical amino acid, glutamine, and estimated the glut- amine concentration in cells using sensors with different affinities. Site-directed mutagenesis was performed to create sensors with different affinities to facilitate physi- ological glutamine analysis. These sensors can be easily used to confirm the properties of glutamine transporters. Many efforts have been made to enrich the family of sen- sors for other amino acids. Okada et al. [80] used bacteri- al periplasmic binding proteins (PBPs) to construct FRET- based sensors, taking advantage of the structure of PBPs to expand the dynamic range of the sensors by circular permutation of the PBP module. 4.6 Fluorescent sensors for sugar detection Sugars involved in key metabolic reactions need to be monitored in real time to determine their subcellular dis- tribution [81]. PBPs were utilized to develop the substrate- binding element of sensors by linking an FP. These were defined as fluorescence indicator protein (FLIP) family sensors (also FRET-based sensors) because their hinge- bend movement leads to highly responsive FRET. From- mer and colleagues [32, 82–85] used PBPs to design a series of FLIP sensors, and monitored the cytosolic distri- bution of glucose, galactose, maltose, ribose, arabinose, and sucrose in vivo. Glucose/galactose-binding protein for glucose and galactose detection [82], maltose-binding protein (for maltose detection [83], ribose-binding protein for ribose detection [84], sucrose-binding protein for sucrose detection [85], and L-arabinose-binding protein for arabinose detection have been used as PBPs to devel- op FLIP biosensors in which the bacterial PBPs are flanked with GFP variants [32]. These FLIP sensors have been improved and optimized to yield large dynamic ranges when binding to target sugar molecules [83]. Effectively monitoring the cytosolic and intracellular dis- tribution and concentration of sugars in real time will shed light on their flux during metabolism. For example, the FLIPGlu detection experiments showed that the con- centration of cytosolic glucose fluctuates by several orders of magnitude depending on the external glucose supply [82]. FLIPmal, on the other hand, paved a way for a better understanding of the transport processes within and between cells. The determination of sucrose (as an energy and carbon skeleton supplier in non-photosyn- thetic organs of plants) and maltose concentrations by this sensor in organelles such as chloroplasts character- ized the actual function of transporters from correspon- ding compartments [85]. Since the cytosolic and intercel- lular sugar concentration and distribution directly relate to carbohydrate metabolism, FLIP sensors additionally have potential applications for fermentation processes in the food, pharmaceutical, and biofuel industries. Sensors for arabinose and maltose were used to determine the concentrations of intracellular ligands in bacterial cul- tures, making it possible to calculate accumulation rates after the addition of the metabolite [86]. The use of PBP- based sensors in metabolite flux research and bioprocess visualization has progressed, but novel and optimized sensors for other important sugar metabolites remain to be developed. 4.7 Fluorescent sensors for redox state Changes in ROS and, by extension, in cellular or intercel- lular redox equilibrium are important for regulating a wide range of physiological and pathological functions. Oster- gaard and colleagues developed a redox-sensitive GFP with two surface-exposed cysteines close to the chro- mophore for real-time visualization of the molecule’s own redox state [87–89]. Subsequently, a high oxidative response sensor based on a redox-sensitive mutant of YFP (rxYFP) was developed [90]. In other cases, sensors were designed by flanking a polypeptide, acting as redox-sen- sitive linker or redox-sensitive switch, with CFP/YFP, leading to a FRET signal change in response to a different redox environment [37]. These sensors depend on ROS- induced intramolecular formation of disulfide bonds, which lead to reversible conformational changes and enable the imaging of ROS-induced oxidoreductase reac- tion processes and signal transduction. HyPer, developed by Belousov et al. [91], is an example of an ROS (mainly H2O2) response domain-based sensor. HyPer consists of a sensing domain, derived from Escherichia coli OxyR, with two redox-active cysteines and cpYFP. The single-site mutation A406V in HyPer (HyPer-2, A233V in OxyR) expands the dynamic range of the probe up to 6-fold, thereby improving the detection of H2O2 levels in the cytosol and peroxisome of tobacco and Arabidopsis [92]. Sakai and colleagues [93, 94] engineered a novel FRET-based redox sensor, named Redoxfluor, by fusing the GFP variants Cerulean and Citrine to the N terminus and C terminus, respectively, of the cysteine-rich domain (I601 to N650) of sensory Yap1. This multifunctional sen- sor has been used to detect peroxisomal and cytosolic redox states in wild-type and mutant cells, screen for drugs that modulate abnormal cytosol redox state in CHO cells, and investigate redox state maintenance via cyto- plasmic thioredoxin in the yeast Saccharomyces cerevisi- ae. Similarly, Chen’s group designed a highly selective and sensitive FRET-based sensor for organic hydroperox- ides (OHPs), which constantly generate cellular stress, by combining the responsive domain of the transcriptional regulatory protein OhrR with an environmentally sensi- tive FP, a modified Venus [15]. Moreover, ROS or redox state probes could be designed based on redox regulato- ry systems from different species to monitor the response
  • 7. Biotechnology Journal Biotechnol. J. 2013, 8, 1280–1291 www.biotechnology-journal.com 1286 © 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim processes of redox systems per se, and further define the mechanisms of these regulatory networks. More conven- ient and effective probes still need to be mined for com- plex mechanism research. 4.8 Other significant fluorescent sensors Traditionally, carbon monoxide (CO) is viewed as a toxi- cant or a pollutant. However, a recent study revealed that CO, like NO, functions as an essential second messenger [95]. Wang et al. [17] constructed the fluorescent probe COSer for CO by fusing cpYFP to a dimeric CO-sensitive heme protein from Rhodospirillum rubrum, CooA, as a CO recognition site. However, further studies are needed to improve the properties of CO probes used in biological systems and other settings, such as industrial facilities. Sato et al. [96] have reported a novel cell-based indicator to visualize NO release from living cells, made by com- bining endogenously expressed guanylate cyclase with a FRET-based cGMP indicator. The indicator was used to visualize NO diffusion from single vascular endothelial cells. It showed good sensitivity, selectivity, and potential for deciphering NO dynamics in biological systems. 2-Oxogluatarate (2OG), a metabolite from the highly conserved Krebs cycle, not only plays a critical role in metabolism, but also acts as a signaling molecule in a Table 1. A list of fluorescent biosensors for studying metabolism in vivo Name Applied for Based on Sensory domain Donor/receptor Ref ATeams ATP Single-chain FRET FoF1-ATP synthase MseCFP/ cpmVenus [46] EAF Modified FoF1 synthase GFP/YFP [49] Perceval [ATP]/[ADP] cpFP GlnK1 cpmVenus [47] ICUE cAMP Single-chain FRET Epac1 ECFP/Citrine [52] HCN2-camps HCN2 channel EYFP/ECFP [54] 22F BRET PKA /RIIβB Luciferase [44] RII-CFP/C-YFP Bimolecular FRET PKA CFP/YFP [102] cGMP Single-chain FRET cNMP-BD/GAF EYFP/ECFP [64] FlincGs cpFP PKG I cpEGFP [65] CGY Single-chain FRET PKG I ECFP/EYFP [62] Cygnet-2.1 cNMP-BD ECFP/Citrine [63, 103, 104] Frex/ FrexH NADH cpFP Rex cpYFP [70] Peredox [NADH]/[NAD+] cpFP Rex cpT-Sapphire [71] GluSnFR Glu Single-chain FRET Gltl ECFP/Citrine [78] FLIPQ-TV3.0 Gln GlnH mTFP1/Venus [79] OGsor 2OG GAF ECFP/EYFP [97] Arg QBP ECFP/Citrine [105] Laconic Lactate LldR mTFP/Venus [98] FLIPsuc-4mu Sucrose Sugar-binding protein ECFP/EYFP [85] FILP-HisJ His HisJ ECFP/Venus [80] FLIPGlu Glucose/galactose GGBP CFP/YFP [82] FLIPmal Maltose MBP ECFP/EYFP [83] FLIPrib Ribose RBP ECFP/EYFP [84] FLIPara Arabinose L-Arabinose-binding protein ECFP/EYFP [86] roGFP1/roGFP2 ROS roFP Surface-exposed cysteine roGFP [89] HyPer H2O2 cpFP OxyR cpYFP [91] Hyper-2 Mutant OxyR [92] CY-RL5 Redox states Single-chain FRET RL5 ECFP/EYFP [37] Redoxfluor Yap1 c-CRD Cerulean/Citrine [93, 94] OHSer OHPs cpFP Xc-OhrR cpVenus [15] CLPY BAI-2 Single-chain FRET LuxP CFP/YFP [106, 107] FLIP-CIT Citrate CitA Venus/CFP [29] COSer CO cpFP CooA cpYFP [95] FRET-MT NO Single-chain FRET hMTIIa ECFP/EYFP [108] Piccell Cell-based indicator sGC CFP/YFP [96] RNA-based ADP RNA-based ADP binding aptamer DFHBI [99, 100] Adenosine fluorescence Ade binding aptamer Guanine enlargement Gua binding aptamer GTP GTP binding aptamer SAM SAM binding aptamer
  • 8. 1287 Biotechnol. J. 2013, 8, 1280–1291 www.biotecvisions.com © 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim selecting the targeted binding domains for signaling mol- ecules, especially in small molecules, is difficult. Recent- ly, fluorescent RNA has been generated as a novel genet- ically encoded biosensor that mimics GFP. It is increas- ingly used in various applications [99]. GFP contains a three-residue fluorophore formation that produces the green fluorescence, whereas native RNA lacks a fluo- rophore element. Therefore, to confer a GFP-like ability to RNA, Paige et al. [100] engineered a fluorophore-binding RNA aptamer, termed Spinach, containing 3,5-difluoro- 4-hydroxybenzylidene imidazolinone (DFHBI), which resembles the fluorophore in GFP, thereby yielding a dif- ferent type of genetically encoded biosensor with GFP- like brightness, photostability, and cellular compatibility. Moreover, RNA-fluorophore complexes with emission wavelengths spanning the entire visible spectrum have been created using different optimized aptamers. These RNA-based probes can be used widely to detect a variety of small molecules, such as adenosine, ADP, SAM, gua- nine, and GTP. DFHBI fluorescence is activated when the targeted small molecule binds to a specific aptamer con- nected to Spinach via an optimized stem (Fig. 4). RNA- based sensors, unlike FP-based FRET sensors, produce approximately 20-fold increases in fluorescence upon metabolite binding. Thus, they are promising tools for imaging various metabolites in vivo. 6 Conclusions The discovery and artificial evolution of FPs has been a prominent contribution to biology [101], making it possi- ble to directly observe biochemical processes inside liv- ing cells and organisms [6]. With developments in FPs and variety of organisms. Environmental inorganic nitrogen is reduced to ammonium by microorganisms, which boast the metabolic pathways via conversion of 2OG to gluta- mate and glutamine. Recently, Zhang et al. [97] developed FRET-based biosensors for tracking 2OG in real-time; the dynamic ranges of the sensors appeared identical to the physiological range observed in E. coli. Citrate is another important intermediate in catabolic pathways involving glycolysis and lipid metabolism. Ewald et al. [29] have described FRET-based citrate sensors. In an optimization process, different peptide linkers were screened to achieve a maximal change ratio, and residues in the cit- rate-binding pocket were modified to construct sensors with the proper affinity for the application. When expressed in E. coli, the indicator showed that cells could respond to a carbon source even after long-term starva- tion. Lactate also plays significant roles in healthy tissues. As an abnormal lactate level is associated with inflamma- tion, hypoxia, ischemia, neurodegeneration, and cancer, tracing lactate levels in cells has diagnostic and thera- peutic applications. San Martin et al. [98] recently devel- oped a lactate sensor that discriminates lactate flux in dif- ferent cells and found that T98G glioma cells have a three to fivefold higher rate of lactate production than normal astrocytes, which is consistent with tumor metabolism. All genetically encoded fluorescent biosensors construct- ed to trace intracellular metabolites are listed in Table 1. 5 RNA-based fluorescent sensors FP-based biosensors are powerful tools for visualizing cel- lular activities, but the fusion of FPs may potentially inter- fere with the function of sensory domains. Moreover, Figure 4. RNA sensor structure and imaging of SAM in E. coli. (A) The sensor consists of Spinach (black), a transducer (orange), and a target-binding aptamer (blue). DFHBI (green) fluorescence is activated when the sensor binds to the target metabolite (purple). (B) Emission spectra of the SAM sensor. (C) Imaging shows distinct SAM accumulation pat- terns. Some cells exhibit higher than average (arrow) or slow (arrowhead) increases in SAM, and one cell shows increased and then decreased SAM levels (double arrow).
  • 9. Biotechnology Journal Biotechnol. J. 2013, 8, 1280–1291 www.biotechnology-journal.com 1288 © 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim design methods, genetically encoded fluorescent probes have enabled scientists to decode the mechanisms of intracellular signal transduction pathways and gather a large amount of biological information from cellular sys- tems. Genetically encoded fluorescent indicators present a popular way to study metabolic processes, take advan- tage of FRET technology, and allow noninvasive, spa- tiotemporal monitoring of signaling molecules in vivo. Future work will expand the range of detection ligands to include molecules such as 2OG, ADP, NAD, and others, and will emphasize the optimization of existing biosen- sors [33]. Moreover, RNA-based fluorescent sensors will provide a new platform for exploration in cells. This study was supported by the China NSF (21276079, 21335003), SRFDP (no. 20120074110009), the Key Grant Project (no. 313019) of the Chinese Ministry of Education, and the Fundamental Research Funds for the Central Uni- versities. The authors declare no commercial or financial conflict of interest. 7 References [1] Morise, H., Shimomura, O., Johnson, F. H., Winant, J., Intermolecu- lar energy transfer in the bioluminescent system of Aequorea. Bio- chemistry 1974, 13, 2656–2662. [2] Prasher, D. C., Eckenrode, V. K., Ward, W. W., Prendergast, F. G., Cormier, M. J., Primary structure of the Aequorea victoria green-flu- orescent protein. Gene 1992, 111, 229–233. [3] Tsien, R. Y., The green fluorescent protein. Annu. Rev. Biochem. 1998, 67, 509–544. [4] Cody, C. W., Prasher, D. C., Westler, W. M., Prendergast, F. G., Ward, W. W., Chemical structure of the hexapeptide chromophore of the Aequorea green-fluorescent protein. Biochemistry 1993, 32, 1212– 1218. [5] Ormo, M., Cubitt, A. B., Kallio, K., Gross, L. A. et al., Crystal struc- ture of the Aequorea victoria green fluorescent protein. Science 1996, 273, 1392–1395. [6] Day, R. N., Davidson, M. W., The Fluorescent protein palette: Tools for cellular imaging. Chem. Soc. Rev. 2009, 38, 2887–2921 [7] Baird, G. S., Zacharias, D. A., Tsien, R. Y., Biochemistry, mutagene- sis, and oligomerization of DsRed, a red fluorescent protein from coral. Proc. Natl. Acad. Sci. USA 2000, 97, 11984–11989. [8] Wang, L., Jackson, W. C., Steinbach, P. A., Tsien, R. Y., Evolution of new nonantibody proteins via iterative somatic hypermutation. Proc. Natl. Acad. Sci. USA 2004, 101, 16745–16749. [9] Wiedenmann, J., Ivanchenko, S., Oswald, F., Schmitt, F. et al., EosFP, a fluorescent marker protein with UV-inducible green-to-red fluo- rescence conversion. Proc. Natl. Acad. Sci. USA 2004, 101, 15905– 15910. [10] Adam, V., Moeyaert, B., David, C. C., Mizuno, H. et al., Rational design of photoconvertible and biphotochromic fluorescent proteins for advanced microscopy applications. Chem. Biol. 2011, 18, 1241– 1251. [11] Baird, G. S., Zacharias, D. A., Tsien, R. Y., Circular permutation and receptor insertion within green fluorescent proteins. Proc. Natl. Acad. Sci. USA 1999, 96, 11241–11246. [12] Kawai, Y., Sato, M., Umezawa, Y., Single color fluorescent indicators of protein phosphorylation for multicolor imaging of intracellular sig- nal flow dynamics. Anal. Chem. 2004, 76, 6144–6149. [13] Mizuno, T., Murao, K., Tanabe, Y., Oda, M., Tanaka, T., Metal-ion- dependent GFP emission in vivo by combining a circularly permu- tated green fluorescent protein with an engineered metal-ion-bind- ing coiled-coil. J. Am. Chem. Soc. 2007, 129, 11378–11383. [14] Leder, L., Stark, W., Freuler, F., Marsh, M. et al., The structure of Ca2+ sensor Case16 reveals the mechanism of reaction to low Ca2+ con- centrations. Sensors (Basel) 2010, 10, 8143–8160. Bang-Ce Ye is Professor of biology and chemistry in the East China University of Science and Technology in China, and Director of the Laboratory of Biosystems and Microanalysis in State Key Laboratory of Bioreactor Engineer- ing. He obtained his PhD degree in 1998. His research now focuses on analytical biotechnology and systems biotechnology. Chang Zhang received his BSc in Life Science in 2009 from the East China University of Science and Technology in China. He is pursuing a PhD in bio- chemistry and molecular biology in the Laboratory of Biosystems and Micro- analysis group at the same institution. His research is centered on tracking intracellular metabolites in real time by developing genetically encoded biosensors. His present work focuses on the study of metabolites such as 2-oxogluatarate in tumor cells to provide a better understanding of the regulatory pathways and signal- ing networks for cancer. Zihan Wei majored in biotechnology at Jiangsu University (B.S. 2007) before pursuing graduate work in the Labora- tory of Biosystems and Microanalysis, State Key Laboratory of Bioreactor Engi- neering at the East China University of Science and Technology. Under the direction of Prof. Bang-Ce Ye, Zihan is currently engineering genetically encoded FRET-based biosensors for detecting metabolites such as ROS and 2-oxoglutarate in Saccharopolyspora erythraea and characteriz- ing its related metabolic pathways.
  • 10. 1289 Biotechnol. J. 2013, 8, 1280–1291 www.biotecvisions.com © 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim [15] Zhao, B. S., Liang, Y., Song, Y., Zheng, C. et al., A highly selective fluorescent probe for visualization of organic hydroperoxides in liv- ing cells. J. Am. Chem. Soc. 2010, 132, 17065–17067. [16] Schwarzlander, M., Logan, D. C., Fricker, M. D., Sweetlove, L. J., The circularly permuted yellow fluorescent protein cpYFP that has been used as a superoxide probe is highly responsive to pH but not super- oxide in mitochondria: Implications for the existence of superoxide ‘flashes’. Biochem. J. 2011, 437, 381–387. [17] Wang, J., Karpus, J., Zhao, B. S., Luo, Z. et al., A selective fluorescent probe for carbon monoxide imaging in living cells. Angew. Chem. Int. Ed. Engl. 2012, 51, 9652–9656. [18] Liu, S., He, J., Jin, H., Yang, F. et al., Enhanced dynamic range in a genetically encoded Ca2+ sensor. Biochem. Biophys. Res. Commun. 2011, 412, 155–159. [19] Carlson, H. J., Cotton, D. W., Campbell, R. E., Circularly permuted monomeric red fluorescent proteins with new termini in the beta- sheet. Protein Sci. 2010, 19, 1490–1499. [20] Shui, B., Wang, Q., Lee, F., Byrnes, L. J. et al., Circular permutation of red fluorescent proteins. PLoS One 2011, 6, e20505. [21] Magliery, T. J., Wilson, C. G., Pan, W., Mishler, D. et al., Detecting protein-protein interactions with a green fluorescent protein frag- ment reassembly trap: Scope and mechanism. J. Am. Chem. Soc. 2005, 127, 146–157. [22] Ohad, N., Yalovsky, S., Utilizing bimolecular fluorescence comple- mentation (BiFC) to assay protein-protein interaction in plants. Methods Mol. Biol. 2010, 655, 347–358. [23] Tian, G., Lu, Q., Zhang, L., Kohalmi, S. E., Cui, Y., Detection of pro- tein interactions in plant using a gateway compatible bimolecular fluorescence complementation (BiFC) system. J. Vis. Exp. 2011, (55). pii: 3473. [24] Shyu, Y. J., Suarez, C. D., Hu, C. D., Visualization of ternary com- plexes in living cells by using a BiFC-based FRET assay. Nat. Protoc. 2008, 3, 1693–1702. [25] Pfeiffer, D., Jendrossek, D., Development of a transferable bimolecu- lar fluorescence complementation (BiFC) system for the investiga- tion of interactions between poly(3-hydroxybutyrate) (PHB) granule- associated proteins in Gram-negative bacteria. Appl. Environ. Microbiol. 2013, 79, 2989–2999. [26] Kodama, Y., Hu, C. D., Bimolecular fluorescence complementation (BiFC) analysis of protein-protein interaction: How to calculate sig- nal-to-noise ratio. Methods Cell Biol. 2013, 113, 107–121. [27] Miranda, J. G., Weaver, A. L., Qin, Y., Park, J. G. et al., New alter- nately colored FRET sensors for simultaneous monitoring of Zn(2)(+) in multiple cellular locations. PLoS One 2012, 7, e49371. [28] Salonikidis, P. S., Niebert, M., Ullrich, T., Bao, G. et al., An ion-insen- sitive cAMP biosensor for long term quantitative ratiometric fluo- rescence resonance energy transfer (FRET) measurements under variable physiological conditions. J. Biol. Chem. 2011, 286, 23419– 23431. [29] Ewald, J. C., Reich, S., Baumann, S., Frommer, W. B., Zamboni, N., Engineering genetically encoded nanosensors for real-time in vivo measurements of citrate concentrations. PLoS One 2011, 6, e28245. [30] de Lorimier, R. M., Smith, J. J., Dwyer, M. A., Looger, L. L. et al., Con- struction of a fluorescent biosensor family. Protein Sci. 2002, 11, 2655–2675. [31] Piljic, A., de Diego, I., Wilmanns, M., Schultz, C., Rapid development of genetically encoded FRET reporters. ACS Chem. Biol. 2011, 6, 685–691. [32] Deuschle, K., Okumoto, S., Fehr, M., Looger, L. L. et al., Construction and optimization of a family of genetically encoded metabolite sen- sors by semirational protein engineering. Protein Sci. 2005, 14, 2304–2314. [33] Nagai, T., Yamada, S., Tominaga, T., Ichikawa, M., Miyawaki, A., Expanded dynamic range of fluorescent indicators for Ca(2+) by cir- cularly permuted yellow fluorescent proteins. Proc. Natl. Acad. Sci. USA 2004, 101, 10554–10559. [34] Zaccolo, M., De Giorgi, F., Cho, C. Y., Feng, L. et al., A genetically encoded, fluorescent indicator for cyclic AMP in living cells. Nat. Cell Biol. 2000, 2, 25–29. [35] van Dongen, E. M., Dekkers, L. M., Spijker, K., Meijer, E. W. et al., Ratiometric fluorescent sensor proteins with subnanomolar affinity for Zn(II) based on copper chaperone domains. J. Am. Chem. Soc. 2006, 128, 10754–10762. [36] Mori, M. X., Imai, Y., Itsuki, K., Inoue, R., Quantitative measurement of Ca(2+)-dependent calmodulin-target binding by Fura-2 and CFP and YFP FRET imaging in living cells. Biochemistry 2011, 50, 4685–4696. [37] Kolossov, V. L., Spring, B. Q., Sokolowski, A., Conour, J. E. et al., Engi- neering redox-sensitive linkers for genetically encoded FRET-based biosensors. Exp. Biol. Med. (Maywood) 2008, 233, 238–248. [38] John, S. A., Ottolia, M., Weiss, J. N., Ribalet, B., Dynamic modulation of intracellular glucose imaged in single cells using a FRET-based glucose nanosensor. Pflugers Arch. 2008, 456, 307–322. [39] Achour, L., Kamal, M., Jockers, R., Marullo, S., Using quantitative BRET to assess G protein-coupled receptor homo- and heterodimer- ization. Methods Mol. Biol. 2011, 756, 183–200. [40] Kocan, M., Pfleger, K. D., Study of GPCR-protein interactions by BRET. Methods Mol. Biol. 2011, 746, 357–371. [41] Xie, Q., Soutto, M., Xu, X., Zhang, Y., Johnson, C. H., Biolumines- cence resonance energy transfer (BRET) imaging in plant seedlings and mammalian cells. Methods Mol. Biol. 2011, 680, 3–28. [42] Newman, R. H., Fosbrink, M. D., Zhang, J., Genetically encodable fluorescent biosensors for tracking signaling dynamics in living cells. Chem. Rev. 2011, 111, 3614–3666. [43] Dragulescu-Andrasi, A., Chan, C. T., De, A., Massoud, T. F., Gamb- hir, S. S., Bioluminescence resonance energy transfer (BRET) imag- ing of protein-protein interactions within deep tissues of living sub- jects. Proc. Natl. Acad. Sci. USA 2011, 108, 12060–12065. [44] Binkowski, B. F., Butler, B. L., Stecha, P. F., Eggers, C. T. et al., A lumi- nescent biosensor with increased dynamic range for intracellular cAMP. ACS Chem. Biol. 2011, 6, 1193–1197. [45] Pellegatti, P., Raffaghello, L., Bianchi, G., Piccardi, F. et al., Increased level of extracellular ATP at tumor sites: In vivo imaging with plas- ma membrane luciferase. PLoS One 2008, 3, e2599. [46] Imamura, H., Nhat, K. P., Togawa, H., Saito, K. et al., Visualization of ATP levels inside single living cells with fluorescence resonance energy transfer-based genetically encoded indicators. Proc. Natl. Acad. Sci. USA 2009, 106, 15651–15656. [47] Berg, J., Hung, Y. P., Yellen, G., A genetically encoded fluorescent reporter of ATP:ADP ratio. Nat. Methods 2009, 6, 161–166. [48] Bell, C. J., Manfredi, G., Griffiths, E. J., Rutter, G. A., Luciferase expression for ATP imaging: Application to cardiac myocytes. Meth- ods Cell Biol. 2007, 80, 341–352. [49] Zadran, S., Sanchez, D., Zadran, H., Amighi, A. et al., Enhanced- acceptor fluorescence-based single cell ATP biosensor monitors ATP in heterogeneous cancer populations in real time. Biotechnol. Lett. 2013, 35, 175–180. [50] Parks, S. K., Mazure, N. M., Counillon, L., Pouyssegur, J., Hypoxia promotes tumor cell survival in acidic conditions by preserving ATP levels. J. Cell Physiol. 2013. [51] Ledur, P. F., Villodre, E. S., Paulus, R., Cruz, L. A. et al., Extracellular ATP reduces tumor sphere growth and cancer stem cell population in glioblastoma cells. Purinergic Signal. 2012, 8, 39–48. [52] DiPilato, L. M., Cheng, X., Zhang, J., Fluorescent indicators of cAMP and Epac activation reveal differential dynamics of cAMP signaling within discrete subcellular compartments. Proc. Natl. Acad. Sci. USA 2004, 101, 16513–16518.
  • 11. Biotechnology Journal Biotechnol. J. 2013, 8, 1280–1291 www.biotechnology-journal.com 1290 © 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim [53] DiPilato, L. M., Zhang, J., The role of membrane microdomains in shaping beta2-adrenergic receptor-mediated cAMP dynamics. Mol. Biosyst. 2009, 5, 832–837. [54] Nikolaev, V. O., Bunemann, M., Schmitteckert, E., Lohse, M. J., Engelhardt, S., Cyclic AMP imaging in adult cardiac myocytes reveals far-reaching beta1-adrenergic but locally confined beta2- adrenergic receptor-mediated signaling. Circ. Res. 2006, 99, 1084– 1091. [55] Fan, F., Binkowski, B. F., Butler, B. L., Stecha, P. F. et al., Novel genet- ically encoded biosensors using firefly luciferase. ACS Chem. Biol. 2008, 3, 346–351. [56] Mazina, O., Reinart-Okugbeni, R., Kopanchuk, S., Rinken, A., Bac- Mam system for FRET-based cAMP sensor expression in studies of melanocortin MC1 receptor activation. J. Biomol. Screen. 2012, 17, 1096–1101. [57] Rybalkin, S. D., Yan, C., Bornfeldt, K. E., Beavo, J. A., Cyclic GMP phosphodiesterases and regulation of smooth muscle function. Circ. Res. 2003, 93, 280–291. [58] Munzel, T., Feil, R., Mulsch, A., Lohmann, S. M. et al., Physiology and pathophysiology of vascular signaling controlled by guanosine 3’,5’-cyclic monophosphate-dependent protein kinase [corrected]. Circulation 2003, 108, 2172–2183. [59] Feil, R., Lohmann, S. M., de Jonge, H., Walter, U., Hofmann, F., Cyclic GMP-dependent protein kinases and the cardiovascular system: Insights from genetically modified mice. Circ. Res. 2003, 93, 907–916. [60] Conti, M., Beavo, J., Biochemistry and physiology of cyclic nucleotide phosphodiesterases: Essential components in cyclic nucleotide signaling. Annu. Rev. Biochem. 2007, 76, 481–511. [61] Hofmann, F., Biel, M., Kaupp, U. B., International Union of Pharma- cology. LI. Nomenclature and structure-function relationships of cyclic nucleotide-regulated channels. Pharmacol. Rev. 2005, 57, 455–462. [62] Sato, M., Hida, N., Ozawa, T., Umezawa, Y., Fluorescent indicators for cyclic GMP based on cyclic GMP-dependent protein kinase Ial- pha and green fluorescent proteins. Anal. Chem. 2000, 72, 5918– 5924. [63] Honda, A., Adams, S. R., Sawyer, C. L., Lev-Ram, V. et al., Spa- tiotemporal dynamics of guanosine 3’,5’-cyclic monophosphate revealed by a genetically encoded, fluorescent indicator. Proc. Natl. Acad. Sci. USA 2001, 98, 2437–2442. [64] Russwurm, M., Mullershausen, F., Friebe, A., Jager, R. et al., Design of fluorescence resonance energy transfer (FRET)-based cGMP indi- cators: A systematic approach. Biochem. J. 2007, 407, 69–77. [65] Nausch, L. W., Ledoux, J., Bonev, A. D., Nelson, M. T., Dostmann, W. R., Differential patterning of cGMP in vascular smooth muscle cells revealed by single GFP-linked biosensors. Proc. Natl. Acad. Sci. USA 2008, 105, 365–370. [66] Kannurpatti, S. S., Joshi, N. B., Energy metabolism and NAD-NADH redox state in brain slices in response to glutamate exposure and ischemia. Metab. Brain Dis. 1999, 14, 33–43. [67] Li, Y., Dash, R. K., Kim, J., Saidel, G. M., Cabrera, M. E., Role of NADH/NAD+ transport activity and glycogen store on skeletal mus- cle energy metabolism during exercise: In silico studies. Am. J. Physiol. Cell Physiol. 2009, 296, C25–46. [68] Zhang, Q., Wang, S. Y., Nottke, A. C., Rocheleau, J. V. et al., Redox sensor CtBP mediates hypoxia-induced tumor cell migration. Proc. Natl. Acad. Sci. USA 2006, 103, 9029–9033. [69] Garriga-Canut, M., Schoenike, B., Qazi, R., Bergendahl, K. et al., 2-Deoxy-D-glucose reduces epilepsy progression by NRSF-CtBP- dependent metabolic regulation of chromatin structure. Nat. Neu- rosci. 2006, 9, 1382–1387. [70] Zhao, Y., Jin, J., Hu, Q., Zhou, H. M. et al., Genetically encoded fluo- rescent sensors for intracellular NADH detection. Cell Metab. 2011, 14, 555–566. [71] Hung, Y. P., Albeck, J. G., Tantama, M., Yellen, G., Imaging cytosolic NADH-NAD(+) redox state with a genetically encoded fluorescent biosensor. Cell Metab. 2011, 14, 545–554. [72] Forde, B. G., Lea, P. J., Glutamate in plants: Metabolism, regulation, and signalling. J. Exp. Bot. 2007, 58, 2339–2358. [73] Paczek, V., Dubois, F., Sangwan, R., Morot-Gaudry, J. F. et al., Cellu- lar and subcellular localisation of glutamine synthetase and gluta- mate dehydrogenase in grapes gives new insights on the regulation of carbon and nitrogen metabolism. Planta 2002, 216, 245–254. [74] Plaitakis, A., Zaganas, I., Regulation of human glutamate dehydro- genases: Implications for glutamate, ammonia and energy metabo- lism in brain. J. Neurosci. Res. 2001, 66, 899–908. [75] Peng, S., Zhang, Y., Zhang, J., Wang, H., Ren, B., Glutamate recep- tors and signal transduction in learning and memory. Mol. Biol. Rep. 2011, 38, 453–460. [76] Dulla, C., Tani, H., Okumoto, S., Frommer, W. B. et al., Imaging of glu- tamate in brain slices using FRET sensors. J. Neurosci. Methods 2008, 168, 306–319. [77] Okumoto, S., Looger, L. L., Micheva, K. D., Reimer, R. J. et al., Detec- tion of glutamate release from neurons by genetically encoded sur- face-displayed FRET nanosensors. Proc. Natl. Acad. Sci. USA 2005, 102, 8740–8745. [78] Hires, S. A., Zhu, Y., Tsien, R. Y., Optical measurement of synaptic glutamate spillover and reuptake by linker optimized glutamate-sen- sitive fluorescent reporters. Proc. Natl. Acad. Sci. USA 2008, 105, 4411–4416. [79] Gruenwald, K., Holland, J. T., Stromberg, V., Ahmad, A. et al., Visu- alization of glutamine transporter activities in living cells using genetically encoded glutamine sensors. PLoS One 2012, 7, e38591. [80] Okada, S., Ota, K., Ito, T., Circular permutation of ligand-binding module improves dynamic range of genetically encoded FRET- based nanosensor. Protein Sci. 2009, 18, 2518–2527. [81] Ha, J. S., Song, J. J., Lee, Y. M., Kim, S. J. et al., Design and applica- tion of highly responsive fluorescence resonance energy transfer biosensors for detection of sugar in living Saccharomyces cerevisi- ae cells. Appl. Environ. Microbiol. 2007, 73, 7408–7414. [82] Fehr, M., Lalonde, S., Lager, I., Wolff, M. W., Frommer, W. B., In vivo imaging of the dynamics of glucose uptake in the cytosol of COS-7 cells by fluorescent nanosensors. J. Biol. Chem. 2003, 278, 19127– 19133. [83] Fehr, M., Frommer, W. B., Lalonde, S., Visualization of maltose uptake in living yeast cells by fluorescent nanosensors. Proc. Natl. Acad. Sci. USA 2002, 99, 9846–9851. [84] Lager, I., Fehr, M., Frommer, W. B., Lalonde, S., Development of a flu- orescent nanosensor for ribose. FEBS Lett. 2003, 553, 85–89. [85] Lager, I., Looger, L. L., Hilpert, M., Lalonde, S., Frommer, W. B., Con- version of a putative Agrobacterium sugar-binding protein into a FRET sensor with high selectivity for sucrose. J. Biol. Chem. 2006, 281, 30875–30883. [86] Kaper, T., Lager, I., Looger, L. L., Chermak, D., Frommer, W. B., Fluo- rescence resonance energy transfer sensors for quantitative moni- toring of pentose and disaccharide accumulation in bacteria. Biotechnol. Biofuels 2008, 1, 11. [87] Ostergaard, H., Henriksen, A., Hansen, F. G., Winther, J. R., Shed- ding light on disulfide bond formation: Engineering a redox switch in green fluorescent protein. EMBO J. 2001, 20, 5853–5862. [88] Dooley, C. T., Dore, T. M., Hanson, G. T., Jackson, W. C. et al., Imag- ing dynamic redox changes in mammalian cells with green fluores- cent protein indicators. J. Biol. Chem. 2004, 279, 22284–22293. [89] Hanson, G. T., Aggeler, R., Oglesbee, D., Cannon, M. et al., Investi- gating mitochondrial redox potential with redox-sensitive green flu- orescent protein indicators. J. Biol. Chem. 2004, 279, 13044–13053. [90] Maulucci, G., Labate, V., Mele, M., Panieri, E. et al., High-resolution imaging of redox signaling in live cells through an oxidation-sensi- tive yellow fluorescent protein. Sci. Signal. 2008, 1, pl3.
  • 12. 1291 Biotechnol. J. 2013, 8, 1280–1291 www.biotecvisions.com © 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim [91] Belousov, V. V., Fradkov, A. F., Lukyanov, K. A., Staroverov, D. B. et al., Genetically encoded fluorescent indicator for intracellular hydrogen peroxide. Nat. Methods 2006, 3, 281–286. [92] Markvicheva, K. N., Bilan, D. S., Mishina, N. M., Gorokhovatsky, A. Y. et al., A genetically encoded sensor for H2O2 with expanded dynamic range. Bioorg. Med. Chem. 2011, 19, 1079–1084. [93] Yano, T., Oku, M., Akeyama, N., Itoyama, A. et al., A novel fluores- cent sensor protein for visualization of redox states in the cyto- plasm and in peroxisomes. Mol. Cell. Biol. 2010, 30, 3758–3766. [94] Oku, M., Hoseki, J., Ichiki, Y., Sakai, Y., A fluorescence resonance energy transfer (FRET)-based redox sensor reveals physiological role of thioredoxin in the yeast Saccharomyces cerevisiae. FEBS Lett. 2013, 587, 793–798. [95] Yuan, L., Lin, W., Tan, L., Zheng, K., Huang, W., Lighting up carbon monoxide: Fluorescent probes for monitoring CO in living cells. Angew. Chem. Int. Ed. Engl. 2013, 52, 1628–1630. [96] Sato, M., Nakajima, T., Goto, M., Umezawa, Y., Cell-based indica- tor to visualize picomolar dynamics of nitric oxide release from liv- ing cells. Anal. Chem. 2006, 78, 8175–8182. [97] Zhang, C., Wei, Z., Ye, B. C., Quantitative monitoring of 2-oxoglu- tarate in Escherichia coli cells by a fluorescence resonance energy transfer-based biosensor. Appl. Microbiol. Biotechnol. 2013, 97, 8307–8316. [98] San Martin, A., Ceballo, S., Ruminot, I., Lerchundi, R. et al., A genetically encoded FRET lactate sensor and its use to detect the Warburg effect in single cancer cells. PLoS One 2013, 8, e57712. [99] Paige, J. S., Wu, K. Y., Jaffrey, S. R., RNA mimics of green fluores- cent protein. Science 2011, 333, 642–646. [100] Paige, J. S., Nguyen-Duc, T., Song, W., Jaffrey, S. R., Fluorescence imaging of cellular metabolites with RNA. Science 2012, 335, 1194. [101] Tsien, R. Y., Constructing and exploiting the fluorescent protein paintbox (Nobel Lecture). Angew. Chem. Int. Ed. Engl. 2009, 48, 5612–5626. [102] Zaccolo, M., Pozzan, T., Discrete microdomains with high concen- tration of cAMP in stimulated rat neonatal cardiac myocytes. Sci- ence 2002, 295, 1711–1715. [103] Honda, A., Sawyer, C. L., Cawley, S. M., Dostmann, W. R., Cygnets: In vivo characterization of novel cGMP indicators and in vivo imag- ing of intracellular cGMP. Methods Mol. Biol. 2005, 307, 27–43. [104] Honda, A., Moosmeier, M. A., Dostmann, W. R., Membrane-per- meable cygnets: Rapid cellular internalization of fluorescent cGMP-indicators. Front. Biosci. 2005, 10, 1290–1301. [105] Bogner, M., Ludewig, U., Visualization of arginine influx into plant cells using a specific FRET-sensor. J. Fluoresc. 2007, 17, 350–360. [106] Rajamani, S., Zhu, J., Pei, D., Sayre, R., A LuxP-FRET-based reporter for the detection and quantification of AI-2 bacterial quo- rum-sensing signal compounds. Biochemistry 2007, 46, 3990–3997. [107] Rajamani, S., Sayre, R. T., A sensitive fluorescence reporter for monitoring quorum sensing regulated protease production in Vib- rio harveyi. J. Microbiol. Methods 2011, 84, 189–193. [108] Pearce, L. L., Gandley, R. E., Han, W., Wasserloos, K. et al., Role of metallothionein in nitric oxide signaling as revealed by a green flu- orescent fusion protein. Proc. Natl. Acad. Sci. USA 2000, 97, 477– 482.
  • 13. © 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.biotechnology-journal.com Editorial: Biotechnology Journal in Asia – the first AFOB special issue Tai Hyun Park and George G.Q. Chen http://dx.doi.org/10.1002/biot.201300415 Review Organ-on-a-chip technology and microfluidic whole-body models for pharmacokinetic drug toxicity screening Jong Bum Lee and Jong Hwan Sung http://dx.doi.org/10.1002/biot.201300086 Review Microfluidic-integrated biosensors: Prospects for point-of-care diagnostics Suveen Kumar, Saurabh Kumar, Md. Azahar Ali, Pinki Anand,Ved Varun Agrawal, Renu John, Sagar Maji and Bansi D. Malhotra http://dx.doi.org/10.1002/biot.201200386 Review Imaging and tracing of intracellular metabolites utilizing genetically encoded fluorescent biosensors Chang Zhang, Zi-Han Wei and Bang-Ce Ye http://dx.doi.org/10.1002/biot.201300001 Review Applications of cell-free protein synthesis in synthetic biology: Interfacing bio-machinery with synthetic environments Kyung-Ho Lee and Dong-Myung Kim http://dx.doi.org/10.1002/biot.201200385 Review Current developments in high-throughput analysis for microalgae cellular contents Tsung-Hua Lee, Jo-Shu Chang and Hsiang-Yu Wang http://dx.doi.org/10.1002/biot.201200391 Research Article Heterologous prime-boost immunization regimens using adenovirus vector and virus-like particles induce broadly neutralizing antibodies against H5N1 avian influenza viruses Shih-Chang Lin, Wen-Chun Liu,Yu-Fen Lin, Yu-Hsuan Huang, Jin-Hwang Liu and Suh-Chin Wu http://dx.doi.org/10.1002/biot.201300116 Research Article Repetitive Arg-Gly-Asp peptide as a cell-stimulating agent on electrospun poly(ε-caprolactone) scaffold for tissue engineering Pacharaporn Chaisri, Artit Chingsungnoen and Sineenat Siri http://dx.doi.org/10.1002/biot.201300191 Research Article Global gene expression analysis of Saccharomyces cerevisiae grown under redox potential-controlled very-high-gravity conditions Chen-Guang Liu,Yen-Han Lin and Feng-Wu Bai http://dx.doi.org/10.1002/biot.201300127 Research Article Automated formation of multicomponent-encapuslating vesosomes using continuous flow microcentrifugation Huisoo Jang, Peichi C. Hu, Sungho Jung, Won Young Kim, Sun Min Kim, Noah Malmstadt and Tae-Joon Jeon http://dx.doi.org/10.1002/biot.201200388 Research Article Size and CT density of iodine-containing ethosomal vesicles obtained by membrane extrusion: Potential for use as CT contrast agents Bomin Na, Byoung Wook Choi and Bumsang Kim http://dx.doi.org/10.1002/biot.201300110 Research Article Site-targeted non-viral gene delivery by direct DNA injection into the pancreatic parenchyma and subsequent in vivo electroporation in mice Masahiro Sato, Emi Inada, Issei Saitoh, Masato Ohtsuka, Shingo Nakamura,Takayuki Sakurai and Satoshi Watanabe http://dx.doi.org/10.1002/biot.201300169 Biotechnology Journal – list of articles published in the November 2013 issue. This issue is the first official special issue of Biotechnology Journal with the Asian Federation of Biotechnology (AFOB). In the cover, we use the grain as a symbol and analogy for biotechnology, as both the grain and biotechnology are common and uniting factors that link the member countries of the AFOB. Gran image credit: © Myimagine – Fotolia.com. Systems & Synthetic Biology · Nanobiotech · Medicine ISSN 1860-6768 · BJIOAM 8 (11) 1245–1364 (2013) · Vol. 8 · November 2013 11/2013 Microfluidics Biosensors Organ-on-chip www.biotechnology-journal.com