Accepted Manuscript Title: Development of fluorescent and luminescent probes for reactive oxygen species Author: Huai-Song Wang PII: DOI: Reference:
S0165-9936(16)30208-4 http://dx.doi.org/doi: 10.1016/j.trac.2016.09.006 TRAC 14825
To appear in:
Trends in Analytical Chemistry
Please cite this article as: Huai-Song Wang, Development of fluorescent and luminescent probes for reactive oxygen species, Trends in Analytical Chemistry (2016), http://dx.doi.org/doi: 10.1016/j.trac.2016.09.006. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Development of fluorescent and luminescent probes for reactive oxygen species
1 2 3
Huai-Song Wanga,b
4 5 6 7 8 9
a
Department of Pharmaceutical Analysis, China Pharmaceutical University, Nanjing, 210009, China.
b
Key Laboratory of Drug Quality Control and Pharmacovigilance (China Pharmaceutical University), Ministry of
Education, Nanjing 210009, China.
Highlights
10 11 12 13 14
Reactive oxygen species (ROS) can mediate a wide variety of biological processes. Due to their reactive and transient nature, ROS are generally difficult to determination. Fluorescent and luminescent probes for monitoring ROS in biological systems. ROS probes (including small organic molecules, metal complexes and nanomaterials) were discussed.
15 16 17
The design strategies and ROS sensing mechanisms of these functional probes were described.
18
ABSTRACT
19
Reactive oxygen species (ROS) are chemically reactive molecules that can mediate a wide
20
variety of biological processes. The imbalance of these reactive intermediates in the metabolism
21
will result in the phenomenon known as oxidative stress. Therefore, a great number of approaches
22
have been developed for measuring ROS in biological systems. Due to their reactive and transient
23
nature, the ROS are generally difficult to determination. Fluorescent and luminescent probes for
24
monitoring ROS have shown advantages such as high sensitivity, selectivity, as well as real-time
25
imaging, which can yield visible information about the ROS. The recent progress in preparing
26
ROS probes (including small organic molecules, metal complexes or nanomaterials) for detecting
27
and imaging of ROS production in living cells or whole organisms were summarized in this
28
review. The design strategies and ROS sensing mechanisms of these functional probes were
29
described.
30 31
Keywords: Reactive oxygen species; Fluorescent probes; Luminescent probes; Multi-functional
32
probes; Ratiometric probes; Biological systems
33 34
1. Introduction
35
Reactive oxygen species (ROS), as a class of highly reactive chemicals, play important roles in
36
varieties of physiological and pathological processes.[1, 2] The balance of oxidation-antioxidation *
Corresponding author. E-mail:
[email protected] (H.-S. Wang) 1
Page 1 of 43
37
modulated by ROS in biosystems is crucial for maintaining normal cell functions. That is,
38
ROS-induced disease can be either resulted from the lack of ROS (e.g., chronic granulomatous
39
disease, certain autoimmune disorders) or excessive production of ROS (e.g., cancer, arthritis,
40
arteriosclerosis).[3, 4] ROS can be neutral molecules [such as hydrogen peroxide (H2O2), singlet
41
oxygen (1O2)], ions [such as superoxide (O2•-), hypochlorite (ClO−) and the nitrogen-containing
42
peroxynitrite (ONOO−)] or radicals [such as hydroxyl radical (•OH)].[5, 6] Due to their reactive
43
and transient nature, the ROS are usually difficult to determination especially in biological
44
systems.
45
Most intracellular ROS are derived from the reduction of molecular oxygen in the process of
46
metabolism. The major ROS (O2•- and H2O2) are from the NADPH oxidases (NOXs), xanthine
47
oxidase (XO), and the mitochondrial electron-transport chain.[7-9] Other ROS can be derived
48
from a cascade of transitions from one species to another (Scheme 1), including the superoxide
49
dismutase (SOD) catalyzed formation of H2O2 from O2•-, the reaction of O2•- with •NO to form
50
ONOO-, the peroxidase-catalyzed formation of HOCl from H2O2, and the iron-catalyzed Fenton
51
reaction leading to the generation of •OH.
52
ROS have been considered as key regulatory molecules for cells, but cellular damage can result
53
from the perturbed equilibrium between the formation and transformation of ROS. Due to their
54
high reactivity, ROS readily react with virtually all of the biological molecules. Overproduction of
55
ROS can cause damage to many cellular constituents, including proteins, carbohydrates, lipids,
56
and nucleic acids. Therefore, oxidative stress caused by high-level ROS may be associated with
57
pathologies. The ROS-metabolising systems have become an important research area for better
58
understanding their biological functions.
59 60 61
In the past decades, many approaches (including electrochemical, spectroscopic, and enzymatic
62
techniques) for detecting ROS have been developed.[10-12] The electron spin resonance (ESR)
63
method has been used as a powerful for detecting ROS (such as O2•-, •OH and 1O2), which need to
64
form stable free radicals (spin adducts) by spin trapping.[13, 14] But the ESR technique requires
65
relatively expensive instruments and cannot be employed readily to acquire quantitative estimates
66
of ROS due to many secondary reactions during spin trapping.[15] Therefore, high-performance 2
Page 2 of 43
67
liquid chromatography (HPLC) has been employed to separate the byproducts after ROS
68
trapping.[16, 17] The aromatic hydroxylation, arising from the reaction of ROS with aromatic
69
compounds (e.g., salicylic acid,[18] phenylalanine[19] and 4-hydroxybenzoate[20]), is usually
70
used for HPLC method. Several other chemical tools including electrochemical method[21] and
71
mass spectrometry analysis[22] have been devoted in detection of ROS. These methods are still
72
very complex during the sample preparation, which hampers their values for routine analysis.
73
The fluorescent and luminescent probes for monitoring ROS have shown advantages such as
74
high sensitivity, selectivity, as well as real-time imaging, which can yield visible information
75
about the ROS in biological systems.[23-25] In recent years, numerous probes (organic molecules
76
or nanomaterials) have been prepared for selective monitoring one kind of the ROS (such as H2O2,
77
O2•- or •OH) or evaluating the oxidative stress generated from the total ROS. Especially, most of
78
these probes have shown great advantages for in vivo real-time sensing ROS. In this review, I
79
mainly summarize the fluorescent and luminescent probes designed for detecting and imaging of
80
ROS production in living cells or whole organisms.
81 82
2. Different ROS production and the corresponding probes
83
2.1 Superoxide (O2•-) probes
84
The O2•- with highly oxidative activity were considered to be the precursor of other ROS.[26] In
85
the early studies, luminescence- and fluorescence-based assays have been employed to measure
86
cell-derived O2•-. The initial oxidation of the phenanthridine moiety by O2•- to generate a radical
87
intermediate was mostly used for designing the organic probes for O2•-.[27] For example,
88
hydroethidine (HE) and its derivatives were frequently used as probes for sensing local O2•- in the
89
mitochondria. The reaction of the non-fluorescent HE with O2•- leads to a specific hydroxylated
90
product with highly fluorescence (Scheme 2).[28, 29] The major drawback of HE related probes is
91
the poor selectivity toward O2•-. Such probes are usually light-sensitive, and the experiment
92
procedures should be performed in dim light.[5, 30, 31]
93 94 95
Recently, two-photon (TP) fluorescent probes for selectively and sensitively monitoring O2•-
96
were synthesized by Tang and co-workers.[32, 33] Based on the scavenging activity of caffeic acid 3
Page 3 of 43
97
(or its derivatives) toward O2•-,[34] the probe 2,5-di(4ʹ-caffeic acid amidestyrene) pyrazine
98
(PY-CA, Scheme 3), conjugated with two caffeic acid molecules and a symmetric styryl-pyrazine
99
as TP absorption cross section, shows high specificity for O2•-, instantaneous response and
100
reversible interaction.[33] The O2•- sensing mechanism is based on the transform from
101
pyrocatechol to benzoquinone. After interacted with O2•-, the generated PY-CAO can emit a
102
maximal fluorescence around 520 nm, and excitation wavelength is 800 nm for TP. This probe can
103
be applied to image O2•- in living tissues and intact organisms. The O2•- images can reach 900 μm
104
into
105
acyl)-1,3,5-triazine-2,4,6- triamine (TCA, Scheme 3), also conjugated with two caffeic acid
106
molecules and show high selectivity for O2•- in live cells and in vivo.[32] Both of the two probes
107
(PY-CA and TCA) exhibit reversible on-off-on type fluorescence response mediated O2•- and
108
glutathione (GSH).
tumor
tissues.
Another
TP
probe,
N,
Nʹ-di-((2E)-3-(3,4-dihydroxyphenyl)acrylic
109 110 111
The cell-penetrating nanotechnology-based probes also offer reliable and durable approach for
112
intracellular biosensing of O2•-. Tian and co-workers prepared a carbon-dot-based probe which
113
employs carbon dots (C-Dots) as the reference fluorophore and HE as the recognition element
114
toward O2•-.[35] The prepared probe CD-HE was used as a dual-emission ratiometric probe: one
115
emission peak at 525 nm from C-Dots as inner reference and one shoulder at 610 nm generated
116
form the reaction between HE and O2•- (Fig. 1). Such inorganic-organic probe demonstrated well
117
stability against pH changes and continuous light illumination and low cytotoxicity. Furthermore,
118
in another work, the HE and fluorescein isothiocyanate (FITC) were loaded in the hollow of the
119
rattle-type silica colloidal particles. The probe with core-shell structure shows similar ratiometric
120
sensing ability toward O2•- compared with the probe CD-HE.[36]
121 122
Even considerable attention has been paid on the development of O2•- sensors, it is still a
123
challenging work to design and synthesize specific and sensitive to O2•-. It might because the
124
instantaneous lifetime and the high oxidation property of O2•-. In biological systems, once O2•- is
125
produced, it will spontaneously or enzymatically changed into H2O2.[37] Therefore, designing
126
fluorescent or luminescent probes that can transiently respond O2•- with well selectivity is still an 4
Page 4 of 43
127
important research area.
128 129
2.2 Hydrogen peroxide (H2O2) probes
130
The H2O2 is a relatively stable species comparing with other ROS. Evidence has shown that
131
H2O2 plays an important role as a second messenger in cellular signal transduction.[26, 38] But
132
the generation, degradation, and diffusion of H2O2 are still not well understood.[39] Thus,
133
attentions have been focus on the direct measurement of H2O2 in living systems.
134
2.2.1 Molecular probes for H2O2
135
Boronate-based fluorescent probes, based intramolecular charge-transfer (ICT) mechanism,
136
have been well used to monitor H2O2 in vitro and in vivo, because the reaction of H2O2 with
137
phenyl-boronate is highly selective and faster compared with other ROS.[40, 41] In recent studies,
138
the chemoselective deprotection of boronate esters to phenols was widely used for H2O2 (Scheme
139
4A). Additionally, if the phenyl-boronate is attached to the fluorophore through an ether-linkage
140
(Scheme 4B), a p-quinone-methide will be released after the reaction between phenyl-boronate
141
and H2O2.[42] When the boronate group is removed from the probe, the generated fluorophore
142
with brightly fluorescent can be activated.
143 144
In recent years, the H2O2-mediated “boronate to phenol” conversion is still very popular for
145
designing probes for H2O2 (Fig. 2).[43-46] Most of the probes have good performance in
146
intracellular H2O2 sensing.[40, 47-51] For example, the probe HP-1 consists of a coumarin unit
147
and a diboron xanthene spiro isobenzofuran group bridged by a disulfide bond.[47] It can be used
148
as “turn-on” dual responsive fluorescent probe for the exogenous H2O2 as well as endogenous
149
thiols in living HeLa cancer cells; The HP-3 is an aggregation-induced emission (AIE)
150
fluorescence probe for monitoring H2O2:[49] In the presence of H2O2, the phenylboronic ester
151
moiety can be convert into the phenol group, meanwhile the aggregation will occur based on the
152
AIE because the hydrophilic and hydrophobic properties of the HP-3 is changed.
153 154
Several ratiometric fluorescent probes containing the boronate moiety were also designed for
155
sensing H2O2.[52-56] The ratiometric fluorescent probes can eliminate most ambiguities by two 5
Page 5 of 43
156
emission bands that enable an accurate internal calibration. Ihmels and co-workers prepared a
157
series of boronobenzo[b]quinolizinium derivatives allowing ratiometric analysis.[53] Among them,
158
the HP-11 exhibited pronounced light-up effect and higher selectivity towards H2O2 in living cells.
159
According to Fig. 3A, the H2O2 sensing progress can be observed by a continuous decrease of the
160
intensity of the emission maximum around 420 nm which is accompanied by the simultaneous
161
increase of a new redshifted emission band around 540 nm. Hong and co-workers also synthesized
162
a series of styryl dyes in which the formylphenyl boronate esters were used as combinatorial
163
blocks.[54] Especially, the probe HP-12 showed a redshift of about 100 nm at the maximum
164
wavelength upon addition of H2O2 (Fig. 3B). It was successfully utilized for the real-time
165
monitoring of glucose oxidation in the presence of glucose oxidase (GOx).
166 167
Yi and co-workers developed a versatile ratiometric H2O2 probe (HP-13, Fig. 4A) based on
168
1,8-naphthalimide and boric acid ester.[55] The HP-13 exhibited high sensitivity toward H2O2
169
with a fluorescence ratio change of up to 1020-fold. And the HP-13 contains an azide group that
170
makes HP-13 versatile and can be potentially linked to biological molecules via the click reaction.
171
Therefore, it was used to imaging endogenous H2O2. HeLa cells incubated with HP-13 showed
172
strong blue fluorescence at first, then the intensity of the blue fluorescence decreased after
173
interacting with H2O2. Meanwhile, the yellow fluorescence intensity was increased (Fig. 4A).
174
Interestingly, when the HP-13 was modified with Nuclear Localization Signal (NLS) peptide (the
175
transmembrane molecular cargo carrier), the generated probe HP-14 can be delivered into nuclei
176
and ratiometric detection of nuclear H2O2 in living cells (Fig. 4B).
177 178
If the boronate-based fluorescent probe contains an ether- or ester-linkage between the boronate
179
moiety and fluorophore moiety, the reaction between the probe and H2O2 usually trigger the
180
remove of boronate followed by the release of p-quinone-methide.[42, 57, 58] Such H2O2 probes
181
are shown in Scheme 5 and Scheme 6.[59-61] The HP-15 and HP-16 were used as ratiometric
182
fluorescent sensor for H2O2. HP-15 was synthesized based on water-soluble hemi-cyanine dye,
183
which can be easily modified and shows superior photostability.[60] It displayed a colour change
184
from pale orange to pink in the presence of H2O2 with fast response. HP-16 can quantitatively
185
detect H2O2 by ratiometric fluorescence method with a 100 nm red-shifted emission.[61] Thus, 6
Page 6 of 43
186
HP-16 can serve as a “naked-eye” probe for H2O2.
187 188
Yin and co-workers designed a chemical probe HP-18 with good stability for detecting
189
H2O2.[62] After interacting with H2O2, the HP-18 exhibited color change from colorless to pink,
190
making it a “naked-eye” probe. The possible mechanism of the reaction of HP-18 with H2O2 was
191
proposed in Scheme 6. This probe is stable under cell culture conditions and can also be used to
192
detect enzymatically generated H2O2. The HP-19 is a turn-on near-infrared (NIR) fluorescent
193
probe for H2O2 based on dicyanomethylene-4H-chromene (DCM) fluorophore.[63] Upon the
194
addition of H2O2, a significant increase in fluorescence intensity at 700 nm can be observed, due
195
to the generation of DCM-OH. The HP-20, as a fluorescence ratiometric sensor molecule, is
196
suitable for trace vapor detection of H2O2.[64] In the presence of H2O2, the fluorescence of HP-20
197
with an emission maximum at 500 nm is converted to an electronic “push-pull” structure (DAT-N),
198
which has an emission peak at 574 nm. The HP-20 exhibits effective vapor sampling of H2O2 with
199
high detection sensitivity (down to 7.7 ppb) and fast sensor response (down to 0.5 s under 1 ppm
200
of H2O2).
201 202
Furthermore, the N-alkylated BODIPY, where p-pinacolborylbenzyl unit was attached through
203
C-N linkage to the fluorophore moiety, has been applied by Shao and co-workers for monitoring
204
and imaging of H2O2 in both living cells and living angelfish.[65, 66] For example, the reaction of
205
HP-21 with H2O2 under physiological conditions can cause the oxidation of the boronate,
206
followed by the 1,6-rearrangement elimination reaction and “Turn-On” fluorescence response with
207
suitable sensitivity (Fig. 5A and B).[66] The probe was used as a mitochondrial-targeting probe
208
for imaging H2O2 in living cells. As shown in Fig. 5C, the HP-21 signal overlaid very well with
209
the fluorescence of Mito Tracker Deep Red (MT DeepRed, commercially available mitochondrial
210
dye). It indicates that the HP-21 was located in the mitochondria and can detect localized H2O2.
211 212
2.2.2 Nanoprobes for H2O2
213
Fluorescent or luminescent nanoparticles have been widely used as fluorescent probes owing to
214
their high photostability and facile surface functionalization for specific targeting. A number of 7
Page 7 of 43
215
nanoprobes were recently developed for sensing H2O2 in living cell systems.[25, 67-70] The
216
chemoselectivity of H2O2-induced deboronation also has been applied to the preparation of
217
nanoprobes.[71-74] Niu and co-workers prepared a fluorescence resonance energy transfer
218
(FRET)-based ratiometric nanoprobe to detect intracellular H2O2.[72] The nanoprobe PMT-F127
219
micelle is a self-assembled polymeric micelle, in which the tetraphenylethene (TPE) was selected
220
as the energy donor for FRET and the fluorescent boronate was selected as H2O2-responsive
221
acceptor (Fig. 6A). In the presence of H2O2, the boronate groups were transformed into phenols,
222
resulting in a green-fluorescent fluorescein product (Fig. 6B). In living HeLa cells, the PMT-F127
223
micelles show a blue color at first. When treated with H2O2, the blue color decreased, but green
224
colors increased (Fig. 6C). Such probe was successfully used for FRET-based ratiometric
225
detection of mitochondrial H2O2. In the self-assembly process, several multi-functional molecules
226
such as targeting and energy acceptor moieties can also be simultaneously integrated into the
227
multi-functional nanoprobe.
228 229
Wu and co-workers reported another multifunctional nanoprobe, similar as PMT-F127 micelle,
230
for detecting and imaging mitochondrial H2O2.[74] The nanoprobe possesses both
231
mitochondria-targeting and FRET-based ratiometric sensing capability. The carbon-dot serves as
232
the donor of energy transfer and a boronate-based H2O2 recognition element (PFl) was covalently
233
linked onto CD. In the presence of H2O2, the PFl moieties undergo structural and spectral
234
conversion, affording the nanoplatform a FRET-based ratiometric signal. In the living cell, the
235
nanoprobe can specifically target and stain the mitochondria, as well as track the exogenous H2O2
236
levels.
237
The carbon nanodots (C-dots), as promising fluorescent materials, have elicited much research
238
interest recently due to their outstanding properties, such as good water solubility, biocompatibility,
239
and tunable fluorescence. These merits make the C-dots especially useful for fluorescent
240
bioimaging or biosensing.[75-78] Zhang and co-workers recently designed a C-dots -based
241
fluorescence turn-on sensor employing photo-induced electron transfer (PET) mechanism for
242
H2O2
243
2-(diphenylphosphino)ethylamine, the blue fluorescence of the C-dots was quenched through
244
PET(Fig. 7A and B). In the presence of H2O2, the diphenylphosphine can selectively react with
monitoring
in
aqueous
solutions.[78]
After
being
modified
with
8
Page 8 of 43
245
H2O2 and the PET derived from the electron pair of the phosphorus atom to the C-dots was
246
cancelled. Then, the fluorescence intensity of C-dots was increased (Fig. 7C). Chang and
247
co-workers developed a one-pot synthetic strategy for decorating the C-dots on graphene oxide
248
(GO).[77] The prepared C-dots@RGO nanomaterial exhibited well photoluminescence (PL) at a
249
wavelength of 440 nm upon excitation at 365 nm. The H2O2 can quench the photoluminescence of
250
the C-dots@RGO through an etching process. Therefore, such probe was used for detection of
251
H2O2 generated in an acetylcholinesterase (AChE)/choline oxidase (ChOx) system.
252 253
Cerium oxide (CeO2) nanowire was recently found to be a superquencher with long-range
254
energy transfer properties.[79-82] Tang and co-workers prepared a H2O2 nanosensor via binding
255
FAM-tagged single-strand (ss) DNA on CeO2 nanowire (Fig. 8A and B).[79] The Ce4+ on the
256
surface of CeO2 exhibits well DNA binding affinity by means of metal coordination. After
257
assembling the CeO2 nanowires with FAM-labeled ssDNA, the fluorescence of FAM groups was
258
quenched. Upon addition of H2O2, the FAM-labeled ssDNA was released form CeO2 nanowires
259
followed by fluorescence signal increasing. The designed CeO2-DNA nanosensor is capable of
260
rapidly (<1 min) and selectively tracking H2O2 in living cells and zebrafish larvae (Fig. 8C and D).
261
Qu and co-workers synthesized a H2O2 nanosenser by doping the Eu3+ into CeO2 nanorods.[81]
262
The fluorescence emission peaks can be assigned to the 5D0-7F J (J = 0-4) transitions of Eu3+ ions.
263
After treating with L(+)-ascorbic acid, the fluorescence intensity of the Eu3+-doped CeO2 nanorods
264
was decreased, resulting from the chemical reduction of Ce4+ to Ce3+ by ascorbic acid and
265
subsequent excitation cut-off of charge transfer from O2- to Ce4+. Interestingly, the fluorescence
266
could be completely recovered to their original intensity when the reduced samples were treated
267
by H2O2. Such Eu3+-doped CeO2 nanorods can be used as fluorescence switcher by alternatively
268
adding ascorbic acid and H2O2.
269 270
Graphene quantum dot (GQD) has shown great promise in the field of biosensing due to its
271
photoluminescence property contributed by quantum confinement and edge effects. Yang and
272
co-workers prepared a kind of AgNP/GQDs hybrid nanocomposite for high performance H2O2
273
detection (Fig. 9A).[83] In such nanocomposite, Ag NPs acted as quencher and recognition unit,
274
and GQDs served as a signal output unit with excellent optical property. The GQDs were 9
Page 9 of 43
275
assembled on the surface of ssDNA-modified AgNPs through π-π stacking (Fig. 9B). Then, the
276
fluorescence of GQDs was quenched by Ag NPs through the resonance energy transfer. Upon
277
H2O2 addition, obvious fluorescence recovery can be observed due to the Ag NPs etching and
278
DNA cleavage. The reaction between H2O2 and Ag NPs can generate •OH that will cleave the
279
DNA-bridge and result in the disassembly of AgNP/GQDs with further signal enhancement (Fig.
280
9C).
281 282
The functionalized fluorescent metal nanoclusters with well quantum yield have been used to
283
determine H2O2.[84, 85] For example, Lu and co-workers used the gold nanoclusters as the model
284
aggregation-induced
285
bis(2,4,6-trichlorophenyl) oxalate (TCPO)-H2O2 chemiluminescence (CL) reaction.[85] This
286
research shows that the unique AIE effect of gold nanoclusters can strongly boost the CL signal of
287
the TCPO-H2O2 system, due to the chemiluminescence resonance energy transfer (CRET) between
288
TCPO energy donors and gold nanocluster aggregate acceptors. Additionally, another CRET based
289
H2O2 probe, constructed by the assembly of CdTe quantum dots (QDs) upon the surface of layered
290
double hydroxide (LDH), has been prepared by Lu and co-workers.[86] The H2O2 recognition is
291
based on the luminol-H2O2 system, and the oriented QD-LDH nanocomposite can efficiently
292
accept the energy from luminol donors for signal amplification. This method exhibited a stable
293
response to H2O2 with a detection limit as low as 0.3 μ M.
emission
(AIE)
molecules
to
study
their
influence
on
the
294 295
2.3 Hydroxyl radical (•OH) probes
296
The •OH is generally considered as the most aggressive radicals among the ROS.[87, 88] It can be
297
generated within cells by the Fenton reaction enabled by transition metals, and plays a critical role
298
in numerous pathological processes. Different methods have been developed for the indirect or
299
direct detection of •OH, including electron paramagnetic resonance (EPR), chemiluminescence,
300
HPLC method, UV-vis spectroscopy and fluorescence methods.[89] In the past decades, the
301
fluorescent probes for •OH have been widely reported employing organic dye molecules or
302
fluorescent nanoparticles (quantum dots, metal nanoclusters, etc.). But, designing biocompatible
303
fluorescent probes for accurate and selective detection and quantification of •OH is still a
304
challenge. 10
Page 10 of 43
305
The spin labeled fluorescent probes (R2-NO•), such as hydroxyl-2,2,6,6-tetramethylpiperidine
306
-N-oxide (TEMPO), have been used for trapping and detecting the •OH. The fluorescence of the
307
fluorophore can be quenched by the nitroxide group through electron exchanges, and can be
308
recovered after interacted with a free radical.[90] For example, the probe HR-1 (Scheme 7) was
309
recently applied for trapping the •OH, then quantitatively detected by HPLC).[91] This probe has
310
the potential to trace •OH production induced by the impairment of the mitochondrial respiratory
311
chain.
312 313
Tae and co-workers prepared a fluorescent probe (HR-2, Scheme 8) for monitoring •OH based
314
on the oxidative C-H abstraction reaction of rhodamine cyclic hydrazide.[92] The •OH-induced
315
reaction of HR-2 takes place rapidly at room temperature and opens the spirocyclic ring system.
316
The probe exhibited excellent selectivity for monitoring intracellular •OH with virtually no
317
interference by other ROS species. Pierre and co-workers prepared a •OH probe (HR-3) that
318
consisted of a terbium complex with open coordination sites and a reactive pre-antenna composed
319
of an aromatic acid.[93] Without •OH, the trimesate does not coordinate, and the emission of the
320
terbium ion was not sensitized. After the reaction with •OH, the hydroxylated trimesate was
321
readily bind to terbium complex, and the terbium-centered emission was increased. Yi and
322
co-workers developed a naphthalimide-naphthyridine derivative (HR-4) for the detection of
323
•OH.[94] The reaction between HR-4 and •OH can generate a hydroxyl modified naphthyridine
324
moiety with blue fluorescence emission. The HR-4 shows excellent photostability, low
325
cytotoxicity and high biocompatibility. It shows no cellular toxicity in 36 h at a concentration of 5
326
mM.
327 328
Au (or Ag) nanoclusters have attracted significant attentions as luminescent nanomaterials due
329
to the strong quantum-confinement effect.[95, 96] Several works have reported for detecting •OH
330
based on Au or Ag nanoclusters.[97, 98] Tian and co-workers prepared a ratiometric fluorescence
331
biosensor, employing Au nanocluster (AuNC) protected by bovine serum albumin (BSA) as
332
reference fluorophore, which was conjugated with HPF (HR-5) as both the response signal and
333
specific recognition element for •OH (Fig. 10A).[98] The prepared AuNC@HPF probe showed
334
only one emission peak at 637 nm ascribed to AuNCs without •OH, because the HR-5 was almost 11
Page 11 of 43
335
non fluorescent. After interacted with •OH, the fluorescence emission at 515 nm was gradually
336
increased due to the oxidation of HR-5 (Fig. 10B), while the emission at 637 nm stays constant.
337
For ratiometric determination of •OH generated in cells, lipopolysaccharides (LPS) was
338
introduced to induce oxidative stress. The results show that the ratiometric sensor was successfully
339
used for bioimaging and monitoring of the •OH changes.
340 341
The C-dots have also been sued for sensing •OH.[99] By employing the polyvinylpyrrolidone
342
(PVP) as the only carbon source, the prepared C dots exhibit good photostability in a wide range
343
of pH solution (from 3.0 to 10.5) and excellent water solubility. Interestingly, the fluorescence
344
intensity of the C dots can be sensitively affected by •OH.
345 346
2.4 Singlet oxygen (1O2) probes
347
Recent researches have shown that the chemistry of 1O2 exhibited large biomedical significance. In
348
the medical studies, the 1O2 plays a key role in photodynamic therapy (PDT), an emerging
349
anticancer treatment using photoirradiation and photosensitizers (Sens).[100, 101] During the PDT,
350
the Sens can transfer the absorbed energy to molecular oxygen to generate 1O2, which can finally
351
destroy the cancer cells.[102] However, the 1O2 can also lead to some diseases because it is prone
352
to destruct the biological molecules (such as proteins, nucleic acids and lipids).
353
There have been several fluorescent probes for monitoring for 1O2 further outstanding
354
importance of 1O2 in photobiological processes.[103] Chemical probes, based on the reaction
355
between 1O2 and the derivatives of anthracene to form stable endoperoxides, were most frequently
356
used.[104-107] The probes with anthracene moiety can react with 1O2 via the highly favorable [4
357
+ 2] cycloaddition mechanism. A typical anthracene based 1O2 probe was shown in Scheme 9: in
358
the presence of
359
anthraquinone were broken, followed by rapid fluorescence recovery of the fluorophores.[108]
1
O2, the linkers between fluorophores (F1 and F2) and the generated
360 361 362
Majima and co-workers recently developed a far-red fluorescence probe (SO-1, Fig. 11A) for
363
monitoring 1O2 based on 9,10-diphenylanthracene (DPA).[109] The SO-1 is composed of
364
silicon-containing rhodamine and an anthracene moiety as 1O2 reactive site. It exhibited a good 12
Page 12 of 43
365
selectivity toward 1O2 out of other ROS with the emission wavelength at 640 nm (Fig. 11B). In the
366
cell experiment, 5-aminolevulinic acid (ALA) was taken up by cancer cells and then metabolized
367
to the heme precursor protoporphyrin IX (PpIX) to become a photosensitizer. The photoirradiation
368
of PpIX can generate 1O2, meanwhile increase the fluorescence intensity of SO-1 rapidly (within
369
10
370
[tetra-(N-methyl-4-pyridyl)porphyrin, TMPyP4], which changes its location from lysosome to
371
cytoplasm and nucleus upon photoirradiation (Fig. 11C). The results proved that the SO-1 can
372
respond to only mitochondrial-originated 1O2.
s).
The
SO-1
does
not
react
with
1
O2
generated
by
another
Sens
373 374
Some luminescent probes were also designed for monitoring 1O2 during the photodynamic
375
therapy.[110, 111] The SO-2 in Scheme 10 is a Eu3+ complex-based luminescence, in which the
376
terpyridine moiety is an antenna for sensitizing the Eu3+ luminescence, and the
377
10-methyl-9-anthryl moiety plays the roles of quenching the luminescence of Eu3+ and selectively
378
trapping 1O2.[110] The SO-2 and PDT drugs, indole-3-acetic acid (IAA) and hematoporphyrin
379
monomethyl ether (HMME), were co-loaded in HeLa cells. The SO-2 can react with the generated
380
1
381
lifetime. The SO-3, constructed based on a fluorescent coumarin group and a phosphorescent Ir3+
382
complex, is a ratiometric luminescent probe for monitoring 1O2.[111] The mechanism for
383
monitoring 1O2 is based on the convering the julolidine of coumarin group to an iminium form
384
(Scheme 10). The SO-3 was successfully used to monitor therapeutic 1O2 dosages by recording
385
the ratiometric photoluminescence changes.
O2 to form its endoperoxide followed by the remarkable increases in luminescence intensity and
386 387 388
2.5 Hypochlorite (ClO−) probes
389
The hypochlorite/hypochlorous acid (HOCl/ClO−) exerts a wide variety of physiological effects in
390
living systems. Endogenous HOCl/ClO− can be produced from the reaction of H2O2 and chloride
391
ions catalyzed by myeloperoxidase (MPO). The function of HOCl/ClO− in physiological processes
392
is mainly for protecting the body against microorganism invasion.[112-114] However, the
393
uncontrolled HOCl/ClO− production may lead to many inflammation-related diseases.[115]
394
A number of probes for visualizing HOCl/ClO− in intercellular systems have been repored.[116] 13
Page 13 of 43
395
Recently, a visible-light-excitable fluorescence ratiometric probe for ClO− was synthesized by
396
Goswami and co-workers.[117] The probe (H-1) exhibits an emission peak at 630 nm when
397
excited at 430 nm. The H-1 can be oxidatively attacked by ClO− to the imino group, which later
398
may lose the diaminomeleonitrile unit (Scheme 11). The final product exhibits a new emission
399
peak at 485 nm with the peak decrease at 630 nm.
400 401 402 403
Wang and co-workers developed a near-infrared fluorescent probe (H-2, Scheme 12) for HOCl
404
synthesized by activating the polymethine chain of a cyanine dye.[118] The reaction between H-2
405
and HOCl can result in the NIR fluorescence quenching at 774 nm. The reaction mechanism was
406
confirmed by mass spectra as electrophilic addition to the polymethine chain, then followed by
407
oxidation cleavage. Yuan and co-workers prepared a ruthenium(II) complex-based phosphorescent
408
probe (H-3) for HOCl.[119] In the probe, the Ru(II) complex was used as the signaling unit and an
409
amide linkage as the specific reaction moiety for HOCl. The oxidation reaction of the amide
410
linkage can be promoted by HOCl, and the generated -N-Cl species can further undergo a
411
hydrolysis process to form a highly luminescent complex.
412 413
The probe H-4 is a ratiometric fluorescence sensor for HOCl based on FRET form coumarin
414
moiety to rhodamine moiety.[120] In the absence of HOCl/ClO−, only donor (coumarin moiety)
415
fluorescence emission can be observed at 470 nm. After interacted with HOCl/ClO−, the
416
thiohydrazide spiro-ring opening reaction was occurred, meanwhile the fluorescence intensity at
417
470 nm diminished and a new emission of the acceptor at 580 nm appeared. The probe H-5
418
contains a coumarin fluorophor and an arylseleno moiety.[121] The HOCl/ClO− sensing
419
mechanism is based on the selenoxide elimination followed by a significant fluorescent turn-on
420
signal. It can rapidly respond to HOCl/ClO− within seconds with well selectivity.
421 422
2.6 Peroxynitrite (ONOO−) probes
423
Peroxynitrite (ONOO−) is formed from the reaction of nitric oxide (•NO) with superoxide O2•- in
424
inflammatory cells such as neutrophils and macrophages. The well-controlled generation of 14
Page 14 of 43
425
ONOO− is involved in cell signal transduction. As a strong oxidant, the ONOO− can also react
426
directly with a wide array of biomolecules (e.g., proteins and DNA). Therefore, sensitive and
427
selective methods (such as using fluorescent and luminescent probes) have been developed for
428
better understanding of the role played by ONOO− in cellular functions.
429
Recent investigations have shown that some boronate-containing fluorogenic compounds can
430
selectively react with ONOO− to yield corresponding hydroxyl derivatives.[122-124] Based on
431
such reaction, several fluorescent probes were designed for detecting ONOO−. Han and
432
co-workers synthesized an ONOO−probe (P-1) composed of pyrene dye and a dioxaborolane
433
group (Scheme 13).[123] In the faintly fluorescent P-1, the boronate, with sp2-hybridized boron
434
atom, is very intensely Lewis acidic. The ONOO− as Lewis base can attack the boron atom eagerly
435
and form a peroxyborate intermediate. Then aryl migration and quantitative hydrolysis give the
436
fluorescent phenol product. Another boronate-based fluorescent probe (P-2) was synthesized by
437
Kim and co-workers.[124] The boronate probe P-2 contains a phenol moiety masked by a
438
p-dihydroxyborylbenzyloxy reaction site, and displays a low fluorescence quantum yield. After
439
the reaction with ONOO−, the arylboronate group will be oxidized to its corresponding phenol,
440
which would undergo rapid elimination of p-quinomethane to produce fluorescent phenolate.
441 442
The P-3 is a ruthenium(II) complex-based fluorescent sensor for ONOO−.[125] The
443
aryloxyphenol group of P-3 was used as the ONOO− accepting unit. The reaction mechanism
444
between P-3 and ONOO− was confirmed by MS and NMR spectra. The addition of to the P-3
445
solution can result in distinct fluorescence quenching at 600 nm due to the O-dealkylation reaction.
446
The probe shows favorable water-solubility, biocompatibility and rapid reaction with ONOO−.
447
James and co-workers developed a probe via self-assembling aromatic boronic acids with
448
alizarin red S (ARS) for colorimetric and fluorometric detection of ONOO−.[126] A boronic acid,
449
2-(N,N-dimethylaminomethyl)phenylboronic acid (NBA), was successfully assembled with ARS
450
(Scheme 14). The assembling of ARS with NBA results in two species (one major and one minor
451
with approximately 2 : 1 ratio), which were used as ONOO− probe (P-4). The ONOO− can react
452
with boronate-based P-4 to produce the phenol analogues and lead to the release of ARS,
453
meanwhile giving a fluorescence decrease and color change.
454 15
Page 15 of 43
455
Yoon and co-workers designed and synthesized another colorimetric and fluorometric probe P-5
456
(Scheme 15) for ONOO−.[127] The P-5 contains a hybrid coumarin–hemicyanine scaffold. The
457
interaction between P-5 and ONOO− was monitored by using NMR and mass spectrometry, which
458
shows
459
coumarin-3-aldehyde. In this process, the emission peak of P-5 at 635 nm decreases in concert
460
with an increase of a new and more emissive band at 515 nm, and the color of the solution was
461
changed from blue violet to faint yellow.
the
final
products
consisting
of
predominantly
1,3,3-trimethyloxiindole
and
462 463
Yang and co-workers designed and synthesized a rhodamine-based ONOO− probe (P-6, Fig.
464
12A).[128] The probe P-6 was non-fluorescent. In the presence of ONOO−, the ONOO− triggered
465
N-dearylation reaction can be occurred followed by fluorescence turn-on response. The P-6
466
derivative (P-7, Fig. 12B) with a caged carboxylate is neutral and can readily diffuse across the
467
cell membrane. Therefore, the P-7 was applied to image the generation of endogenous ONOO− in
468
living cells and issues. As shown in Fig. 12C, the P-7 exhibited high sensitivity and selectivity
469
toward ONOO−.
470 471 472
Sevral chemiluminescence probes for ONOO− were recently developed by Lu and
473
co-workers.[129, 130] Typically, the thioglycolic acid (TGA)-capped CdTe QDs was used for
474
specific detection of ONOO− in living cells.[130] Generally, the ONOO− can decompose into
475
oxidizing and reducing radical pair, which can interact with QDs to produce the CL emissions by
476
electron-transfer annihilation. It was found that the oxidizing radical •OH from ONOOH can
477
inject a hole into the valence band (VB) of the CdTe QDs to produce the oxidized QDs (QDs +).
478
Then, electron-transfer annihilation between QDs + and O2•- (form ONOO−) was occurred followed
479
by light emission. This probe features an excellent selectivity for ONOO−, and it is the first
480
chemiluminescence probe available for the detection of ONOO− in living cells.
•
•
481 482
2.7 Multi-ROS probes
483
Multiple ROS including (H2O2, 1O2, O2•-, ClO−, ONOO− and •OH) generally coexist in
484
physiological progresses. Because the total ROS level has been considered to be one of the major 16
Page 16 of 43
485
characteristics of many diseases, several methods for detecting ROS level have been recently
486
developed using fluorophores, luminophores and quantum dots.[131-134] Over the past decades
487
years, one of the most commonly used fluorescent probes for ROS is the non-fluorescent
488
2',7'-dichlorodihydrofluorescein diacetate (DCFH-DA).[135, 136] DCFH-DA can easily pass the
489
cell membrane and is cleaved by intracellular esterases to 2',7'-dichlorodihydrofluorescein
490
(DCFH). The DCFH is then oxidized by ROS to the highly fluorescent dichlorofluorescein (DCF).
491
Recent years, the research attentions mainly focus on designing nano-probes for ROS level. The
492
Au nanoclusters (AuNCs) synthesized using glutathione template have been used as biosensing
493
substrates for ROS sensing.[137, 138] Typically, Chu and co-workers developed a nanocomplex
494
displaying single-excitation and dual-emission fluorescent properties towards highly reactive
495
oxygen species (hROS), including ClO−, ONOO− and •OH.[138] The nanocomplex is an
496
AuNC-decorated silica particle, in which CF 405S succinimidyl ester (an amine-reactive
497
fluorescent dye with a strong and photostable emission peak) is located in silica nanoparticle and
498
the glutathione templated AuNCs are grafted on the silica surface via streptavidin (SA)-biotin (Fig.
499
13A and B). The nanosensor exhibits two fluorescence peaks: one located at 435 nm arising from
500
CF 405S, and the other at 565 nm originating from the decorated AuNCs. It can selectively react
501
with hROS followed by fluorescence quenching at 565 nm, and shows excellent biocompatibility
502
for sensing of hROS in living cells (Fig. 13C and D).
503 504
The AuNCs synthesized using glutathione template can also assembled with C-dots as hROS
505
sensor.[139] When the AuNCs and C-dots are fabricated in nanoparticle (termed as C-dots-AuNC),
506
the fluorescence of AuNCs was enhanced while the fluorescence of C-dots was stabilized.
507
Interestingly, the fluorescence of AuNCs can be quenched in response to hROS, but the
508
fluorescence from the C-dots negatively decreased. Therefore, such dual-emission property of
509
C-dots-AuNC allows sensitive imaging and monitoring of the hROS signal (Fig. 14).
510 511
There have been several works reported by using C-dots for directly sensing ROS.[140, 141]
512
The phosphorus and nitrogen doped carbon dots (PN-CDs) were prepared by carbonization of
513
adenosine-5′-triphosphate using a hydrothermal treatment.[141] The PN-CDs shows rapid
514
fluorescence quenching in the presence of ROS (especially the ClO−), and can be applied for 17
Page 17 of 43
515
label-free, sensitive and real-time detection of ROS.
516
Fluorescent coordination polymers, a class of hybrid materials assembled from polydentate
517
bridging ligands and metal ions, have shown their excellent properties for sensing ROS. Such
518
materials possess several potential advantages such as structural and chemical diversity and their
519
intrinsic biodegradability.[142] Very recently, xia and co-workers designed a sulfur-tagged
520
europium(III) coordination polymers for monitoring ROS in living cell and aerosols.[143] The
521
probe were prepared by simply mixing the bridging ligand (2,2′-thiodiacetic acid) and Eu3+ in
522
ethanol. The product morphology can be transformed from microcrystals (at 25 ºC) to
523
nanoparticles (at 150 ºC) upon increasing reaction temperature. In the presence of ROS, the
524
emission peaks of Eu3+ are quenched due to the thioether groups of the ligand are oxidized to
525
sulfoxide followed by intramolecular charge transfer from sulfoxide to Eu3+ (Fig. 15).
526 527 528
3.
Conclusion
529
The development of fluorescent and luminescent probes for ROS in recent three years was
530
discussed in this review. Up to now, the great challenge of designing ideal ROS probe is still the
531
selectivity and sensitivity when used in real complex systems, especially in biological samples and
532
living cells. In the living systems, the low concentrations and short lifetimes have hampered the
533
detection of most of ROS, including H2O2, 1O2, O2•-, ClO−, ONOO− and •OH. Furthermore, the
534
traditional fluorescent probes generally have some shortcomings (e.g., background interference,
535
easy to be photobleached and so on). Recently, many probes based on small organic molecules or
536
metal complexes have shown good selectivity toward different kinds of ROS, such as the
537
boronate-based fluorescent probes for H2O2 and DPA-based fluorescent probes for
538
Nevertheless, compared with the probes for H2O2, there are few probes specially designed for
539
sensing 1O2, O2•-, ClO− or ONOO− in recent 3 years.
1
O2.
540
The emerged nanomaterials have provided promising sensing platforms for ROS, because of
541
their unique optical and catalytic properties for translating the biorecognition events to
542
spectroscopic responses. Modification of the surface of such nanomaterials with biomolecules,
543
such as antibodies or peptides, can reduce their cytotoxicity, facilitate their internalization into
544
cells. Therefore, the multifunctional fluorescent nano-probes have shown high potential in the 18
Page 18 of 43
545
field of intracellular ROS sensing. Furthermore, Most of the nanocomposites can be designed as
546
dual-colored ratiometric nanoprobes for more accurately monitoring ROS. And designing
547
near-infrared nano-probes will facilitate the in vivo ROS sensing with low background
548
interference and high penetrability into tissues.
549
Generally, the intracellular ROS level was very high in in pathological processes including
550
cancer, neurodegenerative injury and inflammation. In recent years, many researchers are focus on
551
designing multifunctional nanoprobes with property of ROS sensing combined with ROS-induced
552
drug release, thus offering a step toward the development of theranostic nanomedicines. Designing
553
smart nanomaterials provide a compelling approach for the future development of ROS probes for
554
bioimaging and therapeutic applications.
555 556
Acknowledgements
557
This work was supported by Jiangsu Provincial Natural Science Foundation (No. BK20150689
558
and
559
BE2016745), the Open Project Program of MOE Key Laboratory of Drug Quality Control and
560
Pharmacovigilance (No. DQCP2015QN01),the National Natural Science Foundation of China
561
(Grants 81673390) and the Fundamental Research Funds for the Central Universities (No.
562
2015PY010 and No. 2015ZD008).
No.
BK
20151445),
Jiangsu Provincial Key Research and Development Program (No.
563 564
References
565 566 567 568 569 570 571 572 573 574 575 576 577 578
[1] M. Nishikawa, Reactive oxygen species in tumor metastasis, Cancer Lett., 266 (2008) 53-59. [2] X. Wen, J. Wu, F. Wang, B. Liu, C. Huang, Y. Wei, Deconvoluting the role of reactive oxygen species and autophagy in human diseases, Free. Radical. Bio. Med., 65 (2013) 402-410. [3] K. Brieger, S. Schiavone, F.J. Miller, Jr., K.-H. Krause, Reactive oxygen species: From health to disease, Swiss Med. Weekly, 142 (2012). [4] O. Adegoke, P.B.C. Forbes, Challenges and advances in quantum dot fluorescent probes to detect reactive oxygen and nitrogen species: A review, Anal. Chim. Acta, 862 (2015) 1-13. [5] S.I. Dikalov, D.G. Harrison, Methods for detection of mitochondrial and cellular reactive oxygen species, Antioxid. Redox Sign., 20 (2014) 372-382. [6] C.C. Winterbourn, The challenges of using fluorescent probes to detect and quantify specific reactive oxygen species in living cells, Bba-gen. Subjects., 1840 (2014) 730-738. [7] K. Bedard, K.-H. Krause, The NOX family of ROS-generating NADPH oxidases: Physiology and pathophysiology, Physiol. Rev., 87 (2007) 245-313. [8] Michael P. Murphy, How mitochondria produce reactive oxygen species, Biochem. J., 417 (2009) 19
Page 19 of 43
579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622
1-13. [9] W.M. Nauseef, Biological roles for the NOX family NADPH oxidases, J. Biol. Chem., 283 (2008) 16961-16965. [10] E.A. Robinson, J.D. Johnson, Methods for analysis of free radicals in cigarette smoke, Mini-Rev. Org. Chem., 8 (2011) 401-411. [11] J. Pranczk, D. Jacewicz, D. Wyrzykowski, L. Chmurzynski, Analytical methods for determination of reactive oxygen species, Curr. Pharm. Anal., 10 (2014) 293-304. [12] P. Fernandez-Castro, M. Vallejo, M. Fresnedo San Roman, I. Ortiz, Insight on the fundamentals of advanced oxidation processes. Role and review of the determination methods of reactive oxygen species, J. Chem. Technol. Biot., 90 (2015) 796-820. [13] W. He, Y. Liu, W.G. Wamer, J.-J. Yin, Electron spin resonance spectroscopy for the study of nanomaterial-mediated generation of reactive oxygen species, J. Food Drug Anal., 22 (2014) 49-63. [14] S.L. Baum, I.G.M. Anderson, R.R. Baker, D.M. Murphy, C.C. Rowlands, Electron spin resonance and spin trap investigation of free radicals in cigarette smoke: Development of a quantification procedure, Anal. Chim. Acta, 481 (2003) 1-13. [15] X.F. Yang, X.Q. Guo, Investigation of the anthracene-nitroxide hybrid molecule as a probe for hydroxyl radicals, Analyst., 126 (2001) 1800-1804. [16] W. He, X. Liu, H. Xu, Y. Gong, F. Yuan, Y. Gao, On-line HPLC-ABTS screening and HPLC-DAD-MS/MS identification of free radical scavengers in gardenia (gardenia jasminoides ellis) fruit extracts, Food Chem., 123 (2010) 521-528. [17] A. Sikora, J. Zielonka, M. Lopez, A. Dybala-Defratyka, J. Joseph, A. Marcinek, B. Kalyanaraman, Reaction between peroxynitrite and boronates: EPR spin-trapping, HPLC analyses, and quantum mechanical study of the free radical pathway, Chem. Res. Toxicol., 24 (2011) 687-697. [18] J.-F. Jen, M.-F. Leu, T.C. Yang, Determination of hydroxyl radicals in an advanced oxidation process with salicylic acid trapping and liquid chromatography, J. Chromatogr. A, 796 (1998) 283-288. [19] H. Kaur, B. Halliwell, Aromatic hydroxylation of phenylalanine as an assay for hydroxyl radicals: Measurement of hydroxyl radical formation from ozone and in blood from premature babies using improved HPLC methodology, Anal. Biochem., 220 (1994) 11-15. [20] L. Ste-Marie, D. Boismenu, L. Vachon, J. Montgomery, Evaluation of sodium 4-hydroxybenzoate as an hydroxyl radical trap using gas chromatography–mass spectrometry and high-performance liquid chromatography with electrochemical detection, Anal. Biochem., 241 (1996) 67-74. [21] D. Lu, A. Cagan, R.A.A. Munoz, T. Tangkuaram, J. Wang, Highly sensitive electrochemical detection of trace liquid peroxide explosives at a prussian-blue 'artificial-peroxidase' modified electrode, Analyst., 131 (2006) 1279-1281. [22] H.M. Cocheme, A. Logan, T.A. Prime, I. Abakumova, C. Quin, S.J. McQuaker, J.V. Patel, I.M. Fearnley, A.M. James, C.M. Porteous, R.A.J. Smith, R.C. Hartley, L. Partridge, M.P. Murphy, Using the mitochondria-targeted ratiometric mass spectrometry probe mitob to measure H2O2 in living drosophila, Nat. Protocols, 7 (2012) 946-958. [23] A.R. Lippert, G.C.V. De Bittner, C.J. Chang, Boronate oxidation as a bioorthogonal reaction approach for studying the chemistry of hydrogen peroxide in living systems, Accounts. Chem. Res., 44 (2011) 793-804. [24] Z. Lou, P. Li, K. Han, Redox-responsive fluorescent probes with different design strategies, Accounts. Chem. Res., 48 (2015) 1358-1368. [25] S. Wang, N. Li, W. Pan, B. Tang, Advances in functional fluorescent and luminescent probes for 20
Page 20 of 43
623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666
imaging intracellular small-molecule reactive species, Trac-trend. Anal. Chem., 39 (2012) 3-37. [26] B.C. Dickinson, C.J. Chang, Chemistry and biology of reactive oxygen species in signaling or stress responses, Nat. Chem. Biol., 7 (2011) 504-511. [27] K.M. Robinson, M.S. Janes, M. Pehar, J.S. Monette, M.F. Ross, T.M. Hagen, M.P. Murphy, J.S. Beckman, Selective fluorescent imaging of superoxide in vivo using ethidium-based probes, Proc. Natl Acad. Sci. USA, 103 (2006) 15038-15043. [28] R.P.S. de Campos, J.M. Siegel, C.G. Fresta, G. Caruso, J.A.F. da Silva, S.M. Lunte, Indirect detection of superoxide in raw 264.7 macrophage cells using microchip electrophoresis coupled to laser-induced fluorescence, Anal. Bioanal.Chem., 407 (2015) 7003-7012. [29] K.M. Robinson, M.S. Janes, J.S. Beckman, The selective detection of mitochondrial superoxide by live cell imaging, Nat. Protocols, 3 (2008) 941-947. [30] S. Dikalov, K.K. Griendling, D.G. Harrison, Measurement of reactive oxygen species in cardiovascular studies, Hypertension, 49 (2007) 717-727. [31] H. Maeda, K. Yamamoto, Y. Nomura, I. Kohno, L. Hafsi, N. Ueda, S. Yoshida, M. Fukuda, Y. Fukuyasu, Y. Yamauchi, N. Itoh, A design of fluorescent probes for superoxide based on a nonredox mechanism, J. Am. Chem. Soc., 127 (2004) 68-69. [32] W. Zhang, P. Li, F. Yang, X. Hu, C. Sun, W. Zhang, D. Chen, B. Tang, Dynamic and reversible fluorescence imaging of superoxide anion fluctuations in live cells and in vivo, J. Am. Chem. Soc., 135 (2013) 14956-14959. [33] W. Zhang, X. Wang, P. Li, F. Huang, H. Wang, W. Zhang, B. Tang, Elucidating the relationship between superoxide anion levels and lifespan using an enhanced two-photon fluorescence imaging probe, Chem. Commun., 51 (2015) 9710-9713. [34] J.H. Chen, C.-T. Ho, Antioxidant activities of caffeic acid and its related hydroxycinnamic acid compounds, J. Agr. Food Chem., 45 (1997) 2374-2378. [35] X. Gao, C. Ding, A. Zhu, Y. Tian, Carbon-dot-based ratiometric fluorescent probe for imaging and biosensing of superoxide anion in live cells, Anal. Chem., 86 (2014) 7071-7078. [36] Y. Zhou, J. Ding, T. Liang, E.S. Abdel-Halim, L. Jiang, J.-J. Zhu, Fitc doped rattle-type silica colloidal particle-based ratiometric fluorescent sensor for biosensing and imaging of superoxide anion, ACS Appl. Mater. Interfaces, 8 (2016) 6423-6430. [37] H. Li, Q. Li, X. Wang, K. Xu, Z. Chen, X. Gong, X. Liu, L. Tong, B. Tang, Simultaneous determination of superoxide and hydrogen peroxide in macrophage raw 264.7 cell extracts by microchip electrophoresis with laser-induced fluorescence detection, Anal. Chem., 81 (2009) 2193-2198. [38] E.W. Miller, O. Tulyathan, E.Y. Isacoff, C.J. Chang, Molecular imaging of hydrogen peroxide produced for cell signaling, Nat. Chem. Biol., 3 (2007) 263-267. [39] H. Guo, H. Aleyasin, B.C. Dickinson, R.E. Haskew-Layton, R.R. Ratan, Recent advances in hydrogen peroxide imaging for biological applications, Cell Biosci., 4 (2014) 64. [40] P. Wang, K. Wang, D. Chen, Y. Mao, Y. Gu, A novel colorimetric and near-infrared fluorescent probe for hydrogen peroxide imaging in vitro and in vivo, Rsc Adv., 5 (2015) 85957-85963. [41] B.C. Dickinson, C. Huynh, C.J. Chang, A palette of fluorescent probes with varying emission colors for imaging hydrogen peroxide signaling in living cells, J. Am. Chem. Soc., 132 (2010) 5906-5915. [42] N. Karton-Lifshin, E. Segal, L. Omer, M. Portnoy, R. Satchi-Fainaro, D. Shabat, A unique paradigm for a turn-on near-infrared cyanine-based probe: Noninvasive intravital optical imaging of hydrogen peroxide, J. Am. Chem. Soc., 133 (2011) 10960-10965. [43] M. Xu, J.-M. Han, Y. Zhang, X. Yang, L. Zang, A selective fluorescence turn-on sensor for trace 21
Page 21 of 43
667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710
vapor detection of hydrogen peroxide, Chem. Commun., 49 (2013) 11779-11781. [44] A. Matsumoto, R. Nishiyabu, Y. Kubo, Synthesis of a borylated boron-dibenzopyrromethene dye enabling the visual detection of H2O2 vapor, Rsc Adv., 4 (2014) 37973-37978. [45] M.S. Purdey, J.G. Thompson, T.M. Monro, A.D. Abell, E.P. Schartner, A dual sensor for ph and hydrogen peroxide using polymer-coated optical fibre tips, Sensors, 15 (2015) 31904-31913. [46] W. Sun, Z. Ma, J. Li, W. Li, L. Du, M. Li, Bifunctional fluorescent probes for hydrogen peroxide and diols based on a 1,8-naphthalimide fluorophore, Sci. China Chem., 56 (2013) 1440-1445. [47] C.Y. Ang, S.Y. Tan, S. Wu, Q. Qu, M.F.E. Wong, Z. Luo, P.-Z. Li, S.T. Selvan, Y. Zhao, A dual responsive "turn-on'' fluorophore for orthogonal selective sensing of biological thiols and hydrogen peroxide, J. Mater. Chem. C, 4 (2016) 2761-2774. [48] B.C. Dickinson, V.S. Lin, C.J. Chang, Preparation and use of MitoPY1 for imaging hydrogen peroxide in mitochondria of live cells, Nat. Protoc., 8 (2013) 1249-1259. [49] W. Zhang, W. Liu, P. Li, F. Huang, H. Wang, B. Tang, Rapid-response fluorescent probe for hydrogen peroxide in living cells based on increased polarity of C-B bonds, Anal. Chem., 87 (2015) 9825-9828. [50] X. Zhang, L. Zhang, Y. Liu, B. Bao, Y. Zang, J. Li, W. Lu, A near-infrared fluorescent probe for rapid detection of hydrogen peroxide in living cells, Tetrahedron, 71 (2015) 4842-4845. [51] M. Ren, B. Deng, J.-Y. Wang, X. Kong, Z.-R. Liu, K. Zhou, L. He, W. Lin, A fast responsive two-photon fluorescent probe for imaging H2O2 in lysosomes with a large turn-on fluorescence signal, Biosens. Bioelectron., 79 (2016) 237-243. [52] Y. Wen, K. Liu, H. Yang, Y. Liu, L. Chen, Z. Liu, C. Huang, T. Yi, Mitochondria-directed fluorescent probe for the detection of hydrogen peroxide near mitochondria! DNA, Anal. Chem., 87 (2015) 10579-10584. [53] R. Bortolozzi, S. von Gradowski, H. Ihmels, K. Schaefer, G. Viola, Selective ratiometric detection of H2O2 in water and in living cells with boronobenzo[b]quinolizinium derivatives, Chem. Commun., 50 (2014) 8242-8245. [54] S.W. Lee, H.-W. Rhee, Y.-T. Chang, J.-I. Hong, Ratiometric fluorescent probes for hydrogen peroxide from a focused library, Chem-eur. J., 19 (2013) 14791-14794. [55] Y. Wen, K. Liu, H. Yang, Y. Li, H. Lan, Y. Liu, X. Zhang, T. Yi, A highly sensitive ratiometric fluorescent probe for the detection of cytoplasmic and nuclear hydrogen peroxide, Anal. Chem., 86 (2014) 9970-9976. [56] B. Zhu, H. Jiang, B. Guo, C. Shao, H. Wu, B. Du, Q. Wei, A highly selective ratiometric fluorescent probe for hydrogen peroxide displaying a large emission shift, Sensor Actuat. B-chem., 186 (2013) 681-686. [57] G.C. Van de Bittner, E.A. Dubikovskaya, C.R. Bertozzi, C.J. Chang, In vivo imaging of hydrogen peroxide production in a murine tumor model with a chemoselective bioluminescent reporter, Proc. Natl Acad. Sci. USA, 107 (2010) 21316-21321. [58] L. Yuan, W. Lin, S. Zhao, W. Gao, B. Chen, L. He, S. Zhu, A unique approach to development of near-infrared fluorescent sensors for in vivo imaging, J. Am. Chem. Soc., 134 (2012) 13510-13523. [59] G. Zeng, R. Zhang, W. Xuan, W. Wang, F.-S. Liang, Constructing de Novo H2O2 signaling via induced protein proximity, Acs Chemical Biology, 10 (2015) 1404-1410. [60] Y. Lei, C. Xue, S. Zhang, Y. Sha, A ratiometric fluorescent probe for sensing hydrogen peroxide based on a hemicyanine-naphthol fluorophore, Luminescence, 31 (2016) 660-664. [61] C. Liu, C. Shao, H. Wu, B. Guo, B. Zhu, X. Zhang, A fast-response, highly sensitive and selective 22
Page 22 of 43
711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 754
fluorescent probe for the ratiometric imaging of hydrogen peroxide with a 100 nm red-shifted emission, Rsc Adv., 4 (2014) 16055-16061. [62] Z. Han, X. Liang, X. Ren, L. Shang, Z. Yin, A 3,7-dihydroxyphenoxazine-based fluorescent probe for selective detection of intracellular hydrogen peroxide, Chem.-Asian J., 11 (2016) 818-822. [63] P. Wang, K. Wang, Y. Gu, A highly selective fluorescent turn-on NIR probe for the bioimaging of hydrogen peroxide in vitro and in vivo, Sensor Actuat. B-chem., 228 (2016) 174-179. [64] M. Xu, J.-M. Han, C. Wang, X. Yang, J. Pei, L. Zang, Fluorescence ratiometric sensor for trace vapor detection of hydrogen peroxide, ACS Appl. Mater. Interfaces, 6 (2014) 8708-8714. [65] J. Xu, Q. Li, Y. Yue, Y. Guo, S. Shao, A water-soluble bodipy derivative as a highly selective "turn-on" fluorescent sensor for H2O2 sensing in vivo, Biosens. Bioelectron., 56 (2014) 58-63. [66] J. Xu, Y. Zhang, H. Yu, X. Gao, S. Shao, Mitochondria-targeted fluorescent probe for imaging hydrogen peroxide in living cells, Anal. Chem., 88 (2016) 1455-1461. [67] A. Senthamizhan, B. Balusamy, Z. Aytac, T. Uyar, Ultrasensitive electrospun fluorescent nanofibrous membrane for rapid visual colorimetric detection of H2O2, Anal. Bioanal.Chem., 408 (2016) 1347-1355. [68] Q. Zhao, S. Chen, H. Huang, L. Zhang, L. Wang, F. Liu, J. Chen, Y. Zeng, P.K. Chu, Colorimetric and ultra-sensitive fluorescence resonance energy transfer determination of H2O2 and glucose by multi-functional au nanoclusters, Analyst., 139 (2014) 1498-1503. [69] Y. Zhou, W. Pei, X. Zhang, W. Chen, J. Wu, C. Yao, L. Huang, H. Zhang, W. Huang, J.S.C. Loo, Q. Zhang, A cyanine-modified upconversion nanoprobe for NIR-excited imaging of endogenous hydrogen peroxide signaling in vivo, Biomaterials., 54 (2015) 34-43. [70] J. Yuan, Y. Cen, X.-J. Kong, S. Wu, C.-L. Liu, R.-Q. Yu, X. Chu, MnO2-nanosheet-modified upconversion nanosystem for sensitive turn-on fluorescence detection of H2O2 and glucose in blood, ACS Appl. Mater. Interfaces, 7 (2015) 10548-10555. [71] H. Tan, C. Ma, Q. Li, L. Wang, F. Xu, S. Chen, Y. Song, Functionalized lanthanide coordination polymer nanoparticles for selective sensing of hydrogen peroxide in biological fluids, Analyst., 139 (2014) 5516-5522. [72] J. Qiao, Z. Liu, Y. Tian, M. Wu, Z. Niu, Multifunctional self-assembled polymeric nanoprobes for FRET-based ratiometric detection of mitochondrial H2O2 in living cells, Chem. Commun., 51 (2015) 3641-3644. [73] W.-K. Oh, Y.S. Jeong, S. Kim, J. Jang, Fluorescent polymer nanoparticle for selective sensing of intracellular hydrogen peroxide, Acs Nano, 6 (2012) 8516-8524. [74] F. Du, Y. Min, F. Zeng, C. Yu, S. Wu, A targeted and FRET- based ratiometric fluorescent nanoprobe for imaging mitochondrial hydrogen peroxide in living cells, Small, 10 (2014) 964-972. [75] F. Pan, P. Wei, M. Zhang, C. Lu, Micelle modified-carbon nanosphere enhanced chemiluminescence from reactive oxygen species for the detection of hydrogen peroxide, Anal. Methods, 7 (2015) 5667-5673. [76] J. Wei, L. Qiang, J. Ren, X. Ren, F. Tang, X. Meng, Fluorescence turn-off detection of hydrogen peroxide and glucose directly using carbon nanodots as probes, Anal. Methods, 6 (2014) 1922-1927. [77] C.-I. Wang, A.P. Periasamy, H.-T. Chang, Photoluminescent C-dots@RGO probe for sensitive and selective detection of acetylcholine, Anal. Chem., 85 (2013) 3263-3270. [78] M. Lan, Y. Di, X. Zhu, T.-W. Ng, J. Xia, W. Liu, X. Meng, P. Wang, C.-S. Lee, W. Zhang, A carbon dot-based fluorescence turn-on sensor for hydrogen peroxide with a photo-induced electron 23
Page 23 of 43
755 756 757 758 759 760 761 762 763 764 765 766 767 768 769 770 771 772 773 774 775 776 777 778 779 780 781 782 783 784 785 786 787 788 789 790 791 792 793 794 795 796 797 798
transfer mechanism, Chem. Commun., 51 (2015) 15574-15577. [79] W. Gao, X. Wei, X. Wang, G. Cui, Z. Liu, B. Tang, A competitive coordination-based CeO2 nanowire-DNA nanosensor: Fast and selective detection of hydrogen peroxide in living cells and in vivo, Chem. Commun., 52 (2016) 3643-3646. [80] R. Pautler, E.Y. Kelly, P.-J.J. Huang, J. Cao, B. Liu, J. Liu, Attaching DNA to nanoceria: Regulating oxidase activity and fluorescence quenching, ACS Appl. Mater. Interfaces, 5 (2013) 6820-6825. 3+
[81] W. Gao, J. Li, X. Zhou, Z. Zhang, Y. Ma, Y. Qu, Repeatable fluorescence switcher of Eu -doped CeO2 nanorods by L(+)-ascorbic acid and hydrogen peroxide, J. Mater. Chem. C, 2 (2014) 8729-8735. [82] C. Lv, W. Di, Z. Liu, K. Zheng, W. Qin, Luminescent cepo4: Tb colloids for H2O2 and glucose sensing, Analyst., 139 (2014) 4547-4555. [83] L. Wang, J. Zheng, Y. Li, S. Yang, C. Liu, Y. Xiao, J. Li, Z. Cao, R. Yang, AGNP-DNA@GQDS hybrid: New approach for sensitive detection of H2O2 and glucose via simultaneous AGNP etching and DNA cleavage, Anal. Chem., 86 (2014) 12348-12354. [84] H. Miao, D. Zhong, Z. Zhou, X. Yang, Papain-templated Cu nanoclusters: Assaying and exhibiting dramatic antibacterial activity cooperating with H2O2, Nanoscale, 7 (2015) 19066-19072. [85] L.J. Zhang, N. He, C. Lu, Aggregation-induced emission: A simple strategy to improve chemiluminescence resonance energy transfer, Anal. Chem., 87 (2015) 1351-1357. [86] S.C. Dong, F. Liu, C. Lu, Organo-modified hydrotalcite-quantum dot nanocomposites as a novel chemiluminescence resonance energy transfer probe, Anal. Chem., 85 (2013) 3363-3368. [87] H. Zhao, J. Gao, W. Zhou, Z. Wang, S. Wu, Quantitative detection of hydroxyl radicals in fenton system by Uv-Vis spectrophotometry, Anal. Methods, 7 (2015) 5447-5453. [88] Y. You, W. Nam, Designing photoluminescent molecular probes for singlet oxygen, hydroxyl radical, and iron-oxygen species, Chem. Sci., 5 (2014) 4123-4135. [89] S. Liu, J. Zhao, K. Zhang, L. Yang, M. Sun, H. Yu, Y. Yan, Y. Zhang, L. Wu, S. Wang, Dual-emissive fluorescence measurements of hydroxyl radicals using a coumarin-activated silica nanohybrid probe, Analyst., 141 (2016) 2296-2302. [90]
N.V.
Blough,
D.J.
Simpson,
Chemically
mediated
fluorescence
yield
switching
in
nitroxide-fluorophore adducts: Optical sensors of radical/redox reactions, J. Am. Chem. Soc., 110 (1988) 1915-1917. [91] L. Cao, Q. Wu, Q. Li, S. Shao, Y. Guo, Fluorescence and hplc detection of hydroxyl radical by a rhodamine-nitroxide probe and its application in cell imaging, J. Fluoresc., 24 (2014) 313-318. [92] M. Kim, S.-K. Ko, H. Kim, I. Shin, J. Tae, Rhodamine cyclic hydrazide as a fluorescent probe for the detection of hydroxyl radicals, Chem. Commun., 49 (2013) 7959-7961. [93] K.L. Peterson, M.J. Margherio, P. Doan, K.T. Wilke, V.C. Pierre, Basis for sensitive and selective time-delayed luminescence detection of hydroxyl radical by lanthanide complexes, Inorg. Chem., 52 (2013) 9390-9398. [94] L. Meng, Y. Wu, T. Yi, A ratiometric fluorescent probe for the detection of hydroxyl radicals in living cells, Chem. Commun., 50 (2014) 4843-4845. [95] H. Xu, K.S. Suslick, Sonochemical synthesis of highly fluorescent ag nanoclusters, ACS Nano, 4 (2010) 3209-3214. [96] J. Zheng, P.R. Nicovich, R.M. Dickson, Highly fluorescent noble-metal quantum dots, Annu. Rev. Phys. Chem., 58 (2007) 409-431. [97] L. Zhang, R.-P. Liang, S.-J. Xiao, J.-M. Bai, L.-L. Zheng, L. Zhan, X.-J. Zhao, J.-D. Qiu, C.-Z. Huang, DNA-templated Ag nanoclusters as fluorescent probes for sensing and intracellular imaging of 24
Page 24 of 43
799 800 801 802 803 804 805 806 807 808 809 810 811 812 813 814 815 816 817 818 819 820 821 822 823 824 825 826 827 828 829 830 831 832 833 834 835 836 837 838 839 840 841 842
hydroxyl radicals, Talanta, 118 (2014) 339-347. [98] M. Zhuang, C. Ding, A. Zhu, Y. Tian, Ratiometric fluorescence probe for monitoring hydroxyl radical in live cells based on gold nanoclusters, Anal. Chem., 86 (2014) 1829-1836. [99] X. Ren, J. Liu, X. Meng, J. Wei, T. Liu, F. Tang, Synthesis of ultra-stable fluorescent carbon dots from polyvinylpyrrolidone and their application in the detection of hydroxyl radicals, Chem.-Asian J., 9 (2014) 1054-1059. [100] P. Wang, F. Qin, Z. Zhang, W. Cao, Quantitative monitoring of the level of singlet oxygen using luminescence spectra of phosphorescent photosensitizer, Opt. Express, 23 (2015) 22991-23003. [101] Y. Shen, H.Y. Lin, Z.F. Huang, D.F. Chen, B.H. Li, S.S. Xie, Indirect imaging of singlet oxygen generation from a single cell, Laser Phys. Lett., 8 (2011) 232-238. [102] A.P. Castano, P. Mroz, M.R. Hamblin, Photodynamic therapy and anti-tumour immunity, Nat. Rev. Cancer, 6 (2006) 535-545. [103] H. Wu, Q. Song, G. Ran, X. Lu, B. Xu, Recent developments in the detection of singlet oxygen with molecular spectroscopic methods, Trac-trend. Anal. Chem., 30 (2011) 133-141. [104] B. Song, G. Wang, M. Tan, J. Yuan, A europium(III) complex as an efficient singlet oxygen luminescence probe, J. Am. Chem. Soc., 128 (2006) 13442-13450. [105] M. Higuchi, S. Shiki, K. Ariga, K. Yamamoto, First synthesis of phenylazomethine dendrimer ligands and structural studies, J. Am. Chem. Soc., 123 (2001) 4414-4420. [106] S.K. Pedersen, J. Holmehave, F.H. Blaikie, A. Gollmer, T. Breitenbach, H.H. Jensen, P.R. Ogilby, Aarhus sensor green: A fluorescent probe for singlet oxygen, J. Org. Chem., 79 (2014) 3079-3087. [107] R. Ruiz-González, R. Zanocco, Y. Gidi, A.L. Zanocco, S. Nonell, E. Lemp, Naphthoxazole-based singlet oxygen fluorescent probes, Photochem. Photobiol., 89 (2013) 1427-1432. 1
[108] D. Arian, L. Kovbasyuk, A. Mokhir, 1,9-dialkoxyanthracene as a O2-sensitive linker, J. Am. Chem. Soc., 133 (2011) 3972-3980. [109] S. Kim, T. Tachikawa, M. Fujitsuka, T. Majima, Far-red fluorescence probe for monitoring singlet oxygen during photodynamic therapy, J. Am. Chem. Soc., 136 (2014) 11707-11715. [110] Z. Dai, L. Tian, Y. Xiao, Z. Ye, R. Zhang, J. Yuan, A cell-membrane-permeable europium complex as an efficient luminescent probe for singlet oxygen, J. Mater. Chem. B, 1 (2013) 924-927. [111] Y. You, E.J. Cho, H. Kwon, J. Hwang, S.E. Lee, A singlet oxygen photosensitizer enables photoluminescent monitoring of singlet oxygen doses, Chem. Commun., 52 (2016) 780-783. [112] F.-J. Huo, J.-J. Zhang, Y.-T. Yang, J.-B. Chao, C.-X. Yin, Y.-B. Zhang, T.-G. Chen, A fluorescein-based highly specific colorimetric and fluorescent probe for hypochlorites in aqueous solution and its application in tap water, Sensors and Actuators B: Chemical, 166–167 (2012) 44-49. [113] W. Lin, L. Long, B. Chen, W. Tan, A ratiometric fluorescent probe for hypochlorite based on a deoximation reaction, Chem-eur. J., 15 (2009) 2305-2309. [114] X.-Q. Zhan, J.-H. Yan, J.-H. Su, Y.-C. Wang, J. He, S.-Y. Wang, H. Zheng, J.-G. Xu, Thiospirolactone as a recognition site: Rhodamine B-based fluorescent probe for imaging hypochlorous acid generated in human neutrophil cells, Sensor Actuat. B-chem., 150 (2010) 774-780. [115] S. Chen, J. Lu, C. Sun, H. Ma, A highly specific ferrocene-based fluorescent probe for hypochlorous acid and its application to cell imaging, Analyst., 135 (2010) 577-582. [116] X. Chen, X. Tian, I. Shin, J. Yoon, Fluorescent and luminescent probes for detection of reactive oxygen and nitrogen species, Chem. Soc. Rev., 40 (2011) 4783-4804. [117] S. Goswami, K. Aich, S. Das, B. Pakhira, K. Ghoshal, C.K. Quah, M. Bhattacharyya, H.K. Fun, S. Sarkar, A triphenyl amine-based solvatofluorochromic dye for the selective and ratiometric sensing 25
Page 25 of 43
843 844 845 846 847 848 849 850 851 852 853 854 855 856 857 858 859 860 861 862 863 864 865 866 867 868 869 870 871 872 873 874 875 876 877 878 879 880 881 882 883 884 885 886
-
of OCl in human blood cells, Chem.-Asian J., 10 (2015) 694-700. [118] M. Sun, H. Yu, H. Zhu, F. Ma, S. Zhang, D. Huang, S. Wang, Oxidative cleavage-based near-infrared fluorescent probe for hypochlorous acid detection and myeloperoxidase activity evaluation, Anal. Chem., 86 (2014) 671-677. [119] R. Zhang, B. Song, Z. Dai, Z. Ye, Y. Xiao, Y. Liu, J. Yuan, Highly sensitive and selective phosphorescent chemosensors for hypochlorous acid based on ruthenium(II) complexes, Biosens. Bioelectron., 50 (2013) 1-7. [120] Y.-R. Zhang, Z.-M. Zhao, L. Su, J.-Y. Miao, B.-X. Zhao, A ratiometric fluorescence sensor for hocl based on a fret platform and application in living cells, Rsc Adv., 6 (2016) 17059-17063. [121] G. Li, D. Zhu, Q. Liu, L. Xue, H. Jiang, A strategy for highly selective detection and imaging of hypochlorite using selenoxide elimination, Org. Lett., 15 (2013) 2002-2005. [122] J. Zielonka, A. Sikora, M. Hardy, J. Joseph, B.P. Dranka, B. Kalyanaraman, Boronate probes as diagnostic tools for real time monitoring of peroxynitrite and hydroperoxides, Chem. Res. Toxicol., 25 (2012) 1793-1799. [123] F. Yu, P. Song, P. Li, B. Wang, K. Han, A fluorescent probe directly detect peroxynitrite based on boronate oxidation and its applications for fluorescence imaging in living cells, Analyst., 137 (2012) 3740-3749. [124] J. Kim, J. Park, H. Lee, Y. Choi, Y. Kim, A boronate-based fluorescent probe for the selective detection of cellular peroxynitrite, Chem. Commun., 50 (2014) 9353-9356. [125] J. Ma, J. Wu, W. Liu, P. Wang, Z. Fan, Ruthenium(II) complex-based fluorescent sensor for peroxynitrite, Spectrochim. Acta A, 94 (2012) 340-345. [126] X. Sun, K. Lacina, E.C. Ramsamy, S.E. Flower, J.S. Fossey, X. Qian, E.V. Anslyn, S.D. Bull, T.D. James, Reaction-based indicator displacement assay (RIA) for the selective colorimetric and fluorometric detection of peroxynitrite, Chem. Sci., 6 (2015) 2963-2967. [127] X. Zhou, Y. Kwon, G. Kim, J.-H. Ryu, J. Yoon, A ratiometric fluorescent probe based on a coumarin-hemicyanine scaffold for sensitive and selective detection of endogenous peroxynitrite, Biosens. Bioelectron., 64 (2015) 285-291. [128] T. Peng, X. Chen, L. Gao, T. Zhang, W. Wang, J. Shen, D. Yang, A rationally designed rhodamine-based fluorescent probe for molecular imaging of peroxynitrite in live cells and tissues, Chem. Sci., 7 (2016) 5407-5413. [129] Z. Wang, X. Teng, C. Lu, Orderly arranged fluorescence dyes as a highly efficient chemiluminescence resonance energy transfer probe for peroxynitrite, Anal. Chem., 87 (2015) 3412-3418. [130] W. Zhou, Y. Cao, D. Sui, C. Lu, Radical pair-driven luminescence of quantum dots for specific detection of peroxynitrite in living cells, Anal. Chem., 88 (2016) 2659-2665. [131] R. Zhang, J. Zhao, G. Han, Z. Liu, C. Liu, C. Zhang, B. Liu, C. Jiang, R. Liu, T. Zhao, M.-Y. Han, Z. Zhang, Real-time discrimination and versatile profiling of spontaneous reactive oxygen species in living organisms with a single fluorescent probe, J. Am. Chem. Soc., 138 (2016) 3769-3778. [132] L. Yang, N. Li, W. Pan, Z. Yu, B. Tang, Real-time imaging of mitochondrial hydrogen peroxide and ph fluctuations in living cells using a fluorescent nanosensor, Anal. Chem., 87 (2015) 3678-3684. [133] W. Zhou, Y. Cao, D. Sui, C. Lu, Turn-on luminescent probes for the real-time monitoring of endogenous hydroxyl radicals in living cells, Angew. Chem. Int. Edit., 55 (2016) 4236-4241. [134] Y. Yuan, C.-J. Zhang, S. Xu, B. Liu, A self-reporting aie probe with a built-in singlet oxygen sensor for targeted photodynamic ablation of cancer cells, Chem. Sci., 7 (2016) 1862-1866. 26
Page 26 of 43
887 888 889 890 891 892 893 894 895 896 897 898 899 900 901 902 903 904 905 906 907 908 909
[135] M. Schaeferling, D.B.M. Groegel, S. Schreml, Luminescent probes for detection and imaging of hydrogen peroxide, Microchim. Acta, 174 (2011) 1-18. [136] L.M. Uusitalo, N. Hempel, Recent advances in intracellular and in vivo ROS sensing: Focus on nanoparticle and nanotube applications, Int. J. Mol. Sci., 13 (2012) 10660-10679. [137] S. Kumar, W.-K. Rhim, D.-K. Lim, J.-M. Nam, Glutathione dimerization-based plasmonic nanoswitch for biodetection of reactive oxygen and nitrogen species, Acs Nano, 7 (2013) 2221-2230. [138] T. Chen, Y. Hu, Y. Cen, X. Chu, Y. Lu, A dual-emission fluorescent nanocomplex of gold-cluster-decorated silica particles for live cell imaging of highly reactive oxygen species, J. Am. Chem. Soc., 135 (2013) 11595-11602. [139] E. Ju, Z. Liu, Y. Du, Y. Tao, J. Ren, X. Qu, Heterogeneous assembled nanocomplexes for ratiometric detection of highly reactive oxygen species in vitro and in vivo, Acs Nano, 8 (2014) 6014-6023. [140] B. Das, P. Dadhich, P. Pal, P.K. Srivas, K. Bankoti, S. Dhara, Carbon nanodots from date molasses: New nanolights for the in vitro scavenging of reactive oxygen species, J. Mater. Chem. B, 2 (2014) 6839-6847. [141] Y. Gong, B. Yu, W. Yang, X. Zhang, Phosphorus, and nitrogen Co-doped carbon dots as a fluorescent probe for real-time measurement of reactive oxygen and nitrogen species inside macrophages, Biosens. Bioelectron., 79 (2016) 822-828. [142] J. Della Rocca, D. Liu, W. Lin, Nanoscale metal–organic frameworks for biomedical imaging and drug delivery, Accounts. Chem. Res., 44 (2011) 957-968. [143] H.-S. Wang, W.-J. Bao, S.-B. Ren, M. Chen, K. Wang, X.-H. Xia, Fluorescent sulfur-tagged europium(III) coordination polymers for monitoring reactive oxygen species, Anal. Chem., 87 (2015) 6828-6833.
910 911
27
Page 27 of 43
912 913
Fig. 1. Carbon-dot-based fluorescent probe for imaging O2•-. (A) Working principle of the O2•- sensing. (B and C)
914
Pseudocolored ratiometric images of HeLa cells containing CD-HE probes before (B) and after (C) induced by
915
LPS (a stimulator for production of ROS). Reproduced with permission of the American Chemical Society
916
from ref.[35]
917
918 919
Fig. 2. Structures of H2O2 probes based on “boronate to phenol” conversion.
920
28
Page 28 of 43
921 922
Fig. 3. Fluorimetric monitoring H2O2 with HP-11 and HP-12. Arrows indicate the development of emission bands
923
with time (A) or with various concentrations of H2O2 (B). Reproduced with permission of The Royal Society of
924
Chemistry from ref.[53] and Wiley-VCH from ref.[54]
925
926 927
Fig. 4. Ratiometric fluorescent probes for monitoring cytoplasmic and nuclear H2O2. (A) Confocal laser scanning
928
microscopy (CLSM) ratio (RY/B) images of 5 μM HP-13-loaded HeLa cells stimulated with 200 μM H2O2 for
929
(a) 0, (b) 30, (c) 60 min at 37 °C. (A) CLSM ratio (RY/B) images of 50 μM HP-14-loaded HeLa cells stimulated
930
with 200 μM H2O2 for (a) 0, (b) 30, (c) 75 min at 37 °C. RY/B was constructed by fluorescence detection at
931
yellow channel and blue channel. Reproduced with permission of the American Chemical Society from ref.[55]
932
29
Page 29 of 43
933 934
Fig. 5. Mitochondria-targeted fluorescent probe for H2O2. (A) The probe HP-21 and its reaction with H2O2. (B)
935
The fluorescence spectra of probe HP-21 (5 μM, black line) and the reaction mixture (red line) of 5 μM probe
936
HP-21 with 50 μM H2O2. (C) Confocal fluorescence images of HeLa cells incubated with HP-21: C1,
937
probe-stained HeLa cells treated with 100 μM H2O2 for 90 min; C2, co-staining and imaged with 50 nM MT
938
DeepRed; C3, Merged images of C1 and C2. Reproduced with permission of the American Chemical Society
939
from ref.[66]
940
941 942
Fig. 6. Polymeric nanoprobes for FRET-based ratiometric detection of H2O2. (A) Schematic illustration for 30
Page 30 of 43
943
micelle-based ratiometric sensing of mitochondrial H2O2 in a living cell. (B) Fluorescence emission spectra of
944
PMT-F127 nanoprobe. (C) Confocal laser scanning microscopy (CLSM) images of HeLa cells incubated with
945
PMT-F127 micelles with the addition of 0 µM (control, C1), 50 µM (C2) and 200 µM (C3) H2O2. Reproduced
946
with permission of The Royal Society of Chemistry from ref.[72]
947
948 949
Fig. 7. Carbon dot-based fluorescence turn-on sensor for H2O2. (A) Schematic of the sensing process for H2O2. (B)
950
TEM image of the C-dots. The insets show the particle size distribution histogram (n = 60) and HRTEM image
951
of the C-dots. (C) Fluorescence spectra of the nanoprobes with the gradual addition of H2O2. Reproduced with
952
permission of The Royal Society of Chemistry from ref.[78]
953
954 955
Fig. 8. CeO2 nanowire-DNA nanosensor for H2O2. (A) Competitive coordination mechanism of the CeO2 31
Page 31 of 43
956
nanowire-DNA nanosensor for H2O2 detection. (B) TEM images of CeO2 nanowire. (C) Real-time imaging and
957
quantification of H2O2 by CeO2 nanowire-DNA nanosensor using imaging flow cytometry. (D) Real-time
958
fluorescence imaging of H2O2 production in zebrafish larvae. Reproduced with permission of The Royal
959
Society of Chemistry from ref.[79]
960
961 962
Fig. 9. AgNP/GQDs hybrid nanocomposite for H2O2 detection. (A) Schematic description of H2O2 based on
963
AgNP/GQDs. (B) and (C) TEM image of AgNP/GQDs before (B) and after (C) H2O2 addition. Reproduced
964
with permission of the American Chemical Society from ref.[83]
965
32
Page 32 of 43
966 967
Fig. 10. Ratiometric fluorescence probe for monitoring •OH in live cells based on gold nanoclusters. (A) Working
968
principle of the AuNC@HPF probe for •OH detection. (B) Reaction scheme of HPF with •OH. (C1) The
969
overlay of confocal fluorescence image and the bright-field image of Hela cells before being exposed to •OH.
970
(C2 and C3) The ratio images of Hela cells with AuNC@HPF probe after being stimulated by LPS for 45 and
971
90 min, respectively. Reproduced with permission of the American Chemical Society from ref.[98]
972
33
Page 33 of 43
973 1
974
Fig. 11. Far-red fluorescence probe for monitoring 1O2. (A) The reaction between SO-1 and O2. (B) Selectivity of
975
SO-1 toward 1O2 among other ROS. (C) Fluorescence images of HeLa cells incubated with (a) SO-1 and
976
5-ALA-induced PpIX, and with (b) SO-1, TMPyP4 and lysosome marker. Reproduced with permission of the
977
American Chemical Society from ref.[109]
978
34
Page 34 of 43
979 980
Fig. 12. Rhodamine-based fluorescent probe for molecular imaging ONOO−. (A) The reaction between P-6 and
981
ONOO− to yield a strongly fluorescent product. (B) The structure of cell membrane permeable P-7. (C) The
982
ONOO− imaging in SH-SY5Y human neuroblastoma cells. The cells were incubated with HKYellow-AM
983
firstly, and then treated with H2O2 or the indicated ROS donors for 1 h, followed by fluorescence imaging.
984
NOC-18, MSB, and SIN-1 were used to produce •NO, O2•-, and ONOO− respectively, FeTMPyP was used as an
985
ONOO− decomposition catalyst. The scale bar was 20 μm. Reproduced with permission of The Royal Society
986
of Chemistry from ref.[128]
987
35
Page 35 of 43
988 989
Fig. 13. AuNC-decorated silica particles for live cell imaging of hROS. (A) Schematic illustration for hROS by
990
AuNC-decorated silica particles. (B) TEM image of AuNC-decorated silica particles. (C) Ratiometric
991
responses of the nanosensor for hROS. (D) Confocal fluorescence microscopy images of HL-60 cells treated
992
with (a) no stimulation and (b) H2O2 for 10 min after incubating with AuNC-decorated silica particles.
993
Reproduced with permission of the American Chemical Society from ref.[138]
994
995 996
Fig. 14. C-dots-AuNC nanocomplex for hROS sensing. (A) Schematic illustration of the construction of
997
C-dots-AuNC and the working principle for detecting hROS. (B) TEM image of C-dots-AuNC (Red circles
998
and blue circles represent AuNCs and C-dots, respectively). (C) Fluorescence spectra of C-dots-AuNC in the 36
Page 36 of 43
999
presence of ClO-. Reproduced with permission of the American Chemical Society from ref.[139]
1000
1001 1002 1003
Fig. 15. Fluorescent sulfur-tagged europium(III) coordination polymers for monitoring ROS. Reproduced with permission of the American Chemical Society from ref.[143]
1004 1005
1006 1007
Scheme 1. Typical progresses of generation and transformation of the intracellular ROS.
1008
1009 1010
Scheme 2. Proposed mechanism of the oxidation of HE by O2•-.
1011 37
Page 37 of 43
1012 1013
Scheme 3.The structure and luminescence mechanism of the two-photon fluorescence imaging probes.
1014
1015 1016
Scheme 4. The reaction mechanism of the boronate-based fluorescent probes for H2O2.
1017
38
Page 38 of 43
1018 1019
Scheme 5. The reaction of the boronate-based fluorescent probes for H2O2.
1020
1021 1022
Scheme 6. The reaction of the selected boronate-based probes for H2O2.
1023
39
Page 39 of 43
1024 1025
Scheme 7. The detection mechanism for •OH by HR-1.
1026
1027 1028
Scheme 8. The reaction of fluorescent probes for •OH.
1029
1030 1031
Scheme 9. The reaction mechanism of the anthracene-based fluorescent probes for 1O2.
1032
40
Page 40 of 43
1033 1034
Scheme 10. Reactions of luminescent probes with 1O2.
1035
1036 1037
Scheme 11. Possible mechanism of the response of H1 towards HOCl/ClO−.
1038
41
Page 41 of 43
1039 1040
Scheme 12. Reactions of sensors for HOCl/ClO− detection.
1041
1042 1043
Scheme 13. Reactions of sensors for ONOO− detection.
1044
1045 1046
Scheme 14. Sensing mechanism of P-4 complex probe for ONOO−.
1047
42
Page 42 of 43
1048 1049
Scheme 15. Propose sensing mechanism of P-5 for ONOO−.
43
Page 43 of 43