Metal-Coordinated Fluorescent and Luminescent Probes For ROS andRNS

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 26

Coordination Chemistry Reviews 427 (2021) 213581

Contents lists available at ScienceDirect

Coordination Chemistry Reviews


journal homepage: www.elsevier.com/locate/ccr

Metal-coordinated fluorescent and luminescent probes for reactive


oxygen species (ROS) and reactive nitrogen species (RNS)
Nahyun Kwon a,1, Dayeh Kim a,1, K.M.K. Swamy a,b, Juyoung Yoon a,⇑
a
Department of Chemistry and Nanoscience, Ewha Womans University, Seoul 03760, South Korea
b
Department of Pharmaceutical Chemistry, V. L. College of Pharmacy, Raichur 584103, India

a r t i c l e i n f o a b s t r a c t

Article history: Reactive oxygen species (ROS) are chemically reactive species containing oxygen, which are produced
Received 8 July 2020 from molecular oxygen (O2) during vital processes occurring in humans and other living organisms.
Received in revised form 31 August 2020 These species include hydrogen peroxide (H2O2), hypochlorous acid/hypochlorite (HOCl/ClO–), hydroxyl
Accepted 31 August 2020
radicals (OH), superoxide anion radicals (O– 1
2 ) and singlet oxygen ( O2). ROS play key roles in various sig-
naling and pathological processes and are essential to human life, but overproduction of ROS by exoge-
nous stimuli is harmful because ROS can induce oxidation of DNA, proteins or lipids, resulting in cell
Keywords:
death. Therefore, unusual ROS levels can indicate ailments such as Parkinson’s and Alzheimer’s diseases,
Reactive oxygen species (ROS)
Reactive nitrogen species (RNS)
inflammation, diabetes and cancer. Reactive nitrogen species (RNS) are another group of important
Fluorescent probes for ROS chemically reactive species, which can damage cells via nitrosative stress. RNS include nitric oxide
Luminescent probes for ROS (NO), nitroxyl (HNO), nitrogen dioxide (NO2) and peroxynitrite (ONOO–). Due to their importance in
Metal coordination human life, research into fluorescent and luminescent sensing and imaging of these ROS and RNS has
been very active over the last couple of decades. Metal ions play key roles in the probes as an on–off redox
switch for photoinduced quenching and as a reaction site with ROS, RNS or luminescent cores. Metal
coordination reports the presence of analyte by changing the fluorescence intensity, lifetime, or excita-
tion/emission maxima. Redox-active metal ions can be trigger switches that control fluorescence quench-
ing effects, which can be used to sense ROS or RNS. In addition, metal ions, especially lanthanide metal
ions, can often be themselves a source of light emission. In this review, we cover ROS- and RNS-selective
fluorescent and luminescent probes based on metal-coordinated systems. This review is organized by the
target ROS or RNS, which are H2O2, HOCl/ClO–, OH, O– 1 –
2 , O2, NO, ONOO , HNO and NO2.
Ó 2020 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2. Metal-coordinated fluorescent and luminescent probes for reactive oxygen species (ROS) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.1. Fluorescent and luminescent probes for hydrogen peroxide (H2O2) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.2. Fluorescent and luminescent probes for hydroxyl radicals (OH). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.3. Fluorescent and luminescent probes for singlet oxygen (1O2) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.4. Fluorescent and luminescent probes for superoxide anion radical (O2–) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.5. Fluorescent and luminescent probes for hypochlorous acid (HOCl) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
3. Metal-coordinated fluorescent and luminescent probes for reactive nitrogen species (RNS) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.1. Fluorescent and luminescent probes for peroxynitrite (ONOO–) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.2. Fluorescent and luminescent probes for nitric oxide (NO) and nitrogen dioxide (NO2) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.3. Fluorescent and luminescent probes for nitroxyl (HNO) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4. Summary and outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
Declaration of Competing Interest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
Acknowledgement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

⇑ Corresponding author.
E-mail address: jyoon@ewha.ac.kr (J. Yoon).
1
These authors contributed equally to this work.

https://doi.org/10.1016/j.ccr.2020.213581
0010-8545/Ó 2020 Elsevier B.V. All rights reserved.
N. Kwon et al. Coordination Chemistry Reviews 427 (2021) 213581

1. Introduction Fluorescent and luminescent probes are an inevitable solution


to the issue of tracking a specific ROS or RNS to study its key mech-
Chemically reactive oxygen-containing species, or reactive oxy- anisms in living organisms. Detecting and imaging these ROS and
gen species (ROS), are produced from molecular oxygen (O2) dur- RNS with fluorescent and luminescent compounds has been a
ing vital processes occurring in humans and other living highly active area of research for the last couple of decades due
organisms [1]. Mitochondrial respiration is a major endogenous to the importance of ROS and RNS to human life [14–17]. To suc-
process for generating these species [2]. The NOX pathway is cessfully sense ROS/RNS, several criteria must be considered. It
another relevant endogenous process, as it generates ROS for should have high selectivity and sensitivity for specific ROS or
immune cell signaling when stimulated [3]. RNS to limit off-target effects. In order to detect targets in living
ROS can also be generated by exposure to infections, UV light cells, it is advantageous to display ’turn-on’ fluorescence, which
and other deleterious factors. ROS are essential to human life and is distinct from intracellular auto fluorescence. The reversibility
play key roles in various signaling and pathological processes, of the probes is also important, but most probes use a group of trig-
but overproduced ROS from exogenous stimuli can cause cell death gers that can react with specific ROS and RNS probes, so they are
by inducing oxidation of DNA, proteins or lipids [4]. Therefore, the irreversible. In this regard, probes containing redox metal ion can
amount of ROS can be a sign of various diseases, such as Alzhei- be a solution to solve this problem.
mer’s disease, Parkinson’s disease, inflammation, diabetes and can- Early probes could only detect ROS in solution, but cellular
cer [5]. Cellular redox mechanisms of various ROS and reactive imaging using confocal microscopy and imaging ROS and RNS in
nitrogen species (RNS) are presented in Fig. 1 [6]. live animals have become popular within the past decade [18]. Flu-
‘Highly reactive oxygen species’ were first reported in 1977 by orescent probes that target specific organelles and fluorescently
Weiss et al. [7]. ROS include hydrogen peroxide (H2O2), hypochlor- label ROS/RNS provide information on various diseases including
ous acid/hypochlorite (HOCl/ClO–), the hydroxyl radical (OH), the cancer and neurodegenerative disorders. The localization of the
superoxide anion radical (O– 1
2 ) and singlet oxygen ( O2). The roles probe provided information on oxidative stress specific to the cel-
of these ROS and RNS in biology will be briefly explained in the rel- lular organelles where the probe was accumulated [19]. Probes
evant sections. For example, O–2 is the precursor of most ROS and is using irreversible reactions with specific ROS/RNS are unable to
formed from O2 by electron reduction by NADPH oxidases or via distinguish between chronically elevated ROS levels as a patholog-
the electron transport chain [8]. Disproportionation of O–
2 can form ical feature and transient fluctuations in ROS production in physi-
H2O2, which can decompose to hydroxide and OH [9]. In addition, ological events [20]. A reversible probe was developed to monitor
myeloperoxidase (MPO) can convert H2O2 to HOCl in the presence the redox status in cells in real time [21]. In addition, the develop-
of Cl– [10]. ment of ratiometric probes prevented defects resulting from
Another important class of chemically reactive species is the set changes in fluorescence emission intensity due to differences in
of reactive nitrogen species (RNS), which can damage cells via sensor concentration [22]. Owing to the real-time monitoring
nitrosative stress. RNS include nitric oxide (NO), nitroxyl (HNO), enabled by this approach, these probes have been successfully
nitrogen dioxide (NO2) and peroxynitrite (ONOO–) [11]. For exam- used for imaging-guided surgery.
ple, NO can be endogenously generated by nitric oxide synthases, Many ROS- and RNS-selective fluorescent and luminescent
and failure of NO homeostasis can induce cancer, neurodegenera- probes adopt basic but important organic reactions. Therefore,
tive diseases and stroke [11]. ONOO–, which is produced from NO most of these probes are self-immolative systems [23] or so-
with O–2 , was considered an ROS only after a report by Beckman called chemodosimeters [24]. Probes called chemodosimeters react
in 1990 [12]. Due to its oxidizing and nitrating abilities, ONOO– with analytes and undergo significant chemical modifications,
also commonly oxidizes DNA, proteins and lipids, which can cause including the breaking and formation of covalent bonds. Because
various cancers, inflammation and neurodegenerative diseases these chemical modifications are irreversible, they cannot act as
[13]. Its protonated form (ONOOH, pKa = 6.8) decomposes to NO2 a reversible probe. For example, oxidations of arylboronic acids
and OH. HNO is both the protonated form of NO and the one- to phenols were the most popular organic reactions for sensing
electron reduced form of NO. H2O2 [25], HOCl [26] and ONOO– [27,28]. More recently, conver-
sion from N-heterocyclic carbene (NHC) boranes to imidazolium
salts has been used to detect HOCl [29,30].
Meanwhile, the metal coordination of the metal complex plays
O2 an essential role in the design of photoluminescent probes. Metal
GSSG O2
GS ions play key roles in the probes as an on–off redox switch for pho-
GSH toinduced quenching and as a reaction site with ROS, RNS or lumi-
OH GS
nescent cores. For example, H2O2 is cleaved reductively to OH by
O2
GSOO Cu+ or Fe2+ in the Fenton reaction, and the heme Fe2+–O2 adducts
Fe3+/Cu2+ can generate O– 2 in vivo [31]. Redox-active metal ions can be trig-
NO3 ger switches that control fluorescence quenching effects, which
ascorbate
OH N 2O 3 can be used to sense ROS or RNS. Further, metal ions, especially
Fe2+/Cu+ ONOOH
NO lanthanide metal ions, often can be the source of luminescence
NO2
NO
themselves [32]. They have been widely used as luminescent
HOCl Cl
probes because of their large Stokes, narrow emission bands, and
MPO H O ONOO long-lived luminescence, which can reduce the background signal
2 2
CO2 CO3 [33]. On the other hand, metal nanoparticles have enormous
NO2 NO2 O2 ONOOCO2
2e NO2 potential as sensors because of their unique properties such as high
e
electrical conductivity and surface reaction activity. Formation of
mitochondria, NADPH oxidases, redox cycling of xenobiotics metal nanostructure is one of the effective approaches for a biosen-
sor platform as it can improve biocompatibility and increase quan-
Fig. 1. Cellular redox mechanisms of various reactive oxygen species (ROS) and
tum yield [34]. For example, quantum dots, gold clusters, metal
reactive nitrogen species (RNS). Modified with Copyright 2018 Elsevier B.V.

2
N. Kwon et al. Coordination Chemistry Reviews 427 (2021) 213581

core-shall nanoparticles, etc. are actively investigated to provide boxylate shown in Fig. 2b, as confirmed by 1H NMR spectroscopy
improved fluorescence and increased photostability [35,36]. Espe- and ESI mass spectrometry. Probe 2 displayed excellent selectivity
cially, upconversion nanoparticles (UCNPs), a type of lanthanide- for H2O2 over various other ROS and metal ions, and the reported
doped luminophore, are particularly suitable for biosensing and limit of detection was 29 mM. Finally, probe 2 was successfully
bioimaging applications because they can be excited with NIR used to image lysosomal exogenous H2O2 in HeLa cells.
region light [37]. The metal–organic framework (MOF), on the Hitomo and coworkers reported an Fe3+ complex of a resorufin
other hand, has been extensively utilized as an effective method derivative (3) at pH 7.2 (20 mM HEPES buffer) [44]. Probe 3 was
to form various metal nanostructures in various applications such designed as a mimetic of nonheme Fe3+ peroxidase bearing a
as chemical sensing and drug delivery over the past few years due stable O-alkyl resorufin. The addition of H2O2 induced an enhance-
to its easy-to-adjust structural properties [38]. Metal complexes ment of fluorescence at 590 nm upon excitation at 530 nm, which
can have several advantageous photophysical properties compared was attributed to the release of resorufin after the reaction of the
to organic fluorophore. They are generally advantageous for iron complex with H2O2, as shown in Fig. 2c. Probe 3 was highly
bioimaging because they have a long emission lifetime, high lumi- selective for H2O2 over other ROS and RNS, such as ClO–, O– 2 and
nescence efficiency, and large Stokes shift. In addition, metal ions NO, and was used to image exogenous H2O2 in living HeLa cells.
can be easily quantified in the bio system through inductively cou- Furthermore, probe 3 successfully imaged epidermal growth factor
pled plasma mass spectrometry (ICP-MS) or X-ray fluorescence (EGF)-stimulated H2O2 generation.
(XRF). Das and coworkers reported the synthesis of carbon dots from
The probes discussed in this review are categorized by the tar- hemoglobin. These so-called blood dots (BD) contain heme groups;
get ROS or RNS, which are H2O2, HOCl/–OCl, OH, O– 1
2 , O2, NO, specifically, Fe2+ held in a porphyrin ring. The blue fluorescence of

ONOO , HNO and NO2. In this review, we cover ROS- and RNS- the BD is quenched by OH and O– 2 generated in situ from H2O2 by
selective fluorescent and luminescent probes designed from the BD, as shown in Fig. 3 [45]. BD can distinguish cancer cells
metal-coordinated systems over the last decade in accordance with (HeLa cells) from noncancer cells (CHO cells) because the H2O2
the theme topic of this special issue. We believe that both these concentration is higher in cancer cells, and H2O2 can be converted
previously reported probes and those that will be developed in to OH and O– 2+
2 by the Fe -enriched BD, resulting in fluorescence
the future will find important applications in clinical diagnosis, quenching. At the same time, the Fe2+-enriched BD activate H2O2,
prediction of drug efficacy, imaging-guided surgery and theranos- which acts as a prodrug producing ROS to cause oxidative damage
tics [16,39]. to DNA. BD can selectively kill cancer cells with 3-fold greater effi-
cacy compared to normal cells.
Chen, Lu and coworkers reported the Fe2+ complex 5,10,15,20-
2. Metal-coordinated fluorescent and luminescent probes for (4-sulphonatophenyl)porphyrin (TPPS4) (Fig. 4) as a turn-off fluo-
reactive oxygen species (ROS) rescent sensor for H2O2 at pH 7.4 (Tris-HCl buffer) [46]. Excitation
at 515 nm promoted fluorescence at 645 and 700 nm, which was
2.1. Fluorescent and luminescent probes for hydrogen peroxide (H2O2) turned off by the addition of H2O2. The oxidation of Fe2+ to Fe3+
by H2O2 was attributed to the formation of a TPPS4-Fe3+ complex;
Of the various ROS, H2O2 is one of the major species of interest, the electron transfer (ET) process resulted in fluorescence quench-
along with O– 
2 . It is a long-lived ROS, and it is a precursor for OH ing. The calculated detection limit was 1.24 nM, with a linear
and other ROS. A high concentration of H2O2 is a common feature response range of 0–42 lM.
of cancer cells [40]. Accordingly, many fluorescent probes have Shao and coworkers reported a ratiometric fluorescent probe
been designed to selectively sense H2O2 [41]. The oxidative inter- for H2O2 composed of silver nanoparticles and denatured lyso-
action between iron and H2O2 is used for fluorescence detection zymes (dLys-AgNCs), which is shown in Fig. 5 [47]. In the presence
of H2O2. To design H2O2 detection probes, a variety of probes that of Fe2+, H2O2 is reduced to OH via the Fenton reaction. This can
mimic the bioredox process between the iron porphyrin center of cause oxidation of tyrosine in the lysozyme, resulting in enhance-
the heme enzyme and H2O2 have been developed. In addition, ment of the dLys-AgNCs emission at 450 nm, along with quenching
H2O2 chelation by Eu3+-tetracycline complex, photooxidation of of the original emission of dLys-AgNCs at 640 nm by both OH and
thioether and H2O2 induced cleavage of boronate has been used Fe3+. Lysozyme acts as a stabilizing ligand for the AgNCs as well as
in the design of H2O2 detection probes. a blue fluorescent source. The calculated detection limit was
Zhang et al. reported that probe 1 can selectively image H2O2 in 0.2 mM, with a linear response range of 0.8–200 mM. In the presence
lysosomes when MPO is present (Fig. 2a), which means that probe of O2, GOx catalyzes glucose to generate H2O2. Accordingly, the
1 can image H2O2 in the relatively acidic pH range of 4.5–6 [42]. In addition of glucose, Fe2+ and GOx induced ratiometric fluorescence
the absence of MPO, probe 1 can selectively react with ClO– over changes in dLys-AgNCs, which means dLys-AgNCs can be used to
various other ROS, such as OH, ROO, t-BuOOH, O– –
2 , ONOO and sense glucose. Further, dLys-AgNCs were reported to have imaged
H2O2. However, the addition of H2O2 and Cl– in the presence of endogenous OH in living cells.
MPO resulted in a large increase in fluorescence at 610 nm upon Li et al. reported neutral iridium(III) complex 4, which bears a
excitation at 380 nm. Thus, probe 1 with MPO showed high selec- benzeneboronic acid pinacol ester (bpe) as a reactive moiety to
tivity for H2O2 over other ROS, with a reported detection limit of detect H2O2, as a phosphorescent probe that is selective for H2O2
42 nM. Probe 1 was localized in lysosomes in the cell imaging data over other ROS [48]. As shown in Fig. 6a, 3-(benzothiazol-2-yl)-7-
with MPO, which was confirmed by LysoTracker Green with a Pear- hydroxy-coumarin bearing a bpe moiety was a cyclometalated
son’s correlation coefficient of 0.92. These results indicate that ligand of 4-I, which absorbs at 460 nm and phosphoresces at
probe 1 can successfully image lysosomal H2O2 and lysosomal 560 nm. The reaction with H2O2 induced significant intermolecular
H2O2 flux. aggregation owing to the removal of the steric hindrance caused by
The iron-complexed fluorescein derivative 2 was reported as a the bulky pinacol ester moiety, resulting in decreased phosphores-
H2O2-selective fluorescent probe at pH 7.0 (25 mM PIPES buffer) cence. A boronic acid or ester deprotection strategy was nicely
[43]. The fluorescence of 2 is quenched when the iron center is adopted to sense H2O2 in this work.
paramagnetic, and the addition of H2O2 induced an enhancement A Eu3+-coordinated polymer containing a sulfur tag was
of fluorescence at 528 nm. Activation of the iron center by H2O2 reported as a H2O2-selective luminescent probe that had been syn-
followed by oxidative dealkylation affords the fluorescein biscar- thesized from Eu(NO3)36H2O and 2,20 -thiodiacetic acid [49]. A
3
N. Kwon et al. Coordination Chemistry Reviews 427 (2021) 213581

(a)
NC CN NC CN

N N N N
Zn MPO
Zn
O O O O
H2O2, Cl-
N S S N N S O O S N

(b) H 2O H2 O
N N H2O
H2O
Fe Fe
N N N N CHOO- COO-
HO O O
O O O H2O2

Cl Cl
Cl Cl
-
CO2-
CO2

2
(c)
N +
O
O 2+
H
O N HO O O
H2O2 FeIII N
H O H
N N N
N O resorufin
HO FeIII N
N N
O

Fig. 2. Reaction mechanism of (a) probe 1 with H2O2 in the presence of MPO and that of (b) probe 2 and (c) 3 with H2O2.

Fig. 3. Hemoglobin-based blood dot (BD) and its reaction mechanism with H2O2. Copyright 2018 American Chemical Society.

SO3H
conversion from a microcrystal structure to a nanoparticle struc-
ture with retention of fluorescence occurs upon increasing the
temperature. The peaks in its emission spectra were attributed to
the 5D0 ? 7F04 transitions of Eu3+. The addition of H2O2 oxidizes
the thioether moiety to sulfoxide, as shown in Fig. 6b, which leads
to quenching of the Eu3+ emission. A linear correlation range of 5–
N HN
SO3H
150 lM was reported. The microcrystals could detect ROS in aero-
HO3S
N
sols, whereas the nanoparticles could image exogenous H2O2 in
NH
human embryonic kidney 293 T (HEK 293 T) cells.
Xu and Kuang et al. reported chiral upconversion nanoassem-
blies (UCNP@ZIF-NiSx) that detect ROS, especially H2O2 in living
cells [50]. As shown in Fig. 7, upconversion nanoparticles (UCNPs)
were first stabilized by oleic acid (OA) to form hexagonal-plate-like
SO3H
UCNPs-OA. After a reaction with HCl to remove the OA, the UCNPs
Fig. 4. The structure of 5,10,15,20-(4-sulphonatophenyl)porphyrin (TPPS4). were decorated with polyvinylpyrrolidone (PVP) to form UCNP-

4
N. Kwon et al. Coordination Chemistry Reviews 427 (2021) 213581

thesized from 2-boronobenzene-1,4-dicarboxylic acid and ZrOCl2-


8H2O in the presence of formic acid (HCOOH), as shown in Fig. 8
[51]. The Zr-MOF contains the UiO-66 structure, and 2-
boronobenzene-1,4-dicarboxylic acid acts as a linker. Zr-UiO-66-
B(OH)2 displayed a fluorescence enhancement at 428 nm that
was selective for H2O2 over other ROS at pH 7.4 (HEPES buffer).
The addition of H2O2 induced the oxidative breakage of the  B
(OH)2 moiety into an  OH group. The reported detection limit
was 0.015 lM, and the probe was used to image exogenous H2O2
in MDAMB-231 cells.

2.2. Fluorescent and luminescent probes for hydroxyl radicals (OH)

Owing to the physiological significance of OH, fluorescent


Fig. 5. Ratiometric fluorescent sensing of H2O2 and OH using lysozyme–silver
nanoclusters (dLys-AgNCs). Copyright 2016 American Chemical Society. probes to selectively detect and image OH have been attractive
research targets. However, of the various ROS and RNS, OH is par-
PVP, which was further coated with zeolitic imidazolate framework- ticularly challenging to detect due to its high reactivity and extre-
8 (ZIF) to form UCNP@ZIF. Finally, further reactions with nickel(II) mely short half-life of 109 s [52]. To designing the OH probe,
ions and chiral L-/D-penicillamine (Pen) ligands as a sulfide source aromatic hydroxylation, oxidative dearylation, and upconversion
afforded chiral UCNP@ZIF-NiSx. UCNP@ZIF-NiSx displayed strong NPs (UCNPs) are frequently-used mechanisms. First, with aromatic
circular dichroism (CD) signals at 440 and 530 nm, and upconversion hydroxylation, OH can easily attack the aromatic rings and make it
luminescence (UCL) signals were observed at 540 nm and 660 nm. hydroxylated. When there are lanthanide ions, the hydroxylated
So, dual-mode detection of CD and fluorescent signals could be used moiety can be a more efficient sensitizer, which usually give dis-
to detect ROSH2O2 was chosen as a representative ROS. Upon the tinct optical changes [53]. Next, electron rich units like aniline or
addition of H2O2, the UCL signal of the UCNPs at 540 nm decreased, phenol groups can mask the hydroxyl or amine in probes, resulting
while the UCL signal at 660 nm remained the same. Based on the in quenching through PET. Through OH-triggered oxidative deary-
UCL intensities (I660/I540 ratio), the reported detection limit was lation, the fluorescence get recovered. Finally, in UCNPs, lanthanide
0.037 lM, with a linear response range of 0.05–20 lM. Using normal metal ions such as Yb3+ are doped in an appropriate host lattice.
primary uterine fibroblast cells (PCS-460–010), the authors success- Luminescence resonance energy transfer (LRET) is common feature
fully demonstrated that these chiral nanoassemblies can quantita- of UCNPs-based fluorescent probes from an upconversion donor to
tively monitor exogenous H2O2 when excited with a 980 nm laser. an energy acceptor [54].
The UCL in the green region increased, while the UCL in the red region Time-gated (or time-resolved) luminescence using lanthanide-
was almost the same. Based on the UCL signal, the calculated detec- based compounds is an active area of investigation for luminescent
tion limit was 0.13 lΜ, with a linear response range of 0.18–10.2 lΜ. probes [55]. Pierre and coworkers reported ternary terbium (Tb)
Biswas and coworkers reported a Zr(IV) metal–organic frame- complex 5 (Fig. 9a) as a selective probe for OH via time-gated
work (Zr-MOF) bearing a boronic acid moiety (Zr-UiO-66-B(OH)2) luminescence [53]. Probe 5 can detect OH at concentrations as
as a H2O2-selective fluorescent probe. Zr-UiO-66-B(OH)2 was syn- low as 0.75 fM. Hydroxylation of trimesate to hydroxytrimesate
was followed by coordination to form Tb complex 5 (Fig. 9a), which

(a)
O
B OH
O
+
S N S N
O O
O
Ir Ir
O O
O O
O O
O
H2O 2
O B
O OH 2
2

4-I
4

(b) UV Emission Quenched


UV

O O
+

S
HO OH O R
H2 O 2
+ Eu S
R'
Eu(NO3)3¡¤6H2O

Fig. 6. Proposed mechanism of (a) luminescent probe 4 and (b) a Eu(III)-coordinated polymer with H2O2.

5
N. Kwon et al. Coordination Chemistry Reviews 427 (2021) 213581

Fig. 7. Proposed scheme for UCNP@MOF-NiSx system to detect ROS. Copyright 2019 American Chemical Society.

Tian and coworkers reported AuNC@HPF, which is a gold nan-


ocluster (AuNC) decorated with bovine serum albumin (BSA) and
2-[6-(40 -hydroxy)phenoxy-3H-xanthen-3-on-9-yl]benzoic acid
(HPF), as a ratiometric fluorescent sensor for OH (Fig. 10) [57].
As shown in Fig. 10, only the emission at 632 nm from AuNCs
was observed upon excitation at 488 nm because HPF is weakly
fluorescent. The addition of OH induced additional emission at
515 nm from the reaction product, fluorescein, which results in
ratiometric fluorescence changes. The reported detection limit
was 0.68 lM, with a linear response range of 1  106 to
1.5  104 M. AuNC@HPF was then used to image endogenous

OH in HeLa cells stimulated by lipopolysaccharides (LPS).
AgNCs were stabilized by cytosine-rich oligonucleotides to
detect OH from among various ROS and RNS by Tian and cowork-
ers [58]. Rolling circle amplification (RCA) was employed to pre-
pare a DNA hydrogel, which is cross-linked with cytosine-rich
sequences in the presence of Ag+ via cytosine–Ag+–cytosine
(Fig. 11). After the reduction of Ag+, the resulting RCA-stabilized
Fig. 8. The synthetic route for Zr-UiO-66-B(OH)2 and its fluorescence enhancement
AgNCs (RCA-AgNCs) displayed improved photostability. The green
at pH 7.4 (10 mM HEPES). Copyright 2018 American Chemical Society.
channel detected emissions between 500 and 560 nm upon excita-
tion at 448 nm, and the red channel detected emissions between
610 and 700 nm upon excitation at 552 nm. The addition of OH
induced fluorescence at 545 nm. This approach was highly selec- induced a decrease in red fluorescence and an increase in green flu-
tive for OH over other ROS and RNS, such as O 1  
2 , O2, t-BuO , ClO , orescence. The reported detection limit was 58 nM, with a linear
H2O2 and NO. With the aid of a long time-gate (0.2 ms), selective response range between 0.2 and 80 lM. RCA-AgNCs were subse-
luminescence of Tb was observed without interference from back- quently used to dynamically and ratiometrically monitor endoge-
ground luminescence. nous ROS/RNS in LPS-stimulated A549 cells.
The self-immolative, luminescent terbium(III) complex 6 was AgNCs were reacted with a DNA template to form DNA-AgNCs
reported as a probe for the selective detection of OH (Fig. 9b) (Fig. 12) and were reported as a OH-selective fluorescent probe
[56]. The probe had an emission maximum at 514 nm upon excita- by Lei and coworkers [59]. The DNA-AgNCs showed a strong green
tion at 315 nm. The electron-rich p-nitrophenoxy moiety can effi- emission at 550 nm upon excitation at 440 nm. The addition of OH
ciently quench the luminescence by PET. 49-fold enhancement of cleaved the DNA template and induced aggregation of the AgNCs,
luminescence was observed upon the addition of OH at pH 7.0 resulting in fluorescence quenching. A linear response to OH was
(0.05 M HEPES), which was attributed to the elimination of the observed in the range of 50 nM to 5.0 mM, and the DNA-AgNCs
p-aminophenyl group, as shown in Fig. 9b. The reported detection were successfully used to image endogenous OH in SK-N-SH cells.
limit was 270 nM, with a long time-resolved lifetime (2.76 ms). Tb The corresponding OH concentration in cell extracts stimulated by
complex 6 was successfully used to detect exogenous OH in HeLa PMA was 0.35 lM, which was very close to the value obtained by
cells with background-free luminescence. HPLC (0.31 lM).

6
N. Kwon et al. Coordination Chemistry Reviews 427 (2021) 213581

(a) O
O
OH2
N HN O O
OH2 O O
O Tb
O O N N O
O HO O O N HN O
O O
O OH O O
Tb
O
N N O
O O O
O
O O
O O O 5
O

(b)
O
CH2OH

NH2

N N OH N N
N N N N N N

N N NH - N
CO2- -O2C N O 2C
CO2-
- - -
CO2 O 2C O2C
Tb3+ O CO2- Tb3+
6

Fig. 9. Reaction mechanism of (a) a terbium-based luminescent probe (5) via hydroxylation of trimesate and (b) Tb-based luminescent probe 6 with OH.

Fig. 10. Strategy for the preparation of AuNC@HPF and its reaction scheme with OH. Copyright 2014 American Chemical Society.

7
N. Kwon et al. Coordination Chemistry Reviews 427 (2021) 213581

Zeolitic imidazolate framework-8 (ZIF-8) was employed as a


host for fluorescent sources to afford EuPS@AnC/ZIF-8, a ratiomet-
ric fluorescent sensor for 1O2, as illustrated in Fig. 15 [66]. 9-
Anthracenecarboxylic acid (AnC), 2-methylimidazole and zinc
nitrate were treated with a polystyrene core decorated with a
europium chelate (EuPS) to afford EuPS@AnC/ZIF-8. In this ratio-
metric system, the emission at 615 nm from Eu3+ acts as a refer-
ence signal and the emission at 410 nm from AnC displays
distinct fluorescence quenching after the reaction with 1O2. In
addition, a colorimetric change from blue-violet to red was
observed after the reaction with 1O2. The reported detection limit
was 43 nM, with a linear response range of 0.5–200 lM.
Wang and coworkers reported a dinuclear Ru2+ complex con-
taining an anthryl moiety (8) as a 1O2-selective luminescent probe
(Fig. 16) [67]. Probe 8 reacts with 1O2 to form its endoperoxide at
pH 7.0 (50 mM Tris-HCl buffer) and pH 10.5 (0.1 M carbonate buf-
Fig. 11. Scheme for the synthesis of RCA-AgNCs. Copyright 2018 American fer), resulting in a large enhancement of emission at 609 nm upon
Chemical Society. excitation at 458 nm. The reported detection limits were 3.1 nM at
pH 7 and 2.7 nM at pH 10.5.
Voelcker and coworkers reported MC-EuA, which was synthe-
sized from a porous silicon microcavity (pSiMC) with a
Liu and coworkers reported a luminescence resonance energy
europium-anthracene (EuA) moiety as a luminescent probe for
transfer (LRET)-based upconversion nanoprobe as a OH-selective 1
O2 [68]. On a porous silicon thin film, anthracenes linked by Si–
sensor [60]. The upconversion nanoprobe was prepared by intro-
O bonds react with 1O2 to form the corresponding endoperoxide
ducing an azo dye, which can be oxidized by OH, onto the surface
adducts. Fig. 17 shows that the MC-EuA emission at 615 nm was
of the NaYF4@NaYF4:Yb,Tm@NaYF4 nanoprobe via interactions
quenched with the addition of ClO– and was dramatically
between the carboxylic acid moieties and lanthanide ions. The
enhanced upon the addition of 1O2. The reported detection limit
nanoprobe emits at 487 nm upon excitation at 980 nm. When
 was 37 nM, with a linear response range of 50 nM to 1 M.
OH was added, the luminescence quenching due to LRET from
the upconversion nanoparticles to the azo dyes (Fig. 13) was
revived owing to the oxidative decomposition of the azo dye with 2.4. Fluorescent and luminescent probes for superoxide anion radical

OH. An excellent detection limit of 1.2 fM was reported, with a lin- (O–
2)
ear range of 1.2–194.6 fM. The probe was used to detect endoge-
nous OH in HeLa cells and mouse tissues stimulated by LPS, as The primary ROS is O–2 , which is the one-electron reduced form
shown in Fig. 13. of O2 and a precursor of several other ROS and RNS. For example,
O– –
2 can generate ONOO by a nonenzymatic reaction with NO. In
2.3. Fluorescent and luminescent probes for singlet oxygen (1O2) addition, O–
2 plays key roles in both physiological and pathological
processes [69].
1
O2 is cytotoxic due to its oxidation of proteins, DNA and lipids. Tang and coworkers reported DBZTC/Ag@SiO2 as a nanoprobe
1
O2 is generated via enzymatic and nonenzymatic reactions, but for O– 2 [70]. The 2-chloro-1,3-dibenzothiazoline-cyclohexene
photosensitizers in photodynamic therapy also may generate 1O2 (DBZTC) group on the surface of Ag@SiO2 is oxidized in the pres-
to kill cancer cells [61,62]. ence of O–
2 (Fig. 18), which can induce fluorescence enhancement
(2 + 4) cycloaddition of 1O2 to anthracene to form an endoper- at 520 nm upon excitation at 485 nm at pH 7.4 (20 mM HEPES).
oxide was used as a delivery system for 1O2 to induce antitumor The calculated detection limit was 0.73 nM, with a linear response
efficacy [63,64]. The opposite path of receiving 1O2 instead of deliv- range of 0.02–3.33 mM. DBZTC/Ag@SiO2 was successfully used to
ering it has been a popular strategy for designing a selective lumi- image endogenous O– 2 upon stimulation by PMA in RAW264.7
nescent probe for 1O2. Endoperoxide adducts are generated from cells.
europium complex 7, as shown in Fig. 14 [65]. Upon the addition Li and Xiao reported gold nanodots (Au NDs) as a selective flu-
of 1O2 to form the endoperoxide product, the fluorescence quench- orescent probe for O–
2 in which 2-mercaptobenzothiazole was used
ing due to PET was eliminated, resulting in strong fluorescence at as a stabilizing agent and D-(+)-mannose was used as a reducing
794 nm at pH 7.4 (0.05 M Tris-HCl buffer). The calculated detection agent via a one-pot sonochemical method, as illustrated in
limit was 3.7 nM. When probe 7 was treated with PDT drugs, such Fig. 19 [71]. The Au NDs displayed a strong orange emission at
as hematoporphyrin monomethyl ether (HMME), in HeLa cells, the 532 nm upon excitation at 346 nm, which was efficiently
generation of 1O2 was successfully monitored by time-gated quenched by the addition of O– 2 . The reported detection limit
luminescence. was 445 nM, with a linear response range of 0.6–78 lM. The addi-

AgNO3 OH
NaBH4 DNA-AgNCs
Fluorescence quenching

: DNA : Ag nanocluster

Fig. 12. Synthesis of DNA-AgNCs and the reaction of DNA-AgNCs with OH.

8
N. Kwon et al. Coordination Chemistry Reviews 427 (2021) 213581

COOH

N COOH
N
= N = oxidized azo dye
OH

Azo dye

Fig. 13. Luminescence resonance energy transfer (LRET)-based upconversion nanoprobe, in which oxidative decomposition of an azo dye with OH induced the recovery of
luminescence, and its application to the imaging of endogenous OH in HeLa cells and tissues. Modified with permission from Copyright 2015 American Chemical Society.

O
O

PET 1 PET
N
O2
N +
N N N N
Eu3+ Eu3+
CO2 O 2C CO2 CO O 2C
CO2 N 2
N
N N N
N
7

Fig. 14. Reaction mechanism of probe 7 containing an anthracene moiety with 1O2.

Fig. 15. Preparation of EuPS@AnC/ZIF-8 and its reaction mechanism with 1O2. Copyright 2017 Royal Society of Chemistry.

tion of vitamin C revived the emission at 532 nm. The Au NDs were In the same paper, the authors also reported the application of this
then used to image exogenous O– 2 in HepG2 cells. system as a MOF-based catalyst for Knoevenagel condensation.
Zr-UiO-66-NH-CH2-Py was reported as a new MOF-based fluo-
rescent sensor for O–2 that had been synthesized from Zr
4+
ions 2.5. Fluorescent and luminescent probes for hypochlorous acid (HOCl)
and 2-[(pyridin-4-ylmethyl)amino]terephthalic acid (Fig. 20, left)
[72]. There was a distinct enhancement of fluorescence at HOCl (pKa = 7.6), which often exists as ClO–, is a strong oxidant
430 nm upon excitation at 330 nm when O– 2 was added to the and is produced from H2O2 in the presence of chloride and
aqueous solution of Zr-UiO-66-NH-CH2-Py (Fig. 20). The reported myeloperoxidase (MPO) [73]. HOCl plays an important role in innate
detection limit was 0.21 lM with a fast response time of 240 s. immune systems to destroy invasive bacteria, but excessive amounts
9
N. Kwon et al. Coordination Chemistry Reviews 427 (2021) 213581

4+ occurred at 587 nm with an emission lifetime of 155 ns, which


was attributed to 3[p(ppy) ? p*(L1/L2)] 3LLCT and 3[5d(Ir) ? p*
N N (L1/L2)] 3MLCT. As shown in Fig. 21a, ClO converts probe 9 to
N H H N
N N N N its carboxylic acid product. In addition, luminescent probe 9
Ru Ru showed excellent selectivity towards ClO over other ROS and
N N N N N N
metal ions.
N N
Yuan and coworkers reported ruthenium(II) complex 10 for
HOCl detection at pH 7.4 (50 mM borate buffer) [80]. Very weak
8 emission was observed for Ru(II) complex 10 due to PET, but after
1
O2 it reacted with HOCl, strong luminescence was observed at 612 nm
upon excitation at 456 nm The detailed proposed mechanism is
4+ presented in Fig. 21b. The addition of HOCl induced the formation
of the unstable intermediate of benzoxazolone N-oxide 10 I, which
N N was converted to the strongly emissive product 10-II. The reported
N H H N
N N N N detection limit was 33 nM, with a linear response range of 2.5–
O
Ru
N
O
N
Ru 40 lM. Probe 10 was successfully used to image exogenous and
N N N N
endogenous HOCl in HeLa cells and porcine neutrophils.
N N
The Ru(II) complex 11 was reported as a HOCl-selective lumi-
nescent probe (Fig. 21c) in which the 2,4-dinitrophenyl (DNP)
group acts as an electron acceptor and fluorescence quencher
Fig. 16. The reaction mechanism of a dinuclear Ru2+ complex containing an anthryl [81]. Weak luminescence was observed for probe 11 due to the
moiety (8) with 1O2. DNP moiety, but HOCl selectively ‘‘turned on” the probe
by  190-fold at 626 nm upon excitation at 456 nm at pH 7.4
(0.1 M phosphate buffer). The addition of HOCl cleaves the DNP
of HOCl cause rheumatoid arthritis, atherosclerosis, cardiovascular group, followed by oxidation to a carboxylic acid moiety
diseases and other chronic diseases [74]. Accordingly, a great many (Fig. 21c). The reported detection limit was 53.5 nM, with a linear
probes have been developed to selectively sense HOCl among vari- response range of 2.5–50 mM. Ru(II) complex 11 could image
ous ROS and RNS [75,76]. For detecting HOCl/OCl, the well- endogenous HOCl in RAW264.7 cells.
known recognition sites are reactive groups like rich electron groups An iridium(III) complex (12) was reported as a ClO–-selective
such as S, Se, double bond, hydroximic acid, and hydrazine. Partially- luminescent probe at pH 7.4 (10 mM PBS buffer) [82]. Probe 12
positive-charged Cl atom in HOCl can easily react with recognition exhibited strong fluorescence at 615 nm upon excitation at
group containing partial negative charge to produce chloride. After 370 nm, which was quenched by the addition of ClO. The fluores-
that, the chloride sometimes undergoes b-elimination or hydrolysis cence of probe 12 was selectively quenched by ClO– with a rapid
to cause changes of fluorescence [77]. Recently, DUOX-dependent response time (<30 s) and a reported detection limit of 0.41 lM.
HOCl production was visualized in the intestine following gut infec- As shown in Fig. 21d, the methacrylate group is converted to its
tion using a rhodamine-based fluorescent probe, which demon- phenol product via oxidation by ClO–, which was confirmed by
strates the role of specific ROS- or RNS-selective fluorescent probes NMR and mass spectroscopy. Probe 12 was subsequently used to
[78]. image ClO– in HepG2 cells.
Chen and coworkers presented iridium complex 9 as a ClO- b-Diketonate–europium complexes 13-I and 13-II were
selective luminescent probe. Iridium complex 9 displayed very reported as mitochondria- and lysosome-selective HOCl probes,
weak emission at pH 7.2 (DMF–HEPES 4:1 v/v), which was attrib- respectively [83]. Both probes displayed strong red emission at
uted to isomerization of the C = N–OH bond; strong luminescence 607 nm upon excitation at 330 nm at pH 7.4 (PBS buffer). The addi-
was observed at 77 K in frozen DMF due to inhibition of this rota- tion of HOCl induced efficient fluorescence quenching (Fig. 21e).
tion (Fig. 21a) [79]. Upon the addition of ClO, strong emission The calculated detection limit was less than 15 nM. Due to the

Fig. 17. (a) Emission spectra of MC-EuA with ClO– and MC-EuA+1O2 (solid red curve) upon excitation at 348 nm. (b) Proposed scheme of MC-EuA interacting with 1O2.
Reproduced with permission. Copyright 2017 American Chemical Society.

10
N. Kwon et al. Coordination Chemistry Reviews 427 (2021) 213581

Ag@SiO2 DBZTC/Ag@SiO2 time-gated luminescence microscopy was successfully used to


image the uptake of HOCl by Daphnia magna using these probes.
Probe 14 was reported as a time-gated luminescence (TGL)
DBZTC O2
Ag Ag Ag probe for HOCl in lysosomes, in which luminescent Tb3+ is a fluo-
rescence resonance energy transfer (FRET) donor and a rhodamine
moiety bearing a HOCl trigger group is a FRET acceptor (Fig. 22a)
[84]. The addition of HOCl induced a decrease of the Tb3+ emission
at 540 nm and an increase of the rhodamine emission at 580 nm at
Oxidized DBZTC
DBZTC H pH 7.4 (50 mM PBS). In addition, the lifetime of the Tb3+ emission
N S N S
= = decreased from 588 ls to 254 ls, along with the FRET process. The
S Cl HN S Cl N reported HOCl detection limit was 0.34 lM, with a linear response
range of 0.5–45 lM. Ultimately, probe 14 was successfully used to
Fig. 18. Proposed mechanism of DBZTC/Ag@SiO2 with O
2 .
image lysosomal endogenous HOCl in HepG2 cells, which was con-
firmed with LysoSensor Green.
Pt pincer complexes containing 4,5-bis(diphenylphosphino)acri
dine and either benzenethiolate (15) or 4-thiolatobenzoic acid (16)
were reported as HOCl-selective fluorescent probes [85]. Probe 15
displayed selective fluorescence enhancements with HOCl and
ONOO–. Probe 16, which has the less electron-rich 4-
thiolatobenzoic acid, shows a highly selective fluorescence
enhancement at  490 nm only with HOCl. The oxidation mecha-
nisms of probes 15 and 16 with HOCl are shown in Fig. 22b. The
reported HOCl detection limit of 16 was 33 nM and the reported
HOCl and ONOO– detection limits of 15 were 26 and 33 nM,
respectively.
The nanoscale metal–organic framework (NMOF) UiO-68-ol
was reported as a HOCl-selective fluorescent sensor [86]. The p-
methoxyphenol ligand was first studied as a redox-switchable
ligand, in which the conversion from p-methoxyphenol to quinone
followed by the formation of hydroquinone was confirmed. The p-
methoxyphenol ligand then reacted with ZrCl4 to form the Zr-MOF
UiO-68-ol. A colorimetric change from colorless to yellow was
observed selectively with HOCl over other ROS and RNS with the
naked eye, as shown in Fig. 23. In addition, the original fluores-
cence at 475 nm upon excitation at 405 nm was efficiently
quenched, with a detection limit of 0.1 lM in PBS buffer. UiO-68-
Fig. 19. Synthesis of gold nanodots (Au NDs) and reaction mechanism with O–
2.
ol was subsequently used to image both exogenous and endoge-
Copyright 2017 Royal Society of Chemistry. nous HOCl production in RAW264.7 cells.
The metal–organic framework (MOF) {[Eu2Cu(IN)5(CO3)(H2O)]},
which was prepared from CuI and the lanthanide duplex chain
triphenylphosphonium group in 13-I and the morpholine moiety shown in Fig. 24, was reported as a ClO-selective sensor [87].
in 13-II, 13-I could be selectively located in mitochondria and The MOF complex fluoresces at 486 nm, which is a characteristic
could image exogenous and endogenous HOCl in the organelle, emission of CuI. The MOF sensor showed selective fluorescence
whereas 13-II could selectively image HOCl in lysosomes. Notably, quenching upon the addition of ClO , with a distinct colorimetric

Fig. 20. Zr-UiO-66-NH-CH2-Py as a fluorescent sensor for O–


2 and Zr-UiO-66-NH-CH2-Py as a MOF-based catalyst for Knoevenagel condensation (left) and fluorescent changes
of Zr-UiO-66-NH-CH2-Py at 430 nm with various amounts of O– 2 upon excitation at 330 nm (right). Inset: Photos of cuvettes containing Zr-UiO-66-NH-CH2-Py before and
–
after the addition of O2 under a UV–Vis lamp. Copyright 2019 Royal Society of Chemistry.

11
N. Kwon et al. Coordination Chemistry Reviews 427 (2021) 213581

(a) PF6-
OH
N COOH
N N
N N
ClO-
Ir Ir
N N
N N

(b) NH2 N
+ O
O N OH
O NO2
O-
N N N
N N N
N N N
HOCl HOCl
Ru II RuII RuII
N N N N N
N
N N N

10 10-I 10-II

(c) NO2 NO2


T
PE
S
COOH
N N
N N HOCl N N
Ru2+ Ru2+
N N N N
N N

11

(d)
N N
N N N N
ClO-
Ir Ir
N N
N N

O OH
2
O 2 12 non-fluorescent
fluorescent

(e) R
R

HOCl

O O
Eu3+
O O HO O O OH
C3F7 1/2 C3F7
Non-fluorescent
Fluorescent at 607 nm
Br
13-I: SO2-NH(CH2)6P(Ph)3 13-II: SO2-NH(CH2)6N O

Fig. 21. Proposed reaction mechanism of (a) iridium complex 9 with ClO, (b) ruthenium(II) complex 10, (c) complex 11, (d) iridium complex 12 and (e) europium complex
probes with HOCl.

change from yellow to blue that is readily observed with the naked to the naked eye, which were attributed to a reversible redox pro-
eye. The calculated detection limit was 10 lM. cess with vitamin C (VC). The calculated detection limit was
Dong and coworkers reported the MOF-based UiO-68-PT as a 0.28 lM, with a linear response range of 0–80 lM. UiO-68-PT
selective fluorescent probe for HOCl. UiO-68-PT was synthesized and poly(vinyl alcohol) could form mixed membrane material
from ZrCl4 and a phenothiazine–benzimidazole–dicarboxylate (MMM), which appeared as a deep red color with an absorption
bridge as a ligand (Fig. 25) [88]. The addition of HOCl induced a band at 533 nm upon the addition of HOCl, as shown in Fig. 25.
large fluorescence enhancement at 392 nm upon excitation at A luminescent metal–organic framework (AF@MOF-801) was
323 nm and a highly selective colorimetric change that is distinct reported to selectively detect HOCl by a size-selective effect in
12
N. Kwon et al. Coordination Chemistry Reviews 427 (2021) 213581

NO2
(a)
O 2N
O O O
HN CO2H
N N
N N
N
N N

O O

N N HOCl N N
3+ N pH 7.4 N
Tb Tb3+
CO2 CO2 CO2 CO2 CO2 CO2 CO2 CO2
14 14-I FRET

(b)

PPh2 PPh2

N Pt S HOCl or ONOO-
N Pt S O
O
PHPh2 PPh2
ClO4- ClO4-
15 15-I

CO2H CO2H
PPh2 PPh2

N Pt S HOCl
N Pt S O
O
PPh2 PPh2
ClO4- ClO4-
16 16-I

Fig. 22. The reaction mechanisms of (a) luminescent Tb3+ complex 14 with HOCl and (b) platinum–acridine-based probes 15 and 16 with ONOO–.

Fig. 23. The Proposed structure of UiO-68-ol and its color changes with various ROS and RNS, including HOCl, and confocal fluorescence images of RAW264.7 cells with HOCl.
Copyright 2017 American Chemical Society.

which large interferents can be effectively inhibited from entering Tan and coworkers reported the MOF CD/CCM@ZIF-8 as a FRET
the cage [89]. The authors described a ‘‘ship in a bottle” strategy to sensor for HOCl that exhibits ratiometric fluorescence [90]. As
prepare the luminescent metal–organic framework MOF-801 using shown in Fig. 27, ZIF-8 was selected as a frame for the MOF, which
Zr6 clusters as inorganic nodes and fumaric acid as organic ligands can accommodate curcumin (CCM) and carbon dots (CD) during
(Fig. 26). The luminescent marker 5-aminofluorescein (AF) was the assembly. The addition of HOCl induced an increase of the
encapsulated during the formation of the MOF to afford CD emission at 410 nm and a decrease of the CCM fluorescence
AF@MOF-801. The addition of HOCl induced about 80% quenching at 585 nm. The reported detection limit was 67 nM, with a linear
of AF@MOF-801 emission at pH 7.4 (60 mM HEPES). The size selec- response range of 0.1–50 lM. CD/CCM@ZIF-8 was then used to
tivity for ClO– was attributed to the small pores of the entrance. detect HOCl in serum. The serum was treated with N-ethyl malei-
The calculated detection limit was 0.052 lM, with a linear mide to remove biothiols, such as GSH, before the HOCl was added.
response range of 0–7 lM. AF@MOF-801 was subsequently used The amount of HOCl spiked in the serum was determined from the
to monitor intracellular exogenous ClO in SMMC-7721 and HeLa ratiometric fluorescence of CD/CCM@ZIF-8, which was in good
cells. agreement with the results determined by kit assay.

13
N. Kwon et al. Coordination Chemistry Reviews 427 (2021) 213581

PVP-capped copper nanoclusters (CuNCs) prepared from Cu2+


with PVP as the scaffold material and ascorbic acid (AA) as the
reducing agent showed sensitive, HOCl-selective fluorescence
quenching (Fig. 28) [91]. PVP-capped CuNCs exhibited strong emis-
sion at 416 nm upon excitation at 340 nm, which was selectively
quenched with HOCl. The fluorescence quenching was attributed
to the oxidation of copper on the surface of the nanocluster by
HOCl. The reported detection limit was 55 nM, which decreased
to 19 nM in the presence of excess iodide (I–). HOCl oxidizes iodide
to iodine (I2), which can react with copper on the surface of the
nanocluster to afford cuprous iodide (CuI) as a precipitate.
As shown in Fig. 29, GSH-decorated AuNCs (GSH-AuNCs) were
encapsulated in NMOFs made of 2-methylimidazole (2-HIm) and
Zn2+ to form AuNCs@NMOFs [92]. The fluorescence of the
AuNCs@NMOFs was about 10 times stronger than that of the
GSH-AuNCs, and it was efficiently quenched by highly reactive
oxygen species (hROS). The calculated detection limit was 30 nM,
with a linear response between 80 nM and 1.0 mM. The AuNCs@N-
MOFs could also be used to image exogenous and endogenous ROS
in living cells.

3. Metal-coordinated fluorescent and luminescent probes for


reactive nitrogen species (RNS)

Fig. 24. (a) Structure of Cu(I) complex, (b) a lanthanide duplex chain, (c) 3D 3.1. Fluorescent and luminescent probes for peroxynitrite (ONOO–)
framework of {[Eu2Cu(IN)5(CO3)(H2O)]}. Reproduced with permission. Copyright
2018, Wiley-VCH Verlag GmbH & Co. KGaA.
Like HOCl, ONOO– displays antibacterial efficacy, but it can also
damage DNA and proteins due to its oxidizing power [93]. In recent

Fig. 25. The reaction mechanism of UiO-68-PT with HOCl and its reversible colorimetric change with HOCl and vitamin C (VC) (left) (a) Emission spectra of UiO-68-PT upon
addition of HOCl. (b) Colorimetric changes of UiO-68-PT in the presence of various analytes. (c) Fluorescence changes of UiO-68-PT for various analytes, including HOCl.
Copyright 2019 American Chemical Society.

14
N. Kwon et al. Coordination Chemistry Reviews 427 (2021) 213581

Fig. 26. A ‘‘ship in a bottle” strategy from Zr6 clusters (inorganic node) and fumaric acid (organic ligand) to prepare LMOF-801 NPs and size-selective sensing of ClO.
Copyright 2019 The Royal Society of Chemistry.

Fig. 27. Proposed scheme for the assembly of CD/CCM@ZIF-8 and its ratiometric sensing of HOCl. Copyright 2020 American Chemical Society.

Fig. 28. Proposed mechanism of PVP-Capped CuNCs with HOCl. Copyright 2020 American Chemical Society.

decades, a variety of ONOO–-selective fluorescent and luminescent 403 nm, which corresponds to the MOFs. As shown in Fig. 30,
imaging probes have been actively investigated [28,94,95]. For ONOO– quickly converts arylboronates to hydroxyl derivatives.
sensing ONOO-, the nucleophilicity and oxidizability of ONOO Both MA and MB showed linear relationships with ONOO– in the
are utilized to design the probes. Oxidation of boronates or boronic range of 0.0–0.1 lM. Finally, both MA and MB could image exoge-
acids and oxidative N-dearylation are widely used. Boronic acids or nous ONOO– in HeLa cells via ratiometric fluorescence changes.
esters go through the nucleophilic addition with strong nucle- TD-DC was reported as a ratiometric, time-resolved lumines-
ophile, ONOO–, and at the end, it undergoes the subsequent hydrol- cent nanoprobe for ONOO– [97]. TD-DC is composed of Tb(DPA)3
ysis resulting in different photophysical properties. For oxidative and DC; the reaction mechanism of the DC moiety is detailed in
N-dearylation, PET-based fluorescence quenched group is oxidized Fig. 31. TD-DC displays emissions at 547 nm and 620 nm upon
to produce an iminium ion in the presence of ONOO–, followed by excitation at 330 nm. The addition of ONOO– induced a decrease
the subsequent hydrolysis to induce the release of the amino fluo- at 620 nm and an increase at 547 nm, resulting in a 48.7-fold ratio-
rophore [94]. metric enhancement (I547/I620). The reported detection limit was
Zhang and coworkers reported two nanoprobes, NMOF-PVA- 39 nM, with a fast response time (29 s). The addition of ONOO– also
Abt (MA) and NMOF-PVA-BDY (MB), as ratiometric fluorescent enhanced the lifetime from 100 ns at 620 nm to 1 ms at 547 nm.
sensors for ONOO– in which the Zr4+-encapsulated MOF was coated TD-DC was subsequently used to image endogenous ONOO– in
with poly(vinyl alcohol) (PVA) and the boronic acid derivatives, A375 cells. Imaging with Mito-Tracker Green showed that TD-DC
ABt and BDP, were attached through the formation of boronate was mainly localized in mitochondria, with a Pearson’s colocaliza-
esters (Fig. 30) [96]. When excited at 340 nm in PBS buffer at pH tion coefficient of 0.85.
7.4, MA displayed fluorescence at 530 nm, whereas MB emitted An iridium(III) based nanoprobe (MSN-ONOO) was reported as
at 610 nm via FRET. The addition of ONOO– induced a distinct a ONOO–-selective phosphorescent probe, in which a ONOO–-
ratiometric change in the emissions and a hypsochromic shift at triggered boronate-bearing Ir(III) derivative (Ir1) was embedded

15
N. Kwon et al. Coordination Chemistry Reviews 427 (2021) 213581

N
N

O
HO
HOOC
O
H 2N O
NH
NH

ONOO-
TD TD

TD-DC

Fig. 31. Reaction mechanism of TD-DC with ONOO–.


Fig. 29. Synthesis of AuNCs@NMOFs and hROS detection in living cells. Copyright
2019 Elsevier B.V.
10% CH3CN). The N-dearylation reaction by ONOO– [99] was also
within mesoporous silica nanoparticles (MSN-Ir3*) (Fig. 32) [98].
used in P-ONOO, as shown in Fig. 33a. The addition of ONOO–
At pH 7.4 (0.01 M PBS buffer containing 10% CH3CN), MSN-ONOO
induced a large increase in phosphorescence at 680 nm, but there
shows a strong emission at 474 nm and a moderate emission at
was no significant change in the emission at 600 nm. The calcu-
605 nm, which were attributed to the emissions from Ir1 and
lated detection limit was 0.018 lM, with a linear response range
Ir3*, respectively. The addition of ONOO– induced a large decrease
of 2–8 lM. When 3-morpholinosydnonimine hydrochloride (SIN-
of the emission at 474 nm but barely affected the emission at
1) was used to release endogenous ONOO– in RAW264.7 cells,
605 nm, which was accompanied by the color of the phosphores-
the green emission (575–625 nm) remained constant and the red
cence changing from white-pink to red. A linear relationship was
emission (655–705 nm) increased dramatically. ONOO– can be an
observed between 0 and 10 lM, with a detection limit of
early biomarker for drug-induced liver injury (DILI). A
11.3 nM. Upon the addition of ONOO– in RAW264.7 cells, a
ketoconazole-induced liver injury was successfully detected from
decrease in the green channel emission was observed, while the
the elevated ONOO– concentration using P-ONOO.
red emission remained nearly constant. MSN-ONOO was used to
A water-soluble luminescent Eu(III) probe bearing a quinoline
image endogenous ONOO– in zebrafish and living mice stimulated
moiety (17) (Fig. 34) was reported as a ONOO– selective probe that
with LPS and interferon gamma (IFN-c).
can show time-resolved luminescence at pH 7.4 (0.1 M PBS buffer)
The same group extended their work to a polymeric probe con-
[100]. Emission in the range of 605–720 nm was observed upon
taining the N-(4-hydroxylphenyl)-rhodol group (P-ONOO), which
excitation at 355 nm. The benzyl boronic moiety, which is the trig-
displayed dual emissions at 680 nm and 600 nm in PBS (pH 7.4,
ger group, was cleaved via oxidative cleavage by ONOO– to afford

Fig. 30. Proposed sensing mechanism of MOF-based sensors for ONOO– (MA and MB). Reproduced with permission. Copyright 2017 The Royal Society of Chemistry.

16
N. Kwon et al. Coordination Chemistry Reviews 427 (2021) 213581

Fig. 32. (a) Structures of iridium(III) complexes and reaction mechanism with ONOO–. (b) Synthesis of nanoprobe MSN-ONOO. (c) Reaction mechanism of MSN-ONOO with
ONOO–. Copyright 2018 American Chemical Society.

the 8-hydroxyquinoline moiety, as shown in Fig. 34. This can pro- NO-induced reduction of paramagnetic Cu(II) to diamagnetic Cu
hibit the energy transfer to Eu(III), resulting in a decrease in the (I) [104]. The reaction with the addition of NO undergoes through
time-resolved luminescence. A linear correlation was observed in the metal reduction without fluorophore displacement to form
the range of 1 nM to 1 lM and a nanomolar range of endogenous RONO and H+ [104]. Paramagenetic or heavy metal ion get dis-
ONOO– was successfully detected in living cells. Eu(III) probe 17 placement or coordination upon the interaction with NO. As
was mainly located in the mitochondria of HeLa cells, with a another method, o-diamine can be used and transformed into a flu-
reported Pearson’s coefficient of 0.8. orescent triazole derivative, which is electron-poor by the reacting
with NO. The PET get interfered by this chemical change and makes
3.2. Fluorescent and luminescent probes for nitric oxide (NO) and the probe show different photophysical properties [105].
nitrogen dioxide (NO2) Lippard and coworkers reported seminaphthofluorescein
derivatives 18a–d as NO-selective fluorescent probes (Fig. 35)
NO, which is endogenously generated by nitric oxide synthases, [106]. When NO was introduced to these probes, large fluorescence
can induce ailments such as cancer, neurodegenerative diseases enhancements (20–45-fold) occurred within the range of 550–
and stroke [16]. NO was selected as the molecule of the year in 625 nm upon excitation at 535–575 nm. The proposed explanation
1992 [101]. Among the various approaches to design a NO- for the fluorescence changes was Cu2+ to Cu+ reduction by NO fol-
selective probe [102,103], the Lippard group reported pioneering lowed by the elimination of Cu+ from the ligand and nitrosation of
work in which the manipulation of fluorescence was achieved via the secondary amine. Other RONS or ROS, such as HNO, NO 
2 , NO3 ,

17
N. Kwon et al. Coordination Chemistry Reviews 427 (2021) 213581

Fig. 33. (a) N-Dearylation reaction by ONOO–. (b) Structures of Ir(III) complexes (Ir1, Ir2, Ir3, and Ir4) and their N-dearylation sensing mechanisms with ONOO–. (c) Structures
of P-ONOO and its sensing mechanism with ONOO–. Copyright 2020 American Chemical Society.

HO
B OH

O N O O N O O N O
O O O
O O O O O O O
N ONOO- N Cu2+ NO
N N N N
Eu O O Eu O NH2 NH2 NH2
N N N N NH N N
O O Cu2+ Cu+ N
17 17-I X O
N N N
Emission ON Emission OFF X
19 X = anion

Fig. 34. Reaction mechanism of Eu(III) probe 17 with ONOO . Weak fluorescence Weak fluorescence

N
N N
O
R +
N
R -Cu
N
NO
N N
H N O O
Cu N O ON 19-I
Cl
O O HO O Strong fluorescence

X X Fig. 36. Reaction mechanism of 19 with Cu2+ and NO.


Cu(I) CO2H
CO2H

18a 18b 18c 18d


Ex: 340 nm Ex: 340 nm
X=H Br H Br
R=H H OCH2CO2Et OCH2CO2Et
HN N
Cu2+
Fig. 35. Reaction mechanisms of 18a–d with NO. O O
N N
N N Em: 481 nm
2+
Cu FRET
ONOO, ClO and H2O2, did not induce any significant fluorescence HO O
HO O OH O
change. The endogenous NO in RAW264.7 cells could be success-
fully imaged by these probes upon stimulation by LPS and INF-c. 20
NO Em: 518 nm
Qian and coworkers reported the o-phenylenediamine-bearing 20-Cu+-NO
. 2+
20 Cu
naphthalimide derivative 19 as a NO-selective fluorescent probe Fluorescence OFF
[107]. Probe 19 showed relatively weak fluorescence due to the
electron-withdrawing effect of the quinolone moiety. The addition Fig. 37. Reaction mechanism of 20 with Cu2+ and NO.

18
N. Kwon et al. Coordination Chemistry Reviews 427 (2021) 213581

of Cu2+ to probe 19 resulted in weak emission at 440 nm at pH 7.3 The unique ring-opening processes of rhodamines, which can
(Tris buffer containing 40% ethanol) and a distinct colorimetric induce the dramatic enhancement of fluorescence as well as dis-
change from yellow to pink. The addition of NO induced the tinct color changes, have been used to design various rhodamine-
removal of paramagnetic Cu2+ to afford the benzotriazole product, based fluorescent probes [111,112]. Duan and coworkers reported
as shown in Fig. 36, which induced a dramatic enhancement of flu- rhodamine-Cu2+ 23 as a NO-selective fluorescent probe in which a
orescence at 425 nm upon excitation at 350 nm and a color change rhodamine ring-opening process [113] and NO-induced reduction
from pink to colorless. The reported detection limit was 30 nM, from Cu2+ to Cu+ was followed by a reaction with NO to form the
with a linear response range of 0–80 mM. ring-opened product (Fig. 39b). Upon the addition of NO, strong
Iyer and coworkers reported a Cu2+ complex of a fluorescein fluorescence at 580 nm was observed upon excitation at 510 nm
hydrazone derivative (20) as a NO-selective fluorescent probe at pH 7.4 (0.1 M PBS buffer). The reported detection limit was
(Fig. 37) [108]. The spirolactam-ring-closed form of 20 showed 1 nM, with a linear response range of 1–10 nM. 23 was also used
very weak fluorescence from the indole moiety at 481 nm upon to detect exogenous NO in MCF-7 cells.
excitation at 340 nm. The addition of Cu2+ induced a spiro-ring- Nagano and coworkers reported Yb(2,7-dichlorocalcein) com-
opening process, resulting in strong green fluorescence at plex 24 as a NO-selective near-infrared luminescent probe [114].
518 nm attributed to FRET from indole to fluorescein. When NO Probe 24 is non-fluorescent due to PET from 2,7-dichlorocalcein
was added, fluorescence quenching was observed and attributed to Yb3+. The addition of NO induced the conversion of the o-
to fluorophore displacement or nitrosation. The reported detection diaminophenyl moiety to triazole (Fig. 40), and the corresponding
limit was 13.8 lL. 20 fluoresced strongly in RAW264.7 cells upon triazole product showed a strong emission at 980 nm upon excita-
the addition of Cu2+, which was quenched by the endogenous NO. tion at 500 nm at pH 7.2 (50 mM HEPES buffer). The reported
A Cu2+ complex of a dansyl derivative (21) was reported as a detection limit was 90 nM.
NO-selective fluorescent probe by Mondal and coworkers [109]. Duan and coworkers reported Ce4(H2TTS)4, a metal–organic
Upon excitation at 342 nm, fluorescent emission at 550 nm was polyhedra (MOPs)-based fluorescent probe, as a selective fluores-
revived due to the conversion of paramagnetic Cu2+ to diamagnetic cent probe for NO [115]. The amide-containing tridentate chelating
Cu+ by NO (Fig. 38). The calculated detection limit was  10 nM. ligand was synthesized from salicylaldehyde and 4,40 ,400 -nitrilotri
Another Cu2+ complex of a dansyl derivative (22) was reported benzocarbohydrazide, which reacted with Ce(NO3)36H2O to afford
as a NO-selective fluorescent probe at pH 7.20 (HEPES buffer, CH3- Ce4(H2TTS)4. As shown in Fig. 41, the resulting MOPs complex
CN–H2O 9:1 v/v) by Ali and coworkers [110]. The addition of NO accommodates 2-phenyl-4,4,5,5-tetramethylimidazoline-1-oxyl-3
reduces Cu2+ to Cu+, which can block PET and revive the fluores- -oxide (PTIO), a known NO trapper, in its cavity to detect NO.
cence at 532 nm upon excitation at 345 nm (Fig. 39a). The reported The fluorescence of Ce4(H2TTS)4 at 470 nm upon excitation at
detection limit was 1.6 nM. Probe 22 was subsequently used to 350 nm was quenched upon complexation with PTIO and revived
detect NO by addition of Angeli’s salt (Na2N2O3) in HeLa cells. upon the addition of NO. Other ROS and RNS did not induce any
significant fluorescence change. Ce4(H2TTS)4 successfully imaged
endogenous NO in MCF-7 cells with a detection limit of 5 nM.
The iridium(III) complex 25 bearing an o-diamino moiety was
N
N reported as a luminescent probe for NO (Fig. 42) [116]. Probe 25
showed a large enhancement of the emission at 580 nm at pH
7.4 (50 mM PBS buffer containing 10% DMSO) in the presence of
O S O sodium nitroprusside (SNP), which can controllably release NO.
O S O
N NO N The reported detection limit was 0.18 lM, with a linear response
range of 5–25 lM. Probe 25 was also successfully used to image
N N
N N N N endogenous NO in HeLa cells.
Cu2+ Cu+ UCNPs were investigated as a ratiometric fluorescent probe for
21 NO
NO [117]. As shown in Fig. 43, the UCNP core, which emits at
CH3OH
Fluorescent 540 nm and 656 nm, and a NO-selective rhodamine B derivative
were encapsulated within mesoporous SiO2, which was further
Fig. 38. Reaction mechanism of probe 21 with NO. decorated with a b-cyclodextrin (bCD) layer. The o-
phenylenediamine moiety of the rhodamine B derivative reacted
with NO to induce ring-opening followed by hydrolysis to rho-
(a) N N
damine B, the absorption bands of which (500–600 nm) overlap
with the green emission (540 nm) of UCNPs via LRET. The addition
NO of NO induced an increase of ratiometric change (I656/I540 ratio) in
O S O O S O
N N
N N N N
N N N Cu+ N
Cu2+
NH2 HN N
Cl Cl
NH2 N

22
(b) NH2 COOH COOH
Cl Cl Cl Cl
O NO
Cu2+ Cl
NO
N N NH2
N -
N
-
O O O O O O
NH2 NO O

NH2 COO- -OOC COO- -OOC


3+
N O N N O N
COO- Yb
3+
-
OOC COO- Yb -OOC

23 24

Fig. 39. Reaction mechanism of (a) probe 22 and (b) 23 with NO. Fig. 40. Proposed reaction mechanism of luminescent probe 24 with NO.

19
N. Kwon et al. Coordination Chemistry Reviews 427 (2021) 213581

Fig. 41. Structure of metal–organic polyhedra (MOPs)-based fluorescent probe Ce4(H2TTS)4 and its fluorescence change upon the addition of NO and 2-phenyl-4,4,5,5-
tetramethylimidazoline-1-oxyl-3-oxide (PTIO). Copyright 2011 American Chemical Society.

PET

N N
N NH2 N N
H 3C NO H3C
Ir Ir N
H 3C H 3C N
N NH2 N
H
N N

25
Fluorescent

Fig. 42. Reaction mechanism of Ir-based probe 25 with NO.

PBS (10 mM, pH 7.4). The reported detection limit was 73 nM, with for NO2 (Fig. 45) [119]. The hybrid nanoprobe 27 showed emissions
a linear response range of 7.4–110 lM. This probe system was suc- at 665 nm from the CdTe-QDs and 460 nm from the CDs upon exci-
cessfully used to detect exogenous NO in living HeLa cells, serum tation at 365 nm. Upon the addition of NO2, the emission at
and liver tissues of mouse models. 665 nm was quenched and the emission at 460 nm was almost
A sulforhodamine B fluorophore was linked to dithiocarbamate unchanged, producing a clear ratiometric change. A distinct fluo-
bound to Ni2+, which was reported as a NO2-selective fluorescent rescent color change from orange-red to blue was easily observed
probe in 10 mM PBS (pH 7.4). As shown in Fig. 44, a PET quenching with the naked eye. The reported detection limit was 19 nM, with a
effect due to Ni2+ was observed for 26 [118]. The fluorescence was linear response range of 0.2–1.2 lM.
revived by the oxidation of the dithiocarbamate ligand upon the
addition of NO2 to afford the decomplexed forms 26-I or 26-II. 3.3. Fluorescent and luminescent probes for nitroxyl (HNO)
Strong fluorescence at 590 nm was observed upon excitation at
540 nm when diethylamine NONOate was used as the NO donor. HNO, which is the one-electron reduced form of NO, mostly
The reported detection limit was 1 ppm for gaseous exogenous exists as the protonated form (pKa = 11.4) [120], which displays
NO2 with the naked eye when 26 was deposited on filter paper. different biochemistry than NO does [121]. Recently, the most
Finally, the probe system was used to image NO2 in RAW264.7 common design strategy of fluorescent HNO probes is to use the
cells. Cu2+/Cu+ redox couple, which is also commonly used for NO
Wang and coworkers reported a CdTe quantum dot (QD) and probes. However, in the presence of HNO, paramagnetic Cu2+ is
carbon dot (CD) hybrid system as a ratiometric fluorescent probe reduced to diamagnetic Cu+ and fluorescence is restored with the
20
N. Kwon et al. Coordination Chemistry Reviews 427 (2021) 213581

Fig. 43. Ratiometric upconversion nanoprobe for NO and its optical responses with NO.

S
SO3- R2N Ni NO2
NO2 S
S N N
NO2 S S HS
N Ni R2N Ni NR2 NR2 HN N
O 2S N S S S Br Cu2+ Cu2+ H Br
S
2 HO O O O O O
26
PET
N N
Cu+ HNO
S S O S
CH3OH S NR2 O
S NR2
28 28-Cu2+
OMe R2N S R2N
R2 N S S
S S
O
26-I Fig. 46. A reversible mechanism between 28 and 28-Cu2+ in the presence of HNO.
26-II
-
SO3

= R2N =
N
emission at 623 nm upon excitation at 570 nm, which can induce
Et2N O NEt2 O 2S N the reduction of Cu2+ to Cu+. Interestingly, the reduced Cu+ dissoci-
+
ated from the binding site and regenerated 131, which can provide
Fig. 44. The reaction mechanism of sulforhodamine B-dithiocarbamate-Ni2+ (26), a reversible sensing system. 131 was also reported to image exoge-
with NO2. nous HNO in RAW264.7 and HeLa cells.
Lippard and coworkers reinvestigated their previously reported
[124] fluorescent probe 29, [Cu(BOT1)Cl]Cl, for HNO. Probe 29 was
Ex = 365 nm Ex = 365 nm
prepared using ‘‘click” chemistry to connect BODIPY and tripodal
Cu2+ moieties via a triazole linker in improved yield, and its X-
2

2
2
NH

2
NH

NH

NH

GSH GSH GSH GSH ray crystal structure was newly reported [125]. The addition of
NH2 CDs
O NO2 O
HNO induced a  10-fold fluorescence enhancement at 526 nm.
N C S QDs GSH NH2 CDs N C S QDs GSH
H H As shown in Fig. 47, the fluorescence quenched by Cu2+ was
GSH GSH revived after reduction to Cu+. Probe 29 was used to image HNO
NH

GSH GSH
NH

x
NH

NH
2

2
2

in HeLa cells using 3 mM of Angeli’s salt, which was mostly local-


2

108
460 nm 665 nm 460 nm 665 nm
ized in the endoplasmic reticulum (ER) and Golgi with Pearson’s
Fig. 45. QD-CD hybrid probe 27 for NO2. coefficients of 0.51 and 0.63, respectively.
CuDHX1 was reported as a HNO-selective NIR fluorescent probe
by Lippard and Rivera-Fuentes et al. (Fig. 48) [126]. The addition of
addition of HNO [122]. The fluorescent probes binding with Cu2+ HNO induced the enhancement of fluorescence at 715 nm upon
often show weak fluorescence or even no fluorescence, because excitation at 650 nm at pH 7 (50 mM PIPES, 100 mM KCl). HNO
Cu2+ can quench the fluorescence by PET. HNO can selectively induced the conversion of Cu2+ to Cu+, which dissociated from
reduce Cu2+ to Cu+ which blocks PET and therefore, it recovers the cyclam ligand. CuDHX1 could selectively detect HNO over
the fluorescence of the fluorophore [120]. other ROS and RNS. CuDHX1 is a good NIR fluorescent probe that
A benzoresorufin derivative 28-Cu2+ was reported as a HNO- can detect HNO in HeLa cells even in the presence of GSH, Cys or
selective probe by Lippard et al. (Fig. 46) [123]. When CuCl2 was H2S. Angeli’s salt was used as the HNO precursor for cell imaging,
added, the fluorescence was quenched by the paramagnetic Cu2+. which is known to decompose to HNO and NaNO2 at pH 7 quickly
The addition of Angeli’s salt induced 4.8-fold enhancement of (t1/2 = 3 min).

21
N. Kwon et al. Coordination Chemistry Reviews 427 (2021) 213581

Cl
F Cl F Cl
F B N F B N
Cu2+ HNO Cu+
N N N
N N N
N N N
N
N N
2 2
29

Fig. 47. Proposed reaction mechanism of fluorescent probe 29, [Cu(BOT1)Cl]Cl, with HNO.

O OH

N N
CuDHX1 Cu
= 650 nm N N
ex
em = 715 nm

Fig. 48. Structure of CuDHX1 and its fluorescence microscopy images of HeLa cells. (A) Differential interference contrast (DIC) image; (B) blue region showing nuclei, (C) NIR
region, (D) NIR region after treatment with Angeli’s salt (1.5 mM). Scale bar = 25 lm. Copyright 2014 American Chemical Society.

Yao et al. reported coumarin-based probe 30 as a selective flu- Cu(II)-BTPY showed very weak fluorescence. Upon the addition of
orescent probe for HNO (Fig. 49a) [127]. The addition of HNO to 30 HNO, the conversion from Cu2+ into Cu+ induced enhancement of
induced 17.2-fold enhancement of fluorescence at 499 nm. The cal- emission. BTPY displayed a relatively high fluorescent quantum
culated association constant was 7.9  105 M1. HNO selectively yield of 34.8% with a stock shift of 62 nm. The efficient quenching
enhanced the fluorescence at 499 nm in HEPES buffer (50 mM, of fluorescence was attributed to the paramagnetism of Cu2+. The
0.1 M KNO3, pH 7.4), which was attributed to the reduction of reduction of Cu2+ to Cu+ upon the addition of HNO was confirmed
Cu2+ to Cu+. Probe 30 was further applied to image HNO in A375 by EPR. Cu(II)-BTPY displayed high selectivity for HNO over other
cells by pretreatment with Angeli’s salt (50 lM). ROS and RNS and was successfully used to image HNO in HeLa
Lin and Yao et al. extended their study to a Cu2+ complex of a cells.
coumarin derivative (31) as a HNO-selective fluorescent probe
(Fig. 49b) [128]. The addition of Cu2+ induced efficient quenching
of fluorescence intensity at 499 nm in HEPES buffer (50 mM, pH 4. Summary and outlook
7.4). The reported binding constant of 31 with Cu2+ was
8.34  105 M1. Of the various ROS and RNS, a 6.9-fold enhance- ROS and RNS play key roles in various signaling and pathologi-
ment of fluorescence intensity was observed when HNO was added cal processes and are essential to human life, but overproduction of
to 31-Cu(II), which was explained by the formation of 31-Cu(I). ROS/RNS by exogenous stimuli is harmful because ROS/RNS can
Exogenous HNO was successfully detected in NIH3T3 cells using induce oxidation of DNA, proteins or lipids, resulting in cell death.
31-Cu(II), which was further used to detect the same in the bacte- Therefore, unusual ROS/RNS levels can indicate ailments such as
riovorous nematode C. elegans. Parkinson’s and Alzheimer’s diseases, inflammation, diabetes and
A similar paramagnetic Cu2+ to diamagnetic Cu+ conversion cancer. In this review, we have discussed metal coordination com-
occurred in dansyl derivative DQ468 (Fig. 49c) [129]. PET fluores- plexes that act as fluorescent or otherwise luminescent probes for
cence quenching of the green emission at 543 nm was observed H2O2, HOCl/ClO–, OH, O– 1 –
2 , O2, NO, HNO, NO2 and ONOO . Lumines-
when paramagnetic Cu2+ was added. The reported detection limit cent detection methods have many advantages over other analyti-
of DQ468 for Cu2+ was 0.091 lM, and the reported detection limit cal methods, such as simplicity and sensitivity. Even more
of [CuII(DQ468)Cl]+ for HNO was 0.41 lM. HepG2 cells were used importantly, these methods could be used for cell, tissue and small
to prove the utility of this system. Cu2+ efficiently quenched the animal in vivo imaging as well as imaging-guided human surgery.
green fluorescence of DQ468, which was revived after treatment Self-immolative or chemodosimetric approaches have blos-
with Angeli’s salt at 37 °C; this behavior is attributed to the reduc- somed over the last decade, especially with respect to ROS- or
tion of Cu2+ to Cu+. RNS-related organic reactions that could be applied to imaging
A new BODIPY derivative bearing a terpyridyl moiety was ROS and RNS in living organisms, diagnostics or ROS- or RNS-
reported to selectively sense HNO by utilizing the redox reaction activated drug delivery systems. So far, the main strategies to
of Cu(II) [130]. As shown in Fig. 50, the BODIPY-Cu(II) complex design ROS- or RNS-selective probes have been protecting groups,
22
N. Kwon et al. Coordination Chemistry Reviews 427 (2021) 213581

(a)

N O O
N O O
N
N 2+ N N N Cu+ N
N N Cu HNO

N
N N
N
30

(b)

N O O

N O
N N O
31 N
N

Cu2+

N O O
N O O
N
N HNO N +O O
N 2+O O N Cu
N Cu N
N N
31-Cu(II) N 31-Cu(I)

(c)

N N N

CuCl2 HNO
O O O O O O
S S S
N N N
N Cu N Cu N
N N Cl N

DQ468
[CuII(DQ468)Cl]+
"ON" "ON"
"OFF"

Fig. 49. Reaction scheme for (a) 30, (b) 31-Cu(II) with HNO and (c) [CuII(DQ468)Cl]+ with HNO.

Fig. 50. The reaction scheme of Cu(II)-BTPY with HNO and its fluorescent change.

selective cleavages or destructive approaches and selective addi- to develop new classes of fluorophores with improved photo-
tions with specific ROS- or RNS-induced fluorescence enhance- properties, such as AIE [39], fluorescent probes for super-
ments or ratiometric changes. Researchers are also endeavoring resolution imaging [131], two-photon probes [132], NIR/NIR-II

23
N. Kwon et al. Coordination Chemistry Reviews 427 (2021) 213581

[133], TADF-based probes [134], upconversion probes [135] and [4] C.C. Winterbourn, Reconciling the chemistry and biology of reactive oxygen
species, Nat. Chem. Biol. 4 (2008) 278–286.
MOFs [136].
[5] S. Heinzelmann, G. Bauer, Multiple protective functions of catalase against
Metal coordination of the metal complex plays an essential role intercellular apoptosis-inducing ROS signaling of human tumor cells, Biol.
in the design of photoluminescent probes. Metal ions play key roles Chem. 391 (2010) 675–693.
in the probes as an on–off redox switch for photoinduced quench- [6] J. Zielonka, B. Kalyanaraman, Small-molecule luminescent probes for the
detection of cellular oxidizing and nitrating species, Free Radical Biol. Med.
ing and as a reaction site with ROS, RNS or luminescent cores. As 128 (2018) 3–22.
shown in this review, redox-active metal ions can be trigger [7] S.J. Weiss, G.W. King, A.F. LoBuglio, Evidence for hydroxyl radical generation
switches that control fluorescence quenching effects, which can by human monocytes, J. Clin. Invest. 60 (1977) 370–373.
[8] J.F. Turrens, Mitochondrial formation of reactive oxygen species, J. Physiol.
be used to sense ROS or RNS. Most common example was NO- 552 (2003) 335–344.
induced reduction of paramagnetic Cu2+ to diamagnetic Cu+. In [9] W.M. Nauseef, Detection of superoxide anion and hydrogen peroxide
addition, metal ions, especially lanthanide metal ions, can often production by cellular NADPH oxidases, Biochim. Biophys. Acta 1840 (2)
(2014) 757–767.
be themselves a source of light emission. In most cases, metal com- [10] N.M. Domigan, T.S. Charlton, M.W. Duncan, C.C. Winterbourn, A.J.J. Kettle,
plexes showed several advantageous photophysical properties Chlorination of tyrosyl residues in peptides by myeloperoxidase and human
compared to those of organic fluorophore, such as a long emission neutrophils, Biol. Chem. 270 (1995) 16542–16548.
[11] P.F. Bove, A. van der Vliet, Nitric oxide and reactive nitrogen species in airway
lifetime, high luminescence efficiency, and large Stokes shift. In epithelial signalling and inflammation, Free Radical Biol. Med. 41 (2006) 515–
addition, metal nanoparticles or MOFs are powerful methods to 527.
introduce more than one function or provide a dramatically [12] J.S. Beckman, T.W. Beckman, J. Chen, P.A. Marshall, B.A. Freeman, Apparent
hydroxyl radical production by peroxynitrite: Implications for endothelial
improved photophysical properties using a combination of lumi-
injury from nitric oxide and superoxide, Proc. Natl. Acad. Sci. U.S.A. 87 (1990)
nescent sources. 1620–1624.
Fluorescent probes that target specific organelles and fluores- [13] C. Szabo, H. Ischiropoulos, R. Radi, Peroxynitrite: biochemistry,
cently label ROS/RNS provide information on various diseases pathophysiology and development of therapeutics, Nat. Rev. Drug Disc. 6
(2007) 662–680.
including cancer and neurodegenerative disorders. The localization [14] X. Chen, F. Wang, J.Y. Hyun, T. Wei, J. Qiang, X. Ren, I. Shin, J. Yoon, Recent
of the probe provided important information on oxidative stress progress in the development of fluorescent, luminescent and colorimetric
specific to the cellular organelles where the probe was accumu- probes for detection of reactive oxygen and nitrogen species, Chem. Soc. Rev.
45 (2016) 2976–3016.
lated. Various probes introduced in this review are already suc- [15] Y. You, W. Nam, Designing photoluminescent molecular probes for singlet
cessfully applied for disease diagnostics. For example, a oxygen, hydroxyl radical, and iron–oxygen species, Chem. Sci. 5 (2014) 4123–
ketoconazole-induced liver injury was successfully detected from 4135.
[16] K.-C. Yan, A.C. Sedgwick, Y. Zang, G.-R. Chen, X.-P. He, J. Li, J. Yoon, T.D. James,
the elevated ONOO– concentration by Chen, et al. [99]. For the Sensors, imaging agents and theranostics to help understand and treat
examples to utilize redox-active metal ions, the Lippard group reactive oxygen species (ROS) related diseases, Small Methods 3 (2019)
reported pioneering and important work, in which fluorescent 1900013.
[17] J.-T. Hou, K.-K. Yu, K. Sunwoo, W.Y. Kim, S. Koo, J. Wang, W.X. Ren, S. Wang,
change was achieved via NO-induced reduction of paramagnetic X.-Q. Yu, J.S. Kim, Fluorescent imaging of reactive oxygen and nitrogen
Cu2+ to diamagnetic Cu+ [104]. This work especially influenced species associated with pathophysiological processes, Chem 6 (2020) 832–
the following related works from various other groups. 866.
[18] L. Wu, A.C. Sedgwick, X. Sun, S.D. Bull, X.-P. He, T.D. James, Reaction-based
As probes for tumors or other sites with especially high levels of
fluorescent probes for the detection and imaging of reactive oxygen, nitrogen,
ROS or RNS become essential for imaging-guided surgery and and sulfur species, Acc. Chem. Res. 52 (2019) 2582–2597.
improved therapeutic and theranostic precision, fluorescent probes [19] A. Kaur, E.J. New, Bioinspired Small-Molecule Tools for the Imaging of Redox
that can dynamically monitor the level of ROS or RNS are urgently Biology, Acc. Chem. Res. 52 (2019) 623–632.
[20] X. Jian, L. Wang, S.L. Carroll, J. Chen, M.C. Wang, J. Wang, Challenges and
needed. The obstacles to overcome in the development of probes opportunities for small-molecule fluorescent probes in redox biology
for practical applications include high selectivity with high photo- applications, Antiox. Redox Signal 29 (2018) 518–540.
stability as well as reversibility. Some key issues such as biocom- [21] A. Kaur, J.L. Kolanowski, E.J. New, Reversible Fluorescent Probes for Biological
Redox States, Angew. Chem. Int. Ed. 55 (2016) 1602–1613.
patibility and organ toxicity of metal complexes are still needed [22] D. Andina, J. –C. Leroux, P. luciani, Ratiometric Fluorescent Probes for the
to be improved. It is also necessary to increase the rate of intracel- Detection of Reactive Oxygen Species, Chem. Eur. J. 23 (2017) 13549-13573.
lular absorption. We believe researchers will make great contribu- [23] J. Yan, S. Lee, A. Zhang, J. Yoon, Self-immolative colorimetric, fluorescent and
chemiluminescent chemosensors, Chem. Soc. Rev. 47 (2018) 6900–6916.
tions to this field for decades to come. [24] Y. Tang, D. Lee, J. Wang, G. Li, W. Lin, J. Yoon, Development of fluorescent
probes based on protection-deprotection of aldehyde, hydroxyl, and amino
Declaration of Competing Interest functional groups for biological imaging, Chem. Soc. Rev. 44 (2015) 5003–
5015.
[25] A.E. Albers, V.S. Okreglak, C.J. Chang, A FRET-based approach to ratiometric
The authors declare that they have no known competing finan- fluorescence detection of hydrogen peroxide, J. Am. Chem. Soc. 128 (2006)
cial interests or personal relationships that could have appeared 9640–9641.
[26] Q. Xu, K.-A. Lee, S. Lee, K.M. Lee, W.-J. Lee, J. Yoon, A highly specific fluorescent
to influence the work reported in this paper. probe for hypochlorous acid and its application in imaging microbe-induced
HOCl production, J. Am. Chem. Soc. 135 (2013) 9944–9949.
Acknowledgement [27] Sun, Q. Xu, G. Kim, S. E. Flower, J. P. Lowe, J. Yoon, J. S. Fossey, X. Qian, S. D.
Bull, Tony D. James, A water-soluble boronate-based fluorescent probe for the
selective detection of peroxynitrite and imaging in living cells, Chem. Sci. 5
This work was supported by the National Research Foundation (2014) 3368-3373.
of Korea (NRF) grant funded by the Korea government (MSIP) (No. [28] D. Lee, C.S. Lim, G. Ko, D. Kim, M.K. Cho, S.-J. Nam, H.M. Kim, J. Yoon, A two
photon fluorescent probe for imaging endogenous ONOO- near NMDA
2012R1A3A2048814).
receptors in neuronal cells and hippocampal tissues, Anal. Chem. 90 (2018)
9347–9352.
References [29] Y.L. Pak, S.J. Park, D. Wu, B. Cheon, H.M. Kim, J. Bouffard, J. Yoon, N-
Heterocyclic carbene boranes as reactive oxygen species-responsive
materials: Application to the two-photon imaging of hypochlorous acid in
[1] B. Halliwell, J.M.C. Gutteridge, Free Radicals in Biology and Medicine, Oxford
living cells and tissues, Angew. Chem. Int. Ed. 57 (2018) 1567–1571.
University Press, Oxford, 2007, pp. 1–677.
[30] Y.L. Pak, S.J. Park, G. Song, Y. Yim, H. Kang, H.M. Kim, J. Bouffard, J. Yoon,
[2] B. Kalyanaraman, G. Cheng, M. Hardy, O. Ouari, B. Bennett, J. Zielonka,
Endoplasmic Reticulum-Targeted Ratiometric N-Heterocyclic Carbene Borane
Teaching the basics of reactive oxygen species and their relevance to cancer
Probe for Two-Photon Microscopic Imaging of Hypochlorous Acid, Anal.
biology: Mitochondrial reactive oxygen species detection, redox signaling,
Chem. 90 (2018) 12937–12943.
and targeted therapies, Redox Biol. 15 (2017) 347–362.
[31] F.A. Villamena, in: Molecular Basis of Oxidative Stress: Chemistry,
[3] G.J. Maghzal, K.H. Krause, R. Stocker, V. Jaquet, Detection of reactive oxygen
Mechanisms, and Disease Pathogenesis, John Wiley & Sons Inc, Hoboken, N.
species derived from the family of NOXNADPH oxidases, Free Radic. Biol.
J, 2013, pp. 1–48.
Med. 53 (2012) 1903–1918.

24
N. Kwon et al. Coordination Chemistry Reviews 427 (2021) 213581

[32] A.B. Aletti, D.M. Gillen, T. Gunnlaugsson, Luminescent/colorimetric probes [59] L. Zhang, R.P. Liang, S.J. Xiao, J.M. Bai, L.L. Zheng, L. Zhan, X.J. Zhao, J.D. Qiu, C.
and (chemo-) sensors for detecting anions based on transition and Z. Huang, DNA-templated Ag nanoclusters as fluorescent probes for sensing
lanthanide ion receptor/binding complexes, Coord. Chem. Rev. 354 (2018) and intracellular imaging of hydroxyl radicals, Talanta 118 (2014) 339–347.
98–120. [60] Z. Li, T. Liang, S.W. Lv, Q.G. Zhuang, Z.H. Liu, A rationally designed
[33] X. Wang, H. Chang, J. Xie, B. Zhao, B. Liu, S. Xu, W. Pei, N. Ren, L. Huang, w. upconversion nanoprobe for in vivo detection of hydroxyl radical, J. Am.
Huang, Recent developments in lanthanide-based luminescent probes, Coord. Chem. Soc. 137 (2015) 11179–11185.
Chem. Rev. 273-274 (2014) 201-212. [61] V.-N. Nguyen, S. Qi, S. Kim, N. Kwon, G. Kim, Y. Yim, S. Park, J. Yoon, An
[34] S.M. Ng, M. Koneswaran, R. Narayanaswamy, A review on fluorescent emerging molecular design approach to heavy-atom-free photosensitizers for
inorganic nanoparticles for optical sensing applications, RSC Adv. 6 (2016) enhanced photodynamic therapy under hypoxia, J. Am. Chem. Soc. 141
21624–21661. (2019) 16243–16248.
[35] Y. Zhang, C. Zhang, C. Xu, X. Wang, C. Liu, G.I.N. Waterhouse, Y. Wang, H. Yin, [62] X. Li, C-y. Kim, S. Lee, D. Lee, H.-M. Chung, G. Kim, S.-H. Heo, C. Kim, K. S.
Ultra small Au nanoclusters for biomedical and biosensing applications: A Hong, J. Yoon, Nanostructured Phthalocyanine Assemblies with Protein-
mini-review, Talanta 200 (2019) 432–442. Driven Switchable Photoactivities for Biophotonic Imaging and Therapy, J.
[36] O. Adegoke, P.B.C. Forbes, Challenges and advances in quantum dot Am. Chem. Soc. 139 (2017) 10880-10886.
fluorescent probes to detect reactive oxygen and nitrogen species: A [63] S. Kolemen, T. Ozdemir, D. Lee, G.M. Kim, T. Karatas, J. Yoon, E.U. Akkaya,
review, Anal. Chim. Acta 862 (2015) 1–13. Remote controlled release of singlet oxygen via plasmonic heating of gold
[37] W. Zheng, P. Huang, D. Tu, En. Ma, H. Zhu, X. Chen, Chem. Soc. Rev. 44 (2015) nanorods: Towards a paradigm change in photodynamic therapy, Angew.
1379-1415. Chem. Int. Ed. 55 (2016) 3606–3610.
[38] P. Horcajada, R. Gref, T. Baati, P.K. Allan, G. Maurin, P. Couvreur, G. Férey, R.E. [64] X. Li, N. Kwon, T. Guo, Z. Liu, J. Yoon, Innovative strategies for hypoxic-tumor
Morris, C. Serre, Metal-organic frameworks in biomedicine, Chem. Rev. 112 photodynamic therapy, Angew. Chem. Int. Ed. 57 (2018) 11522–11531.
(2012) 1232–1268. [65] Z.C. Dai, L. Tian, Y.N. Xiao, Z.Q. Ye, R. Zhang, J.L. Yuan, A cell-membrane-
[39] H.-B. Cheng, Y. Li, B.Z. Tang, J. Yoon, Assembly strategies of organic-based permeable europium complex as an efficient luminescent probe for singlet
imaging agents for fluorescence and photoacoustic bioimaging applications, oxygen, J. Mater. Chem. B 1 (2013) 924–927.
Chem. Soc. Rev. 49 (2020) 21–31. [66] J. Gao, C. Wang, H. Tan, Dual-emissive polystyrene@zeolitic imidazolate
[40] T.P. Szatrowski, C.F. Nathan, Production of large amounts of hydrogen framework-8 composite for ratiometric detection of singlet oxygen, J. Mater.
peroxide by human tumor cells, Cancer Res. 51 (1991) 794–798. Chem. B 5 (2017) 9175–9182.
[41] D. Kim, G. Kim, S.-J. Nam, J. Yin, J. Yoon, Visualization of endogenous and [67] H.-J. Yin, Y.-J. Liu, J. Gao, K.-Z. Wang, A highly sensitive and selective visible-
exogenous hydrogen peroxide using a lysosome-targetable fluorescent probe, light excitable luminescent probe for singlet oxygen based on a dinuclear
Scientific Reports 5 (2015) 8488. ruthenium complex, Dalton Trans. 46 (2017) 3325–3331.
[42] J. Jing, J.L. Zhang, Combining myeloperoxidase (MPO) with fluorogenic [68] S.N.A. Jenie, S.E. Plush, N.H. Voelcker, Singlet oxygen detection on a
ZnSalen to detect lysosomal hydrogen peroxide in live cells, Chem. Sci. 4 nanostructured porous silicon thin film via photonic luminescence
(2013) 2947–2952. enhancements, Langmuir 33 (2017) 8606–8613.
[43] D. Song, J.M. Lim, S. Cho, S.J. Park, J. Cho, D. Kang, S.G. Rhee, Y. You, W. Nam, A [69] I. Fridovich, Superoxide anion radical (O– 2 ), superoxide dismutases, and
fluorescence turn-on H2O2 probe exhibits lysosome-localized fluorescence related matters, J. Biol. Chem. 272 (1997) 18515–18517.
signals, Chem. Commun. 48 (2012) 5449–5451. [70] N. Li, H. Wang, M. Xue, C.Y. Chang, Z.Z. Chen, L.H. Zhuo, B. Tang, A highly
[44] Y. Hitomi, T. Takeyasu, M. Kodera, Iron complex-based fluorescent probes for selective and sensitive nanoprobe for detection and imaging of the
intracellular hydrogen peroxide detection, Chem. Commun. 49 (2013) 9929– superoxide anion radical in living cells, Chem. Commun. 48 (2012) 2507–
9931. 2509.
[45] D. Chakraborty, S. Sarkar, P.K. Das, Blood dots: Hemoglobin-derived carbon [71] Z. Li, L. Xiao, Facile sonochemical synthesis of water-soluble gold nanodots as
dots as hydrogen peroxide sensors and pro-drug activators, ACS Sustainable fluorescent probes for superoxide radical anion detection and cell imaging,
Chem. Eng. 6 (2018) 4661–4670. Anal. Methods 9 (2017) 1920–1927.
[46] J. Chen, Y. Gao, Q. Ma, X. Hu, Y. Xu, X. Lu, Turn-off fluorescence sensor based [72] A. Das, N. Anbu, M. SK, A. Dhakshinamoorthy, S. Biswas, A functionalized UiO-
on the 5,10,15,20-(4-sulphonatophenyl) porphyrin (TPPS4)-Fe2+system: 66 MOF for turn-on fluorescence sensing of superoxide in water and efficient
Detecting of hydrogen peroxide (H2O2) and glucose in the actual sample, catalysis for Knoevenagel condensation, Dalton Trans., 2019, 48, 17371-
Sens. Actuators B 268 (2018) 270–277. 17380.
[47] F. Liu, T. Bing, D. Shangguan, M. Zhao, N. Shao, Ratiometric fluorescent [73] S.J. Klebanoff, Myeloperoxidase: Friend and Foe, J. Leukocyte. Biol. 77 (2005)
biosensing of hydrogen peroxide and hydroxyl radical in living cells with 598–625.
lysozyme-silver nanoclusters: Lysozyme as stabilizing ligand and [74] N. Güngör, A.M. Knaapen, A. Munnia, M. Peluso, G.R. Haenen, R.K. Chiu, R.W.L.
fluorescence signal unit, Anal. Chem. 88 (2016) 10631–10638. Godschalk, F.J. Schooten, Genotoxic effects of neutrophils and hypochlorous
[48] C. Li, S. Wang, Y. Huang, Q. Wen, L. Wang, Y. Kan, Photoluminescence acid, Mutagenesis. 25 (2010) 149–154.
properties of a novel cyclometalated iridium(III) complex with coumarin- [75] D. Wu, L. Chen, Q. Xu, X. Chen, J. Yoon, Design principles, sensing mechanisms
boronate and its recognition of hydrogen peroxide, Dalton Trans. 43 (2014) and applications of highly specific fluorescent probes for HOCl/OCl, Acc.
5595–5602. Chem. Res. 52 (8) (2019) 2158–2168.
[49] H.S. Wang, W.J. Bao, S.B. Ren, M. Chen, K. Wang, X.H. Xia, Fluorescent sulfur- [76] Q. Xu, C.H. Heo, G. Kim, H.W. Lee, H.M. Kim, J. Yoon, Development of
tagged europium(III) Coordination polymers for monitoring reactive oxygen imidazoline-2-Thiones Based two-photon fluorescence probes for imaging
species, Anal. Chem. 87 (2015) 6828–6833. hypochlorite generation in a co-culture system, Angew. Chem. Int. Ed. 54
[50] C. Hao, X. Wu, M. Sun, H. Zhang, A. Yuan, L. Xu, C. Xu, H. Kuang, Chiral core- (2015) 4890–4894.
shell upconversion nanoparticle@MOF nanoassemblies for quantification and [77] Y.-R. Zhang, Y. Liu, X. Feng, B.-X. Zhao, Recent progress in the development of
bioimaging of reactive oxygen species in Vivo, J. Am. Chem. Soc. 141 (2019) fluorescent probes for the detection of hypochlorous acid, Sensors and
19373–19378. Actuators B 240 (2017) 18–36.
[51] M. Sk, S. Banesh, V. Trivedi, S. Biswas, Selective and sensitive sensing of [78] X. Chen, K.-A. Lee, X. Ren, J.-C. Ryu, G. Kim, J.-H. Ryu, W.-J. Lee, J. Yoon,
hydrogen peroxide by a boronic acid functionalized metal-organic framework Synthesis of a highly HOCl -selective fluorescent probe and its use for imaging
and its application in live-cell imaging, Inorg. Chem. 57 (2018) 14574–14581. HOCl in cells and organisms, Nat. Protocol. 11 (2016) 1219–1228.
[52] H. Sies, Strategies of antioxidant defense, Eur. J. Biochem. 215 (1993) 213– [79] N. Zhao, Y.-H. Wu, R.-M. Wang, L.-X. Shi, Z.-N. Chen, An iridium(III) complex
219. of oximated 2,20 -bipyridine as a sensitive phosphorescent sensor for
[53] S.E. Page, K.T. Wilke, V.C. Pierre, Sensitive and selective time-gated hypochlorite, Analyst 136 (2011) 2277–2282.
luminescence detection of hydroxyl radical in water, Chem. Commun. 46 [80] Z.Q. Ye, R. Zhang, B. Song, Z.C. Dai, D.Y. Jin, E.M. Goldys, J.L. Yuan,
(2010) 2423–2425. Development of a functional ruthenium(II) complex for probing
[54] J.-T. Hou, M. Zhang, Y. Liu, X. Ma, R. Duan, X. Cao, F. Yuan, Y.-X. Liao, S. Wang, hypochlorous acid in living cells, Dalton Trans. 43 (2014) 8414–8420.
W.X. Ren, Fluorescent detectors for hydroxyl radical and their applications in [81] R. Zhang, Z.Q. Ye, B. Song, Z.C. Dai, X. An, J.L. Yuan, Development of a
bioimaging: A review, Coord. Chem. Rev. 421 (2020) 213457. Ruthenium(II) Complex-Based Luminescent Probe for Hypochlorous Acid in
[55] M.L. Aulsebrook, B. Graham, M.R. Grace, K.L. Tuck, Lanthanide complexes for Living Cells, Inorg. Chem. 52 (2013) 10325–10331.
luminescence-based sensing of low molecular weight analytes 375 (2018) [82] Z. Lu, M. Shangguan, X. Jiang, P. Xu, L. Hou, T. Wang, A water-soluble
191–220. cyclometalated iridium(III) complex with fluorescent sensing capability for
[56] G.F. Cui, Z.Q. Ye, J.X. Chen, G.L. Wang, J.L. Yuan, Development of a novel hypochlorite, Dyes Pigments 171 (2019) 107715.
terbium(III) chelate-based luminescent probe for highly sensitive time- [83] H. Ma, B. Song, Y. Wang, C. Liu, X. Wang, J. Yuan, Development of organelle-
resolved luminescence detection of hydroxyl radical, Talanta 84 (2011) 971– targetable europium complex probes for time-gated luminescence imaging of
976. hypochlorous acid in live cells and animals, Dyes Pigments 140 (2017) 407–
[57] M. Zhuang, C. Ding, A. Zhu, Y. Tian, Ratiometric fluorescence probe for 416.
monitoring hydroxyl radical in live cells based on gold nanoclusters, Anal. [84] L. Tian, H. Ma, B. Song, Z. Dai, X. Zheng, R. Zhang, K. Chen, J. Yuan, Time-gated
Chem. 86 (2014) 1829–1836. luminescence probe for ratiometric and luminescence lifetime detection of
[58] J. Li, J. Yu, Y. Huang, H. Zhao, L. Tian, Highly stable and multiemissive silver Hypochorous acid in lysosomes of live cells, Talanta 212 (2020) 120760.
nanoclusters synthesized in situ in a DNA hydrogel and their application for [85] Y. Gao, J.H.K. Yip, Selective hypochlorous acid detection by electronic tuning
hydroxyl radical sensing, ACS Appl. Mater. Interfaces 10 (2018) 26075– of platinum4,5-bis(diphenylphosphino)acridinethiolate complexes, Inorg.
26083. Chem. 58 (2019) 9290–9302.

25
N. Kwon et al. Coordination Chemistry Reviews 427 (2021) 213581

[86] Y.-A. Li, S. Yang, Q.-Y. Li, J.-P. Ma, S. Zhang, Y.-B. Dong, UiO-68-ol NMOF-based [112] X. Chen, T. Pradhan, F. Wang, J.S. Kim, J. Yoon, Fluorescent chemosensors
fluorescent sensor for selective detection of HClO and its application in based on spiroring-opening of xanthenes and related derivatives, Chem. Rev.
bioimaging, Inorg. Chem. 56 (2017) 13241–13248. 112 (2012) 1910–1956.
[87] H. Xu, C.-S. Cao, J.-Z. Xue, Y. Xu, B. Zhai, B. Zhao, A cuprous/lanthanide-organic [113] X.Y. Hu, J. Wang, X. Zhu, D.P. Dong, X.L. Zhang, S.O. Wu, C.Y. Duan, A copper
framework as the luminescent sensor of hypochlorite, Chem.-Eur. J. 24 (2018) (II) rhodamine complex with a tripodal ligand as a highly selective
10296. fluorescence imaging agent for nitric oxide, Chem. Commun. 47 (2011)
[88] Q.-Y. Li, Y.-A. Li, Q. Guan, W.-Y. Li, X.-J. Dong, Y.-B. Dong, UiO-68-PT MOF- 11507–11509.
based sensor and its mixed matrix membrane for detection of HClO in Water, [114] T. Terai, Y. Urano, S. Izumi, H. Kojima, T. Nagano, A practical strategy to create
Inorg. Chem. 58 (2019) 9890–9896. near-infrared luminescent probes: conversion from fluorescein-based
[89] Y. Ye, L. Zhao, S. Hu, A. Liang, Y. Li, Q. Zhuang, G. Tao, J. Gu, Specific detection sensors, Chem. Commun. 48 (2012) 2840–2842.
of hypochlorite based on the size-selective effect of luminophore integrated [115] J. Wang, C. He, P.Y. Wu, J. Wang, C.Y. Duan, An amide-containing metal-
MOF-801 synthesized by a one-pot strategy, Dalton Trans. 48 (2019) 2617– organic tetrahedron responding to a spin-trapping reaction in a fluorescent
2625. enhancement manner for biological imaging of NO in living cells, J. Am.
[90] H. Tan, X. Wu, Y. Weng, Y. Lu, Z.-Z. Huang, Self-assembled FRET nanoprobe Chem. Soc. 133 (2011) 12402–12405.
with metal-organic framework as a scaffold for ratiometric detection of [116] C. Wu, K.-J. Wu, T.-S. Kang, H.-M.D. Wang, C.-H. Leung, J.-B. Liu, D.-L. Ma,
hypochlorous acid, Anal. Chem. 92 (2020) 3447–3454. Iridium-based probe for luminescent nitric oxide monitoring in live cells, Sci.
[91] W. Dong, C. Sun, M. Sun, H. Ge, A.M. Asiri, H.M. Marwani, R. Ni, S. Wang, Rep. 8 (2018) 12467.
Fluorescent copper nanoclusters for the iodide-enhanced detection of [117] N. Wang, X. Yu, K. Zhang, C.A. Mirkin, J. Li, Upconversion nanoprobes for the
hypochlorous acid, ACS Appl. Nano Mater. 3 (2020) 312–318. ratiometric luminescent sensing of nitric oxide, J. Am. Chem. Soc. 139 (2017)
[92] X. Cao, S. Cheng, Y. You, S. Zhang, Y. Xian, Sensitive monitoring and 12354–12357.
bioimaging intracellular highly reactive oxygen species based on gold [118] Y. Yan, S. Krishnakumar, H. Yu, S. Ramishetti, L.W. Deng, S.H. Wang, L. Huang,
nanoclusters@nanoscale metal-organic frameworks, Anal. Chim. Acta 1092 D.J. Huang, Nickel(II) Dithiocarbamate complexes containing sulforhodamine
(2019) 108–116. b as fluorescent probes for selective detection of nitrogen dioxide, J. Am.
[93] R.T. Dean, S. Fu, R. Stocker, M.J. Davies, Biochemistry and pathology of radical- Chem. Soc. 135 (2013) 5312–5315.
mediated protein oxidation, Biochem. J. 324 (1997) 1–18. [119] Y. Yan, J. Sun, K. Zhang, H. Zhu, H. Yu, M. Sun, D. Huang, S. Wang, Visualizing
[94] S. Wang, L. Chen, P. Jangili, A. Sharma, W. Li, J.-T. Hou, C. Qin, J. Yoon, J.S. Kim, gaseous nitrogen dioxide by ratiometric fluorescence of carbon nanodots-
Design and applications of fluorescent detectors for peroxynitrite, Coord. quantum dots hybrid, Anal. Chem. 87 (2015) 2087–2093.
Chem. Rev. 374 (2018) 36–54. [120] B. Dong, X. Kong, W. Lin, Reaction-based fluorescent probes for the imaging of
[95] X. Zhou, Y. Kwon, G. Kim, J.-H. Ryu, J. Yoon, A ratiometric fluorescent probe nitroxyl (HNO) in biological systems, ACS Chem. Biol. 13 (2018) 1714–1720.
based on a coumarin-hemicyanine scaffold for sensitive and selective [121] Vladimir Shafirovich 1, Sergei V Lymar, Nitroxyl and Its Anion in Aqueous
detection of endogenous peroxynitrite, Biosens. Bioelectron. 64 (2015) 285– Solutions: Spin States, Protic Equilibria, and Reactivities Toward Oxygen and
291. Nitric Oxide, Proc Natl Acad Sci U S A 2002 May 28;99(11):7340-5. doi:
[96] Z. Ding, J. Tan, G. Feng, Z. Yuan, C. Wu, X. Zhang, Nanoscale metal-organic 10.1073/pnas.112202099.
frameworks coated with poly(vinyl alcohol) for ratiometric peroxynitrite [122] J. Alday, A. Mazzeo, S. Suarez, Selective detection of gasotransmitters using
sensing through FRET, Chem. Sci. 8 (2017) 5101–5106. fluorescent probes based on transition metal complexes, Inorganica Chimica
[97] Y. Wang, B. Li, X. Song, R. Shen, D. Wang, Y. Yang, Y. Feng, C. Cao, G. Zhang, W. Acta 510 (2020) 119696.
Liu, Mito-specific Ratiometric Terbium(III)-complex-based luminescent [123] U.P. Apfel, D. Buccella, J.J. Wilson, S.J. Lippard, Detection of nitric oxide and
probe for accurate detection of endogenous peroxynitrite by time-resolved nitroxyl with benzoresorufin-based fluorescent sensors, Inorg. Chem. 52
luminescence assay, Anal. Chem. 91 (2019) 12422–12427. (2013) 3285–3294.
[98] Z. Chen, P. Yan, L. Zou, M. Zhao, J. Jiang, S. Liu, K.Y. Zhang, W. Huang, Q. [124] J. Rosenthal, S.J. Lippard, Direct detection of nitroxyl in aqueous solution
Zhao, Using ultrafast responsive phosphorescent nanoprobe to visualize using a tripodal Copper(II) BODIPY complex, J. Am. Chem. Soc. 132 (2010)
elevated peroxynitrite in vitro and in vivo via ratiometric and time- 5536–5537.
resolved photoluminescence imaging, Adv. Healthcare Mater. 7 (2018) [125] M. Royzen, J.J. Wilson, S.J. Lippard, Physical and structural properties of [Cu
1800309. (BOT1)Cl]Cl, a fluorescent imaging probe for HNO, J. Inorg. Biochem. 118
[99] Z. Chen, X. Meng, L. Zou, M. Zhao, S. Liu, P. Tao, J. Jiang, Q. Zhao, A dual- (2013) 162–170.
emissive phosphorescent polymeric probe for exploring drug-induced liver [126] A.T. Wrobel, T.C. Johnstone, A.D. Liang, S.J. Lippard, P. Rivera-Fuentes, A fast
injury via imaging of peroxynitrite elevation in vivo, ACS Appl. Mater. and selective near-infrared fluorescent sensor for multicolor imaging of
Interfaces 12 (2020) 12383–12394. biological nitroxyl (HNO), J. Am. Chem. Soc. 136 (2014) 4697–4705.
[100] C. Breen, R. Pal, M.R.J. Elsegood, S.J. Teat, F. Iza, K. Wende, B.R. Buckley, S.J. [127] Y. Zhou, K. Liu, J.Y. Li, Y.A. Fang, T.C. Zhao, C. Yao, Visualization of nitroxyl in
Butler, Time-resolved luminescence detection of peroxynitrite using a living cells by a chelated copper(II) coumarin complex, Org. Lett. 13 (2011)
reactivity-based lanthanide probe, Chem. Sci. 11 (2020) 3164–3170. 1290–1293.
[101] E. Culotta, D.E. Jr, Koshland, NO news is good news, Science 258 (1992) 1862– [128] Y. Zhou, Y.W. Yao, J.Y. Li, C. Yao, B.P. Lin, Nitroxyl induced fluorescence
1864. enhancement via reduction of a copper(II) coumarin-ester complex: Its
[102] L. Chen, D. Wu, J. Yoon, An ESIPT based fluorescence probe for ratiometric application for bioimaging in vivo, Sensor Actuat. B-Chem. 174 (2012) 414–
monitoring of nitric oxide, Sensor Actuat. B-Chem. 259 (2018) 347–353. 420.
[103] X. Sun, G. Kim, Y. Xu, J. Yoon, T.D. James, A water-soluble copper(II) complex [129] D. Maiti, A.S. Musha Islam, A. Dutta, M. Sasmal, C. Prodhan, M. Ali, Dansyl-
for the selective fluorescence detection of nitric oxide/nitroxyl and imaging appended CuII-complex-based nitroxyl (HNO) sensing with living cell
in living cells, Chem. Plus Chem. 81 (2016) 30–34. imaging application and DFT studies, Dalton Trans. 48 (2019) 2760–2771.
[104] M.H. Lim, S.J. Lippard, Metal-based turn-on fluorescent probes for sensing [130] X. Zhao, C. Gao, N. Li, F. Liu, S. Huo, J. Li, X. Guan, N. Yan, BODIPY based
nitric oxide, Acc. Chem. Res. 40 (2007) 41–51. fluorescent turn-on sensor for highly selective detection of HNO and the
[105] D.A. Jose, N. Sharma, R. Sakla, R. Kaushik, S. Gadiyaram, Fluorescent application in living cells, Tetrahedron Lett. 60 (2019) 1452–1456.
nanoprobes for the sensing of gasotransmitters hydrogen sulfide, H2S), [131] Q. Qi, W. Chi, Y. Li, Q. Qiao, J. Chen, L. Miao, Y. Zhang, J. Li, W. Ji, T. Xu, X. Liu, J.
nitric oxide (NO) and carbon monoxide (CO, Methods 168 (2019) 62–75. Yoon, Z. Xu, A H-Bond strategy to develop acid-resistant photoswitchable
[106] M.D. Pluth, M.R. Chan, L.E. McQuade, S.J. Lippard, Seminaphthofluorescein- rhodamine spirolactams for super-resolution single-molecule localization
based fluorescent probes for imaging nitric oxide in live cells, Inorg. Chem. 50 microscopy, Chem. Sci. 10 (2019) 4914–4922.
(2011) 9385–9392. [132] Y.L. Pak, S.J. Park, Q. Xu, H.M. Kim, J. Yoon, A ratiometric two-photon
[107] X.L. Sun, Y.F. Xu, W.P. Zhu, C.S. He, L. Xu, Y.J. Yang, X. Qian, Copper-promoted fluorescent probe for detecting and imaging hypochlorite, Anal. Chem. 90
probe for nitric oxide based on o-phenylenediamine: Large blue-shift in (2018) 9510–9514.
absorption and fluorescence enhancement, Anal. Methods 4 (2012) 919–922. [133] D. Wu, L. Chen, W. Lee, G. Ko, J. Yin, J. Yoon, Recent progress in the
[108] B. Muthuraj, R. Deshmukh, V. Trivedi, P.K. Iyer, Highly selective probe detects development of organic dye based near-infrared fluorescence probes for
Cu2+ and endogenous NO Gas in living cell, ACS Appl. Mater. Inter. 6 (2014) metal ions, Coord. Chem. Rev. 354 (2018) 74–97.
6562–6569. [134] X. Xiong, F. Song, J. Wang, Y. Zhang, Y. Xue, L. Sun, N. Jiang, P. Gao, L. Tian, X.
[109] B. Mondal, P. Kumar, P. Ghosh, A. Kalita, Fluorescence-based detection of Peng, Thermally activated delayed fluorescence of fluorescein derivative for
nitric oxide in aqueous and methanol media using a copper(II) complex, time-resolved and confocal fluorescence imaging, J. Am. Chem. Soc. 136
Chem. Commun. 47 (2011) 2964–2966. (2014) 9590–9597.
[110] R. Alam, T. Mistri, P. Mondal, D. Das, S.K. Mandal, A.R. Khuda-Bukhsh, M. Ali, A [135] J. Xu, A. Gulzar, P. Yang, H. Bi, D. Yang, S. Gai, F. He, J. Lin, B. Xing, D. Jin, Recent
novel copper(II) complex as a nitric oxide turn-on fluorosensor: Intracellular advances in near-infrared emitting lanthanide-doped nanoconstructs:
applications and DFT calculation, Dalton Trans. 43 (2014) 2566–2576. Mechanism, design and application for bioimaging, Coord. Chem. Rev. 381
[111] J.Y. Kwon, Y.J. Jang, Y.J. Lee, K.-M. Kim, M.-S. Seo, W. Nam, J. Yoon, A highly (2019) 104–134.
selective fluorescent chemosensor for Pb2+, J. Am. Chem. Soc. 127 (2005) [136] Y. Zhang, S. Yuan, G. Day, X. Wang, X. Yang, H.-C. Zhou, Luminescent sensors
10107–10111. based on metal-organic frameworks, Coord. Chem. Rev. 354 (2018) 28–45.

26

You might also like