Organic & Biomolecular Chemistry

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Organic &

Biomolecular Chemistry
View Article Online
PAPER View Journal | View Issue
Published on 01 March 2019. Downloaded by Southeast University - Jiulonghu Campus on 1/21/2024 8:51:12 AM.

A mitochondria-targeted red-emitting probe for


Cite this: Org. Biomol. Chem., 2019,
imaging hydrogen sulfide in living cells and
17, 3389 zebrafish†
Ismail Ismail, ‡a Dan Wang,‡a Dawei Wang,a Cuili Niu,a Haojie Huang,b
Long Yi *b,c and Zhen Xi*a,c

Hydrogen sulfide (H2S) is an important endogenous signaling molecule with multiple biological functions.
Based on the thiolysis of NBD (7-nitro-1,2,3-benzoxadiazole) amine, herein we rationally design and syn-
thesize a new fluorescent probe 1 for the detection of mitochondrial H2S in living cells. Probe 1 displays
Received 30th December 2018, bright red-emitting fluorescence after H2S activation (ϕ, 0.77). 1 shows higher selectivity and sensitivity for
Accepted 1st March 2019
H2S than the red-emitting probe Rho-NBD (ChemBioChem, 2016, 17, 962). Moreover, 1 is water-soluble,
DOI: 10.1039/c8ob03219j cell-membrane-permeable, of low cytotoxicity and mitochondria-targeting, and can be employed for
rsc.li/obc monitoring mitochondrial H2S in living cells.

Introduction nate the reactivity of millimolar biothiols and micromolar H2S.


To address this challenge, H2S-triggered chemical reac-
Hydrogen sulfide (H2S) is an important endogenous signaling tions,13,14 including nucleophilic addition,15–19 reduction of
molecule with multiple biological functions.1–6 H2S could be azide or nitro to amine,20–25 copper precipitation,26 thiolysis of
enzymatically produced in vivo by three distinctive pathways dinitrophenyl ether,27,28 thiolysis of NBD dyes,29–43 and clea-
including cystathionine β-synthase (CBS), cystathionine γ-lyase vage of CvC bonds,44 have been successfully employed to
(CSE) and 3-mercaptopyruvate sulfur transferase (3-MPST)/ develop fluorescent and/or colorimetric probes for the detec-
cysteine aminotransferase (CAT) in different organs and tion of biological H2S. Among these, the thiolysis of NBD
tissues.7–10 Recent studies have shown that the H2S level amine is a successful reaction for the development of H2S-
in vivo is correlated with numerous diseases, including the specific fluorescent probes.29,30 In such a probe, the fluo-
symptoms of Alzheimer’s disease, Down’s syndrome, diabetes rescence of the fluorophore can be quenched by the NBD
and liver cirrhosis.1,11,12 Although H2S has been recognized to moiety in aqueous solution, which will be released after the
be linked to numerous physiological and pathological pro- thiolysis reaction (Scheme 1). To this end, a series of NBD-
cesses, many of its underlying molecular events in vivo remain based H2S probes were developed by our groups and
largely unknown and could be further explored. Therefore, others,31–39,45 which showed emissions from the blue to near-
developing efficient methods for in situ and real-time detection infrared range (Table S1†).
of biological H2S is of significant research value.
One major challenge in the development of H2S probes is
the development of a chemical reaction to effectively discrimi-

a
State Key Laboratory of Elemento-Organic Chemistry and Department of Chemical
Biology, National Engineering Research Center of Pesticide (Tianjin),
College of Chemistry, Nankai University, Tianjin 300071, China.
E-mail: zhenxi@nankai.edu.cn
b
State Key Laboratory of Organic-Inorganic Composites and Beijing Key Laboratory of
Bioprocess, Beijing University of Chemical Technology (BUCT), Beijing 100029,
China. E-mail: yilong@mail.buct.edu.cn
c
Collaborative Innovation Center of Chemical Science and Engineering (Tianjin),
Tianjin 300071, China
† Electronic supplementary information (ESI) available. See DOI: 10.1039/
c8ob03219j Scheme 1 (A) Thiolysis of NBD amine for fluorescence sensing of H2S.
‡ These authors contributed equally to this work. (B) Chemical structures of three NBD-based H2S probes.

This journal is © The Royal Society of Chemistry 2019 Org. Biomol. Chem., 2019, 17, 3389–3395 | 3389
View Article Online

Paper Organic & Biomolecular Chemistry

On the other hand, Grimm et al. showed that replacing the


N,N-dialkyl group with the azetidine group can greatly improve
the brightness and quantum yield of the fluorophores.46 Yang
et al. reported that blocking the twisting N,N-diethylamino
group of coumarin can improve the fluorescence properties of
the fluorophore and, in turn, enhance the sensing perform-
Published on 01 March 2019. Downloaded by Southeast University - Jiulonghu Campus on 1/21/2024 8:51:12 AM.

ance.47,48 We find that compared with Cou-NBD, the probe


Cyclo-Cou-NBD showed significantly improved performance
with respect to response time, detection limit and fluorescence
enhancement.49 In this work, we envision that rhodamine
with the twisting-blockage of the N,N-diethylamino group can
be used to develop H2S probes with better properties than
Rho-NBD, which showed only 4.5-fold turn-on fluorescence
during H2S activation.33 Herein, we rationally design and syn-
thesize a new azetidinyl-rhodamine-NBD dyad 1 for fast
sensing of H2S in the red light range. To our delight, 1 displays
19-fold fluorescence enhancement in the presence of H2S. Fig. 2 (A) Time-dependent absorbance spectra of 1 (5 µM) and Na2S
Moreover, 1 can be employed for imaging biological H2S in (100 µM). (B) Time-dependent fluorescence spectra (excitation =
living cells and zebrafish. 565 nm) of 1 (5 µM) and Na2S (100 µM). (C) The time-dependent fluor-
escence intensities at 585 nm for 1 (5 µM) in the presence of different
concentrations of H2S. The solid lines represent the best single-expo-
nential fitting. (D) The linear relationship of kobs versus H2S
concentrations.
Results and discussion
Azetidinyl-rhodamine 3 was prepared from fluorescein in a
straightforward way with good yields (Fig. 1). Probe 1 was pre- HPLC studies (Fig. S1†) further suggested the fast conversion
pared by the coupling of 3 and NBD-piperazine. The structure of 1 into 4 in the presence of H2S.
of 1 was confirmed by 1H NMR, 13C NMR, and high-resolution As shown in Fig. 2, we tested the fluorescence spectra of the
mass spectroscopy (HRMS). Such a facile and economic syn- reaction between 1 and H2S. Probe 1 showed weak fluo-
thesis is important for the wide use of the probe. rescence (quantum yield ϕ, 0.17) upon excitation at 565 nm,
With the probe in hand, we firstly tested the absorbance implying that the fluorescence in 1 is efficiently quenched by
and fluorescence spectra of the reaction between 1 and H2S NBD via the PET effect. After the reaction with H2S, a large
(using Na2S as an equivalent) in PBS buffer (20 mM, pH 7.4). fluorescence change for 1 was observed, with about 19-fold
As shown in Fig. 2, 1 exhibited absorbance peaks at 565 nm intensity increase at 585 nm. The quantum yield of 4 was
and 530 nm, which should be assigned to the rhodamine and determined as ϕ = 0.77. The H2S-triggered fluorescence
NBD fluorophores, respectively. After the reaction with H2S, enhancement folds of 1 are improved significantly compared
the absorbance peaks appeared at 565 nm and 500 nm, which with those of Rho-NBD (4.5 folds),33 implying that the twist-
might be due to the production of 4 and NBD-SH, respectively. ing-blockage of the N,N-diethylamino group in rhodamine B
Such a thiolysis reaction was also characterized by HRMS for can significantly improve the performance of azetidinyl-rhoda-
the peak at 479.2446 (the calculated value for [4]+: 479.2442). mine-NBD probe 1. These results encouraged us to further
check the sensitivity and selectivity of probe 1.
The reaction kinetics of the probe with H2S is an important
parameter to examine its biological applicability on account of
the rapid catabolism of H2S under physiological conditions.
To this end, time-dependent fluorescence intensities at
585 nm for the probe in the presence of H2S were acquired for
data analysis (Fig. 2c). The pseudo-first-order rate, kobs, was
obtained by fitting the data with a single exponential function.
The reaction rate k2 (27.8 M−1 s−1) was obtained by the linear
fitting of kobs versus H2S concentrations (Fig. 2d). The H2S-reac-
tion rate of 1 is about one order of magnitude faster than that
of the o-fluorinated-azido-based probe,8 implying that such
Fig. 1 Synthetic route to probe 1 and its reaction with H2S. Reagents
NBD-based probes can be employed for the fast detection of
and conditions: (a) Trifluoromethanesulfonic anhydride, pyridine,
CH2Cl2, 0 °C to room temperature, 83%; (b) Cs2CO3, Pd2dba3, XPhos,
H2S.
azitidine, dioxane, 95%; and (c) NBD-piperazine, HATU, DIPEA, DMF, Because the biological H2S level normally exists in the
59%. range of nanomolar to micromolar concentrations,1–3 we pro-

3390 | Org. Biomol. Chem., 2019, 17, 3389–3395 This journal is © The Royal Society of Chemistry 2019
View Article Online

Organic & Biomolecular Chemistry Paper

molecule was nearly negligible except for H2S (Fig. 4), implying
that 1 can be used to selectively detect H2S in biological
samples. Additionally, we also investigated the fluorescence
response of 1 to H2S at different pH values. The results indi-
cated that the probe can be functionalized over pH ranging
from 7.0 to 9.0 (Fig. S3†).
Published on 01 March 2019. Downloaded by Southeast University - Jiulonghu Campus on 1/21/2024 8:51:12 AM.

Encouraged by the above results, we proceeded to study the


biological applications of 1. Firstly, the cytotoxicity of the
probe was evaluated using HEK293 cells by the detection of
adherent cell proliferation by varying the electrical current
(Fig. 5, S4†).50 The probe did not show obvious cytotoxicity in
the 0–10 µM range upon 24 h of incubation, implying that the
Fig. 3 The fluorescence titration of probe 1 (5 µM) toward H2S probe is biocompatible at these concentrations. An obvious
(0–200 µM). The solid line represents the best fitting using one-site inhibition of cell proliferation was observed at 20 µM 1, and
binding of pharmacology in Origin software. therefore, 10 µM 1 was used in the following in vivo imaging.
We further examined whether the probe can be used to
detect H2S in living cells. HEK293A cells were treated with
ceeded to study the fluorescence response of 1 upon increasing probe 1 and then washed with PBS to remove excess 1. The 1-
the concentrations of H2S in the micromolar range (Fig. 3, loaded cells were incubated with H2S and subsequently
S2†). The fluorescence intensity was linearly related to imaged using confocal fluorescence microscopy (Fig. 6). The
1–20 μM H2S. Specifically, the detection limit of 1 was deter- addition of both 1 and H2S resulted in obvious red fluo-
mined to be 0.36 μM based on the 3σ/slope method, implying rescence while the cells treated with only probe 1 did not show
that the probe is highly sensitive toward H2S. The complete obvious fluorescence. Bright-field images show that cells
titration resulted in a formation constant of 7.1 × 104 M−1 for retained good morphology after incubation with the probe,
H2S, implying a relatively good affinity of 1. which further suggested the good biocompatibility of 1. To test
A major challenge for H2S detection in biological systems is whether 1 could detect the endogenous production of H2S, the
to develop a highly selective probe that exhibits notably dis- cells were co-incubated with L-Cys and 1. After using L-Cys to
tinctive response to micromolar H2S over millimolar biothiols induce endogenous H2S production, significant fluorescence
and other cellular molecules. To this end, 1 was incubated enhancement could be observed (Fig. 6C). These preliminary
with various biologically relevant species including reactive studies suggested that probe 1 could be used for the visualiza-
sulfur species (biothiols, SO32−), reactive oxygen species (H2O2, tion of H2S in cells efficiently and selectively.
NaOCl, •OH and ONOO−), NaNO2, and ions. These results indi- Since the probe could react quickly with H2S, we also tested
cated that fluorescence intensity enhancement for any tested the time-dependent L-Cys-induced H2S biogenesis in living
cells. The average fluorescence of the time-dependent images
(Fig. S5†) implies that H2S biogenesis from L-Cys occurs con-
tinuously at a relatively stable rate. Taken together, these
results indicate that probe 1 can be used for bioimaging
endogenous H2S in living cells.
Furthermore, the probe contains a positive charge, which
should be mitochondria-targeting. A fluorescence co-localiz-

Fig. 4 Fluorescence intensity at 585 nm of probe 1 (5 µM) in the pres-


ence of various species (100 µM). Lane 1, GSH; 2, Hcy; 3, CH3COOH; 4,
Cys; 5, H2O2; 6, K2C2O4; 7, KF; 8, mercaptoethanol; 9, Na2SO3; 10,
NaHCO3; 11, NaI; 12, NaNO2; 13, NaOCl; 14, •OH; 15, ONOO−; 16, Na2S Fig. 5 Concentration-dependent maximum cell index in the presence
in PBS ( pH 7.4) at room temperature for 30 min. of probe 1 (0–20 µM) for 24 h.

This journal is © The Royal Society of Chemistry 2019 Org. Biomol. Chem., 2019, 17, 3389–3395 | 3391
View Article Online

Paper Organic & Biomolecular Chemistry


Published on 01 March 2019. Downloaded by Southeast University - Jiulonghu Campus on 1/21/2024 8:51:12 AM.

Fig. 6 Fluorescence images for H2S detection in living cells using 1.


Fluorescence images of cells upon incubation with 1 (10 µM) for 30 min
(A); with 1 and Na2S (100 µM) (B); with 1 and L-Cys (100 µM) (C); (D–F) a
merged image of the bright-field and fluorescence images for (A–C),
respectively. Scale bars: 50 μm. Fig. 8 Confocal microscopy images of exogenous and L-Cys-induced
endogenous H2S detection in zebrafish larvae. (A) The zebrafish was
incubated with probe 1 (10 μM) for 1 h and imaged in the red field (left)
and bright field (right). (B) The zebrafish was incubated with probe 1
ation assay with Mito-Tracker Green FM (a well-known mito- (10 μM) for 1 h, then with Na2S (200 μM) for 3 h and imaged in the red
chondria-specific dye) and probe 1 with L-Cys was performed. field (left) and bright field (right). (C) The zebrafish was incubated with
As shown in Fig. 7, the images in the red channel and green probe 1 (10 μM) for 1 h, then with L-Cys (200 μM) for another 3 h, and
channel merged very well, implying that probe 1 was well co- imaged in the red field (left) and bright field (right).

localized with Mito-Tracker Green FM (Pearson’s coefficient


0.947) (Fig. 7D). The above data clarified that probe 1 is an
excellent tool for imaging mitochondrial H2S. for another 3 h. The zebrafish stained with 1 alone was used as
To further demonstrate the biological applicability of the the control group. No obvious fluorescence was observed in
NBD-based probe, we examined whether it can be used to the 1-loaded control group (Fig. 8A), whereas strong red fluo-
detect H2S in living animals, and zebrafish larvae were selected rescence was seen in the H2S-treated zebrafish (Fig. 8B). A sig-
as the biological model. Firstly, the zebrafish was incubated nificant red fluorescence enhancement was also observed in
with 1 for 1 h, washed, and then treated with Na2S (200 μM) the Cys-treated zebrafish (Fig. 8C). These results imply that
probe 1 can be used to visualize H2S in vivo.

Conclusions
A new NBD-based fluorescent turn-on probe is developed for
H2S detection in aqueous buffer and in living cells. The probe
is based on the fast and selective thiolysis of NBD amine,
which can be a general design strategy for H2S probes.52 Probe
1 shows 19-fold turn-on fluorescence at 585 nm upon H2S acti-
vation. 1 contains a positive charge, which is useful for fast
sensing of H2S (k2 = 27.8 M−1 s−1) and mitochondria-targeting.
1 exhibits excellent selectivity toward H2S over millimolar
biothiols in physiological buffer. Moreover, 1 is successfully
applied in the bioimaging of H2S in living cells and zebrafish.
This probe could be a useful tool for potential applications in
H2S biology.

Experimental
Fig. 7 Confocal microscopy images of the mitochondrion in living
cells. Cells were co-stained with 1 (10 μM), L-Cys (100 μM) and commer- Materials and methods
cial MitoTracker® Green FM (2.5 μM). (A) Red channel (555–655 nm,
All chemicals and solvents used for the synthesis were pur-
excitation at 543 nm); (B) green channel (500–530 nm, excitation at
488 nm); (C) a merged graph from fluorescence images of 1 and
chased from commercial suppliers and applied directly in the
MitoTracker® Green FM. Scale bars: 50 μm. (D) Intensity correlation plots experiments without further purification. The progress of the
of the green channel and the red channel. Pearson’s coefficient: 0.947. reaction was monitored by TLC on pre-coated silica plates

3392 | Org. Biomol. Chem., 2019, 17, 3389–3395 This journal is © The Royal Society of Chemistry 2019
View Article Online

Organic & Biomolecular Chemistry Paper

(Merck 60F-254, 250 µm in thickness), and spots were visual- δ 8.00 (d, J = 7.6 Hz, 1H), 7.64 (dt, J = 22.4, 7.2 Hz, 2H), 7.15 (d,
ized by basic KMnO4, UV light or iodine. Merck silica gel 60 J = 7.2 Hz, 1H), 6.60 (d, J = 8.4 Hz, 2H), 6.22 (s, 2H), 6.14 (d, J =
(100–200 mesh) was used for general column chromatography 8.4 Hz, 2H), 3.94 (t, J = 7.2 Hz, 8H), 2.48–2.31 (m, 4H); 13C
purification. 1H NMR and 13C NMR spectra were recorded on a NMR (101 MHz, CD2Cl2) δ 169.62, 154.26, 153.33, 152.49,
Bruker 400 spectrometer. Chemical shifts are reported in parts 134.87, 129.77, 129.16, 128.18, 125.34, 124.61, 52.50, 17.01.
per million with respect to the internal standard tetramethyl- HRMS (ESI): m/z [M + H]+ calcd for C26H22N2O3: 410.1630,
Published on 01 March 2019. Downloaded by Southeast University - Jiulonghu Campus on 1/21/2024 8:51:12 AM.

silane (Si(CH3)4 = 0.00 ppm) or residual solvent peaks (CD2Cl2 found: 411.1706.
= 5.32 ppm; CDCl3 = 7.26 ppm; DMSO-d6 = 2.5 ppm). 1H NMR
coupling constants ( J) are reported in hertz (Hz), and multi- Azetidinyl-rhodamine-NBD dyad 1
plicity is indicated as the following: s (singlet), d (doublet), 2-(3-(Azetidin-1-ium-1-ylidene)-6-(azetidin-1-yl)-3H-xanthen-9-yl)
t (triplet), dd (doublet of doublets), m (multiple). High-resolu- benzoate (0.134 g, 0.33 mmol) was dissolved in 8 ml DMF, fol-
tion mass spectra (HRMS) were obtained on an Agilent lowed by the addition of HATU (0.144 g, 0.38 mmol) and
6540 UHD Accurate-Mass Q-TOF LC/MS or Varian 7.0 T DIPEA (148 μl, 0.75 mmol). The solution was stirred for 5 min
FTICR-MS. The UV-visible spectra were recorded on a UV-3600 and NBD-piperazine33,51 (0.068 g, 2.83 mmol) was added to
UV-VIS-NIR spectrophotometer (Shimadzu, Japan). Fluorescence the solution; stirring was continued for 12 h at room tempera-
study was carried out using an F-280 spectrophotometer ture. After completion of the reaction DMF was removed
(Tianjin Gangdong Sci & Tech., Development. Co., Ltd). in vacuo. The residue was purified by silica gel column chrom-
atography to yield dark-red solid 1 (100 mg, 59%). Rf = 0.7
3-Oxo-3H-spiro[isobenzofuran-1,9′-xanthene]-3′,6′-diyl bis (CH3OH/CH2Cl2/NH3 = 2/10/0.5). 1H NMR (400 MHz, DMSO-
(trifluoromethanesulfonate) 2 d6) δ 8.53 (d, J = 9.0 Hz, 1H), 7.77 (s, 3H), 7.53 (s, 1H), 7.14 (d,
A two-necked round-bottom flask was charged with fluorescein J = 9.2 Hz, 2H), 6.64 (d, J = 8.8 Hz, 2H), 6.59–6.46 (m, 3H), 4.25
sodium (5 g, 15.04 mmol) in anhydrous CH2Cl2 (60 mL) and (t, J = 7.2 Hz, 8H), 4.01 (s, 4H), 3.71 (s, 2H), 3.54 (s, 2H),
cooled to 0 °C. Pyridine (9.74 ml, 12.04 mmol, 8.0 eq.) and tri- 2.48–2.36 (m, 4H). 13C NMR (101 MHz, DMSO-d6) δ 166.80,
fluoromethanesulfonic anhydride (10.12 ml, 60.2 mmol, 156.46, 155.92, 155.69, 145.26, 144.73, 136.29, 135.05, 131.71,
4.0 eq.) were added slowly into the mixture under an inert 130.82, 130.37, 129.94, 127.54, 121.49, 113.14, 112.47, 103.38,
atmosphere. The reaction mixture was stirred at room tempera- 93.97, 51.79, 15.41. HRMS (ESI): calcd for C36H32N7O5+:
ture for 4 h. The reaction was quenched with water, and the 642.2459, found m/z [M]+: 642.2462.
reaction mixture extracted with CH2Cl2, washed with saturated
brine, and dried over anhydrous sodium sulfate. After the Solution preparation method for spectroscopic studies
solvent was removed in vacuo, recrystallization from ethyl All spectroscopic measurements were performed in PBS buffer
acetate and hexanes afforded the pure product as an off-white (20 mM, pH 7.4) in a 3 ml cuvette with 2 ml solution. 1 was
solid (7.4 g) with 83% yield. Rf = 0.7 (ethyl acetate/hexanes = 3/ dissolved in DMSO to prepare the stock solutions with a con-
10); 1H NMR (400 MHz, DMSO-d6) δ 8.09 (d, J = 7.2 Hz, 1H), centration of 5 mM. 1–1000 mM stock solutions of Na2S in
7.87–7.76 (m, 4H), 7.50 (d, J = 7.6 Hz, 1H), 7.32 (dd, J = 8.8, 2.4 degassed PBS buffer were used as the H2S source. The probe
Hz, 2H), 7.12 (d, J = 8.8 Hz, 2H); 13C NMR (101 MHz, DMSO- was diluted in PBS buffer to achieve the final concentration of
d6) δ 168.12, 151.85, 150.74, 149.82, 136.24, 130.89, 130.62, 2–5 µM. For the selectivity tests, different biologically relevant
125.28, 125.02, 124.20, 119.80, 119.38, 118.01, 116.61, 110.84, species (100 mM) were prepared as stock solutions in PBS
79.66; HRMS (ESI): m/z [M]+ calcd for C22H10F6O9S2: 595.9670, buffer or DMSO. The •OH was generated by the Fenton reac-
found: 596.9746. tion in which ferrous chloride was added in the presence of
10 equiv. of H2O2. The ONOO− was generated by the reaction
2-(3-(Azetidin-1-ium-1-ylidene)-6-(azetidin-1-yl)-3H-xanthen- of nitrous acid with hydrogen peroxide in acidic solution
9-yl)benzoate 3 (the concentration was determined by UV spectrophotometry
A 25 ml two-necked round bottom flask was charged with at 302 nm).53 All species were incubated at room temperature
fluorescein ditriflate (0.15 g, 25.2 mmol), 2-dicyclohexyl- for 30 min and then the spectra were recorded under excitation
phosphino-2′,4′,6′-triisopropylbiphenyl (Xphos, 0.036 g, at 565 nm.
75.4 μmol. 0.3 eq.), Cs2CO3 (0.23 g, 704 μmol, 2.8 eq.) and tris
(dibenzylideneacetone)dipalladium(0) (Pd2dba3, 0.023 g, Cell culture
25.2 mmol. 0.1 eq.) in dioxane (2 mL). Argon was fluxed before Complete growth medium was used to culture human embryo-
adding azetidine (40.6 μL, 604 μmol, 2.4 eq.), and the reaction nic kidney HEK-293A cells, which contained high glucose
mixture was stirred at 100 °C under an argon atmosphere for DMEM (Gibco) supplemented with FBS (10%), penicillin (100
18 h. After completion the reaction system was cooled to room U ml−1), streptomycin (10 mg ml−1), and L-glutamine (4 mM).
temperature, diluted with CH3OH, filtered with Celite, and The cells were maintained in exponential growth, and then
concentrated in vacuo. Purification by silica gel chromato- seeded in a 35 mm glass-bottom plate. The cells were passaged
graphy using a solvent system (0–7% CH3OH (2 M NH3)/ every 2–3 days and used between passages 3 and 10.
CH2CI2) afforded a purple solid (0.098 g, 95%). Rf = 0.6 Cytotoxicity assay. The cytotoxicity of 1 was determined
(CH3OH/CH2Cl2/NH3 = 2/10/0.5); 1H NMR (400 MHz, CD2Cl2) using a xCELLigence RTCA S16 (ACEA Biosciences),50 which is

This journal is © The Royal Society of Chemistry 2019 Org. Biomol. Chem., 2019, 17, 3389–3395 | 3393
View Article Online

Paper Organic & Biomolecular Chemistry

an instrument that uses microelectronic plates integrated with Notes and references
gold microelectrode arrays on a glass substrate in the bottom
of the wells to measure the proliferation of adherent cells. 1 C. Szabó, Nat. Rev. Drug Discovery, 2007, 6, 917.
When cells replicate and grow, they can cover the electrosensi- 2 L. Li, P. Rose and P. K. Moore, Annu. Rev. Pharmacol.
tive elements and cause higher impedance, and the electronic Toxicol., 2011, 51, 169.
current will be changed. The software connected to a computer 3 B. L. Predmore, D. J. Lefer and G. Gojon, Antioxid. Redox
Published on 01 March 2019. Downloaded by Southeast University - Jiulonghu Campus on 1/21/2024 8:51:12 AM.

can monitor these changes in real time, reflecting the prolifer- Signaling, 2012, 17, 119.
ation of adherent cells. Briefly, HEK293A cells growing in the 4 G. K. Kolluru, X. G. Shen, S. C. Bir and C. G. Kevil, Nitric
log phase were seeded into a 16-well cell-culture plate (100 µL Oxide, 2013, 35, 5.
per well). When the amount of the cells reached 5 × 104 cells 5 H. Kimura, Exp. Physiol., 2011, 96, 833.
per well, probe 1 was added into each well with final concen- 6 G. Yang, L. Wu, B. Jiang, W. Yang, J. Qi, K. Cao, Q. Meng,
trations of: 0, 1, 2, 5, 10, 15 and 20 µM. These cells were incu- A. K. Mustafa, W. Mu, S. Zhang, S. H. Snyder and R. Wang,
bated at 37 °C under 5% CO2, and the data were collected Science, 2008, 322, 587.
every 15 min for the duration of the assay. 7 H. Kimura, Amino Acids, 2011, 41, 113.
Cell bioimaging. An FV1000 inverted fluorescence confocal 8 L. Wei, Z. T. Zhu, Y. Y. Li, L. Yi and Z. Xi, Chem. Commun.,
microscope (Olympus, Japan) with a 40× objective lens was 2015, 51, 10463.
used for imaging. For exogenous H2S, HEK293A cells were 9 N. Shibuya, S. Koike, M. Tanaka, M. Ishigami-Yuasa,
incubated with the probe (2, 5 or 10 μM) for 30 min, washed, Y. Kimura, Y. Ogasawara, K. Fukui, N. Nagahara and
and then incubated with Na2S (100 μM) for another 30 min. H. Kimura, Nat. Commun., 2013, 4, 1366.
The medium was washed and replaced by PBS, and the cells 10 L. Wei, L. Yi, F. B. Song, C. Wei, B. F. Wang and Z. Xi, Sci.
were imaged via the red channels (Ex. = 543 nm, Em. = Rep., 2014, 4, 4521.
555–655 nm). The control cells were treated with only the 11 S. Fiorucci, E. Antonelli, A. Mencarelli, S. Orlandi,
probe (10 μM) for 30 min. The HEK293A cells were incubated B. Renga, G. Rizzo, E. Distrutti, V. Shah and A. Morelli,
with 100 μM L-Cys for 30 min to induce endogenous H2S pro- Hepatology, 2005, 42, 539.
duction, washed with PBS, and then incubated with the probe 12 S. K. Bae, C. H. Heo, D. J. Choi, D. Sen, E.-H. Joe, B. R. Cho
(5 or 10 μM) for another 30 min. The medium was washed and and H. M. Kim, J. Am. Chem. Soc., 2013, 135, 9915.
replaced by PBS, and the cells were imaged via the red chan- 13 V. S. Lin, W. Chen, M. Xian and C. J. Chang, Chem. Soc.
nels. The imaging results were analyzed using Olympus Rev., 2015, 44, 4596.
FV1000-ASW software. 14 F. B. Yu, X. Y. Han and L. X. Chen, Chem. Commun., 2014,
Co-localization. Living HEK293 cells were co-incubated with 50, 12234.
L-Cys (100 μM) and the probe (10 μM) for 30 min and washed 15 Y. Qian, J. Karpus, O. Kabil, S. Y. Zhang, H. L. Zhu,
with PBS; then Mito-Tracker Green FM (2.5 μM) was added for R. Banerjee, J. Zhao and C. He, Nat. Commun., 2011, 2, 495.
another 30 min. The images were collected from band pass 16 L. W. He, W. Y. Lin, Q. Y. Xu and H. P. Wei, Chem.
555–655 nm upon excitation at 543 nm, from band pass Commun., 2015, 51, 1510.
500–530 nm upon excitation at 488 nm, and bright-field. 17 Y. C. Chen, C. C. Zhu, Z. H. Yang, J. J. Chen, Y. F. He,
In vivo imaging in zebrafish. Imaging in zebrafish was per- Y. Jiao, W. J. He, L. Qiu, J. J. Cen and Z. J. Guo, Angew.
formed according to the institutional ethical guidelines for Chem., Int. Ed., 2013, 52, 1688.
animal experiment. The accreditation number of the labora- 18 Y. Qian, L. Zhang, S. T. Ding, X. Deng, C. He, X. E. Zheng,
tory is SYXK (Jin) 2014-00003 promulgated by Tianjin Science H. L. Zhu and J. Zhao, Chem. Sci., 2012, 3, 2920.
and Technology Commission. 7-day-old zebrafish was incu- 19 X. Wang, J. Sun, W. Zhang, X. Ma, J. Lv and B. Tang, Chem.
bated with the probe (10 μM) for 1 h, washed with culture Sci., 2013, 4, 2551.
medium, then incubated with Na2S (200 μM) or L-Cys (200 μM) 20 A. R. Lippert, E. J. New and C. J. Chang, J. Am. Chem. Soc.,
for 3 h, and imaged with an Olympus fluorescence microscope 2011, 133, 10078.
and imaged in the red channel. 21 H. J. Peng, Y. F. Cheng, C. F. Dai, A. L. King,
B. L. Predmore, D. J. Lefer and B. H. Wang, Angew. Chem.,
Int. Ed., 2011, 50, 9672.
Conflicts of interest 22 Q. Q. Wan, Y. C. Song, Z. Li, X. H. Gao and H. M. Ma,
Chem. Commun., 2013, 49, 502.
There are no conflicts to declare. 23 M. D. Hartle and M. D. Pluth, Chem. Soc. Rev., 2016, 45, 6108.
24 L. Zhang, W. Q. Meng, L. Lu, Y. S. Xue, C. Li, F. Zou, Y. Liu
and J. Zhao, Sci. Rep., 2014, 4, 5870.
Acknowledgements 25 H. A. Henthorn and M. D. Pluth, J. Am. Chem. Soc., 2015,
137, 15330.
This work was supported by the National Key R&D Program of 26 K. Sasakura, K. Hanaoka, N. Shibuya, Y. Mikami,
China (2017YFD0200500, 16JCYBJC20200) and the NSFC Y. Kimura, T. Komatsu, T. Ueno, T. Terai, H. Kimura and
(21702111, 21572019, 21877008). T. Nagano, J. Am. Chem. Soc., 2011, 133, 18003.

3394 | Org. Biomol. Chem., 2019, 17, 3389–3395 This journal is © The Royal Society of Chemistry 2019
View Article Online

Organic & Biomolecular Chemistry Paper

27 X. W. Cao, W. Y. Lin, K. B. Zheng and L. W. He, Chem. 41 M. D. Hammers and M. D. Pluth, Anal. Chem., 2014, 86,
Commun., 2012, 48, 10529. 7135.
28 Z. J. Huang, S. S. Ding, D. H. Yu, F. H. Huang and 42 G. D. Zhou, H. L. Wang, Y. Ma and X. Q. Chen, Tetrahedron,
G. Q. Feng, Chem. Commun., 2014, 50, 9185. 2013, 69, 867.
29 C. Wei, L. Wei, Z. Xi and L. Yi, Tetrahedron Lett., 2013, 54, 43 D. Jiménez, R. Martínez-Máñez, F. Sancenón, J. V. Ros-Lis,
6937. A. Benito and J. Soto, J. Am. Chem. Soc., 2003, 125,
Published on 01 March 2019. Downloaded by Southeast University - Jiulonghu Campus on 1/21/2024 8:51:12 AM.

30 L. A. Montoya, T. F. Pearce, R. J. Hansen, L. N. Zakharov 9000.


and M. D. Pluth, J. Org. Chem., 2013, 78, 6550. 44 C. Wang, X. Cheng, J. Tan, Z. Ding, W. Wang, D. Yuan,
31 C. Y. Zhang, L. Wei, J. Zhang, R. Y. Wang, Z. Xi and L. Yi, G. Li, H. Zhang and X. Zhan, Chem. Sci., 2018, 9,
Chem. Commun., 2015, 51, 7505. 8369.
32 C. Wei, R. Y. Wang, C. Y. Zhang, G. C. Xu, Y. Y. Li, 45 L. Yi and Z. Xi, Org. Biomol. Chem., 2017, 15, 3828.
Q. Z. Zhang, L. Y. Li, L. Yi and Z. Xi, Chem. – Asian J., 2016, 46 J. B. Grimm, B. P. English, J. Chen, J. P. Slaughter,
11, 1376. Z. Zhang, A. Revyakin, R. Patel, J. J. Macklin, D. Normanno,
33 R. Y. Wang, Z. F. Li, C. Y. Zhang, Y. Y. Li, G. C. Xu, R. H. Singer, T. Lionnet and L. D. Lavis, Nat. Methods,
Q. Z. Zhang, L. Y. Li, L. Yi and Z. Xi, ChemBioChem, 2016, 2015, 12, 244.
17, 962. 47 Q. Sun, S. H. Yang, L. Wu, W. C. Yang and G. F. Yang, Anal.
34 J. Zhang, R. Y. Wang, Z. T. Zhu, L. Yi and Z. Xi, Tetrahedron, Chem., 2016, 88, 2266.
2015, 71, 8572. 48 J. Li, C. F. Zhang, S. H. Yang, W. C. Yang and G. F. Yang,
35 C. Y. Zhang, R. Y. Wang, L. H. Cheng, B. J. Li, Z. Xi and Anal. Chem., 2014, 86, 3037.
L. Yi, Sci. Rep., 2016, 6, 30148. 49 I. Ismail, D. Wang, D. Wang, Z. Wang, C. Zhang, L. Yi and
36 K. Zhang, J. Zhang, Z. Xi, L.-Y. Li, X. Gu, Q.-Z. Zhang and Z. Xi, Dyes Pigm., 2019, 163, 706.
L. Yi, Chem. Sci., 2017, 8, 2776. 50 Y. A. Abassi, B. Xi, W. F. Zhang, P. F. Ye, S. Li, L. Kirstein,
37 Y. F. Zhang, H. Y. Chen, D. Chen, D. Wu, X. Q. Chen, M. R. Gaylord, S. C. Feinstein, X. B. Wang and X. Xu, Chem.
S. H. Liu and J. Yin, Org. Biomol. Chem., 2015, 13, 9760. Biol., 2009, 16, 712.
38 Y. L. Pak, J. Li, K. C. Ko, G. Kim, J. Y. Lee and J. Yoon, Anal. 51 F. Song, Z. Li, J. Li, S. Wu, X. Qiu, Z. Xi and L. Yi, Org.
Chem., 2016, 88, 5476. Biomol. Chem., 2016, 14, 11117.
39 B. Roubinet, L. Bailly, E. Petit, P.-Y. Renard and A. Romieu, 52 C. Zhang, Q.-Z. Zhang, K. Zhang, L.-Y. Li, M. D. Pluth, L. Yi
Tetrahedron Lett., 2015, 56, 1015. and Z. Xi, Chem. Sci., 2019, 10, 1945.
40 C. Wei, Q. Zhu, W. W. Liu, W. B. Chen, Z. Xi and L. Yi, Org. 53 J. W. Reed, H. H. Ho and W. L. Jolly, J. Am. Chem. Soc.,
Biomol. Chem., 2014, 12, 479. 1974, 96, 1249.

This journal is © The Royal Society of Chemistry 2019 Org. Biomol. Chem., 2019, 17, 3389–3395 | 3395

You might also like