You are on page 1of 6

pubs.acs.

org/ac Article

Intratumoral Glutathione Activatable Nanoprobes for Fluorescence


and 19F Magnetic Resonance Turn-On Imaging
Yangyang Zhang, Qian Ma, Yunhe Yan, Chang Guo, Suying Xu,* and Leyu Wang*
Cite This: Anal. Chem. 2020, 92, 15679−15684 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Tumor microenvironment turn-on nanoprobes that


could specifically detect the occurrence of diseases possess great
potential in early diagnosis. Here, a GSH activated nanoprobe was
Downloaded via POLITECHNIKA SLASKA on March 10, 2021 at 09:14:06 (UTC).

designed for fluorescence and 19F magnetic resonance (MR) dual-


modal turn-on imaging of tumors. Specifically, fluorescence AgInS2
quantum dots (QDs for fluorescence imaging) were co-
encapsulated with perfluoro-15-crown-5-ether (P19FCE for19F
MRI) by amphiphilic polymers and further coated with in situ
formed manganese dioxide (MnO2) nanoshells, which served as
efficient fluorescence and 19F MR quenchers due to energy transfer
and paramagnetic relaxation effects, respectively. The over-
expressed GSH in tumors would decompose the MnO2 nanoshells,
resulting in remarkable enhancement of both fluorescence and 19F
MRI signals of the nanoprobes, accordingly lighting up the tumor site.

accommodate for fluorescence imaging and 19F MRI is very


D evelopment of smart activatable nanoprobes with
different imaging modalities for early diagnosis of various
diseases has attracted tremendous attention.1 In particular,
intriguing.
Among various biological stimulus imaging probes such as
those dual-modal imaging systems such as use high spatial pH,20,21 enzyme,17,22 redox, and so on, redox-activated imaging
resolution of fluorescence modality2−4 and deep tissue probes are extremely intriguing because of the important role
penetration of magnetic resonance imaging (MRI)5−9 are of redox processes in organisms.23−26 The reduced glutathione,
capable of taking advantage of features of each imaging a peptide composed of glutamic acid, cysteine, and glycine, is
modality and overcoming the corresponding limitation. 1H an essential endogenous antioxidant that plays a vital part in
MRI has been widely utilized in clinics, of which the signal was physiological functions such as defense against toxins and free
originated from the protons in biological systems.10 Given that radicals.27,28 In addition, the concentration of GSH is
diagnosis of diseases is based on the slight signal variation implicated in many diseases typically associated with cancer,28
between health and pathological regions, it is particularly dementia,29 or neurodegenerative diseases.30 To date, GSH-
challenging for discriminating areas of interest. Thus, various responsive 19F MRI probes were mainly constructed by
1
H MRI contrast agents were explored to adjust the relaxation introducing redox sensitive moieties such as disulfide linkers31
rates (longitudinal relaxation time (T1) and transverse and 4-dinitrobenzenesulfonyl groups32 in fluorinated polymers.
relaxation time (T2)) of naturally occurring water molecules, The presence of GSH would either induce assemble or
so as to increase the MRI image contrast,11 yet, only to a limit disassemble of polymers or modulate paramagnetic relaxation
degree. Different from traditional 1H MRI, 19F MRI utilized enhancement (PRE) effects toward fluorinated moieties,
exogenous fluorinated contrast agents, thus with negligible resulting in activation or attenuation of 19FMRI signals.33,34
endogenous background noise, which further provides In addition, most of the reported probes can hardly achieve
quantitative information with high specificity.12−14 Therefore, “one stone for two birds” results, that is, using one single
19
F MRI may become a promising supplementary technique trigger to induce different signal responses, which would
for 1H MRI because it can provide accurate and intuitive
images in pathological imaging through the anatomical Received: October 13, 2020
position of 1H MRI. In addition to combined imaging Accepted: October 30, 2020
modalities, activated nanoprobes could gain information Published: November 11, 2020
about dynamic processes and obtain functional information
in real time, which can facilitate more specific diagnosis.15−20
In this regard, exploration of activatable nanoprobes that could

© 2020 American Chemical Society https://dx.doi.org/10.1021/acs.analchem.0c04301


15679 Anal. Chem. 2020, 92, 15679−15684
Analytical Chemistry pubs.acs.org/ac Article

improve the diagnosis accuracy. There is only one GSH- medium (DMEM) and trypsin were purchased from M&C
activatable nanoprobe with dual-mode turn-on imaging Gene Technology. Fetal calf serum (FBS) was obtained from
properties,15 yet, the 19F MRI signal of which is far from Hangzhou Sijiqing Bioengineer Materials Ltd. All of the above
enough for in vivo imaging. Herein, we synthetized a GSH chemicals were used directly without further purification.
activated dual-modal nanoprobe for19 F MRI and fluorescence Deionized (DI) water was used throughout all experiments.
imaging of tumors. As shown in Scheme 1, 19F moiety and Instrumentation. Transmission electron microscope
(TEM) images were acquired by using a JEOL JEM-1200EX
Scheme 1. Illustration Scheme of GSH-Activated FAN@ (100 kV), and high-resolution transmission electron micros-
MnO2 Nanoprobes for 19F MRI and Fluorescence Turn-On copy (HRTEM) images were obtained via a JEOL JEM-2100F
Imaging transmission electron microscope (200 kV). Dynamic light
scattering (DLS) particle size analysis and zeta potentials were
measured by using a Zeta sizer Nano-ZS90 zeta and size
analyzer from Malvern. Fluorescence spectra were recorded on
an F-4600 fluorescence spectrophotometer (Hitachi, Japan).
19
F NMR spectra were obtained from a Bruker Avance-III 400
MHz spectrometer. The cytotoxicity experiment was tested
using an ELISA plate reader (F50, TECAN). Cell imaging was
carried out using a laser scanning confocal microscope (Leica
SP8). Energy-dispersive spectroscopy (EDS) was acquired by
using a scanning electron microscope (SEM, SU-8010). X-ray
diffraction (XRD) patterns were recorded on a Shimadzu
XRD-7000 X-ray diffractometer using Cu Kα radiation (λ =
1.5418 Å). X-ray photoelectron spectroscopy (XPS) was
measured on an X-ray photoelectron spectrometer from
AgInS2 QD loaded nanoparticles (FANs) were successfully ThermoFisher Scientific USA. All MRI experiments were
prepared by encapsulating AgInS2 and perfluoro-15-crown-5- performed using a 7.0 T Bruker Bio-Spec70/20USR MRI
ether (P19FCE) with amphiphilic polymers (oleylamine and 1- system. Fluorescence imaging was performed on a small animal
(3-aminopropyl) imidazole functionalized polysuccinimide), live imaging system from PE company (IVIS Spectrum). The
which were further coated with manganese dioxide (MnO2) Mn2+ concentration tests were carried out on an inductively
nanoshells through in situ reduction of KMnO4 with 2-(N- coupled plasma optical emission spectroscopy (ICP-OES,
morpholino) ethanesulfonic acid (MES) to afford FAN@ Thermo Scientific iCAP 6000 series).
MnO2 nanoprobes. The MnO2 coating could effectively Preparation of Both 19F Moiety and AgInS2 QD
quench the fluorescence and 19F NMR signals through the Loaded Nanoparticles (FANs). In brief, the as-synthesized
energy transfer process and PRE effect, respectively, yet the AgInS2 QDs (2 mg) (details in the Supporting Information),
elevated GSH level in tumors would induce the decomposition amphiphilic polymer (25 mg), and P19FCE (5 μL) were
of MnO2 into Mn2+, accordingly achieving recovery of both dissolved in 1.0 mL of CHCl3. Then, the mixture was added
fluorescence and 19F NMR signals. into 10 mL of DI water (containing 0.32 mg of NaOH) with

■ EXPERIMENTAL SECTION
Reagents and Chemicals. NaOH, NaH2PO4·2H2O,
ultrasonication treatment and magnetic stirring (600 W for 6
min, pulsed working as 3 s “on” and 3 s “off”). Once the
resultant mixture solution changed into an emulsion, the
Na2HPO4·12H2O, KMnO4, cyclohexane, 1-dodecanethiol CHCl3 was evaporated at 38 °C. After being centrifuged at a
(DDT), chloroform, ethanol, methanol, N, N-dimethylforma- speed of 18,000 rpm for 20 min, the precipitation was
mide (DMF), and dimethyl sulfoxide (DMSO) were supplied dispersed into DI water (1.0 mL) and stored for later use.
by Beijing Chemical Factory. 1-(3-Aminopropyl) imidazole Preparation of Water-Dispersible FAN@MnO2. In situ
and N-methylmaleimide (NMM) were purchased from Alfa formation of MnO2 nanoshells on FANs was carried out by
Aesar Chemical Company. Polysuccinimide (PSI, MW ∼7000 adding 500 μL of KMnO4 aqueous solution (10 mM) into the
Da) was obtained from Shijiazhuang Desai Chemical as-prepared FAN colloid solution (1.0 mL) in the presence of
Company (China). Oleylamine (OAm) and 1-octadecene MES buffer (4.0 mL, 0.1 M, pH = 6.0). It was observed that
(ODE) were obtained from Sigma-Aldrich. The PSI was the color of solution immediately became dark brown. Then,
functionalized with both OAm and 1-(3-aminopropyl) the solution was vigorously stirred for 30 min at room
imidazole to get the amphiphilic polymer for the surface temperature before centrifugation at 15,000 rpm for 10 min.
coating of hydrophobic AgInS2 QDs. Indium nitrate hydrate The as-obtained precipitate was washed with DI water twice
and glutathione (GSH) were purchased from Aladdin. Stearic and then redispersed in DI water. Finally, PEG-2000 (10 mg)
acid was supplied by TCI. AgNO3 was purchased from was added. The colloidal solution was stirred at room
Innochem company. 2-(N-morpholino) ethanesulfonic acid temperature for 30 min. FAN@MnO2 was collected by
(MES) was obtained from Amethyst company. Sublimed sulfur centrifugation at 1000 rpm for 6 min to remove large
and PEG-2000 were purchased from Xilong Chemical nanoparticles. The obtained supernatant solution was centri-
Industrial Company (China). Perfluoro-15-crown-5-ether fuged at 15,000 rpm for 10 min. The precipitate was dispersed
(P19FCE) was supplied by Matrix scientific company. NO in 1.0 mL of DI water and stored in the dark at 4 °C.
was prepared from diethylamine NONOate sodium salt In Vivo Experiments. All animal experiments and
hydrate (DEA NONOate). For the cell culture, phosphate- treatments were approved by the local ethics review committee
buffered solution and methyl thiazolyltetrazolium were and conducted in accordance with the guidelines of the Animal
obtained from Amresco Inc. Dulbecco’s modified eagle Care and Use Committee of Beijing University of Chemical
15680 https://dx.doi.org/10.1021/acs.analchem.0c04301
Anal. Chem. 2020, 92, 15679−15684
Analytical Chemistry pubs.acs.org/ac Article

Technology. Four-week old female nude mice (Balb/c) were overlap between the absorption spectrum of MnO2 and the
used for animal experiments. The 4T1 cells (∼107) were excitation spectrum of FAN (Figure S3), a dramatic
injected subcutaneously into the right hind limb of mice. When fluorescence quenching was observed along with the increased
the tumor size grew to a size of about 900 mm3, FAN@MnO2 dosages of KMnO4 (Figure S4a). Similar to the trend of
nanoprobe colloidal solution (200 μL, 12 mg/mL) was fluorescence, the 19F NMR signal was gradually decreased as
injected into the tumor, and as a control, an equal amount well (Figure 4a), and a final dosage of KMnO4 (10 mM) was
of FAN@MnO2 nanoprobe colloidal solution was injected fixed at 0.5 mL for the formation of MnO2 nanoshells. The
subcutaneously into the healthy tissues in the left hind limb of formation of MnO2 was further supported by the zeta potential
the same mouse. Those mice were anesthetized by using changes. As shown in Figure S4b, the pristine FAN was
isoflurane prior to MRI measurements. The T1 RARE positively charged with the zeta potential of +62 mV. After
sequence was applied for 19F MRI (FOV = 50 mm × 50 coating MnO2 nanoshells, the zeta potential of FAN@MnO2
mm, slice thickness = 5 mm, TR = 3000 ms, effective TE = nanoprobes was gradually decreased. Furthermore, the XRD,
4.24 ms, echo spacing = 4.24 ms, RARE factor = 4, NEX = 20, EDS analysis results, and XPS results (Figure S5) demon-
matrix size = 100 × 100, acquisition time = 19 min). strated the successful construction of FAN@MnO2. The as-

■ RESULTS AND DISCUSSION


AgInS2 QDs were prepared through a previously published
prepared FAN@MnO2 displayed a good stability as no obvious
size changes were observed for at least 20 days both in the DI
water and FBS medium (Figures S6 and S7).
method with slight modification.35 As shown in Figure 1a, the Initially, the 19F NMR signal recovery ability of FAN@
MnO2 nanoprobes was investigated in the presence of GSH
(10 mM) at pH 6.0, as it was assumed that the in situ formed
MnO2 nanoshells could be degraded under tumor micro-
environments.37,38 As for FAN@MnO2 nanoprobes, due to the
intense PRE effect of MnO2, the 19F NMR signal intensities of
nanoprobes would be attenuated. When in the presence of
GSH at pH 6.0, an obvious increase of the 19F NMR signal
intensity, a singlet peak at −91.8 ppm, was observed, as shown
in Figure S8. It was reasoned that the MnO2 nanoshells of
nanoprobes were decomposed to Mn (II) ions by GSH at pH
6.0, and free Mn (II) would be away from nanoprobes to
impair the PRE effect, resulting in recovery of 19F NMR signal
intensities. The impairment of the PRE effect was also
supported by the changes of measured 19F NMR relaxation
time, listed in Table S1, where a clear increase of T2 from 129
to 708 ms was observed after incubation with GSH (10 mM)
at pH 6.0. Inversely proportional to T2 relaxation time, the half
peak widths of nanoprobes decreased from 33.5 to 14.3 Hz,
again demonstrating the proposed mechanism in FAN@MnO2
nanoprobes. Moreover, the decomposition of MnO2 was also
verified by the decreased size and morphology changes of
Figure 1. (a) TEM and HRTEM (inset of 1a) images of AgInS2 QDs;
(b) excitation (black line, λex = 500 nm) and emission (red line, λem =
FAN@MnO2 nanoprobes (Figure S9). At the same time, the
654 nm) spectra of AgInS2 QDs; the inset photo was hydrophobic ICP-OES results showed that the Mn2+ ions in the supernatant
AgInS2 QDs dispersed in chloroform under daylight (left) and 365 had increased from 1.216 to 5.345 μg/mL after adding 10 mM
nm light (right); (c) TEM image and (d) DLS size distribution of of GSH. All these observations strongly supported the
FAN@MnO2 nanoprobes, respectively. assumption that MnO2 would be decomposed under tumor
microenvironments.
TEM and HRTEM images showed that the average size of Then, we investigated the influence of incubation time by
AgInS2 QDs was around 4.0 nm, with a clear lattice spacing of measuring the 19F NMR signal of a mixed solution containing
0.33 nm, well consistent with the (112) facet of the tetragonal FAN@MnO2 nanoprobes after incubation with different
AgInS2 phase observed by XRD patterns (Figure S1).36 The concentrations of GSH for different time intervals. As shown
acquired hydrophobic AgInS2 QD suspension displayed in Figure 2a, at pH 6.0, the 19F NMR signal was gradually
intense fluorescence with maximal emission at 654 nm (Figure recovered by adding GSH, and the 19F signal-to-noise ratio
1b). Under ultrasonication treatment, the FAN with a particle (SNR) reached a plateau at 2 h. Moreover, the 19F NMR signal
size of 78.3 ± 25.8 nm was successfully fabricated by of nanoprobes was gradually recovered along with the increase
encapsulating AgInS2 QDs and P19FCE with the as-synthesized of GSH in a weak acid buffer solution (pH 6.0). Yet, the signal
amphiphilic polymer (Figure S2a). The fluorescence of the was stable, and no significant enhancement was observed
FAN dispersed in water shows a slight decrease, compared to under neutral buffer solution (pH 7.4), as seen in Figure 2b,
that of AgInS2 QDs dispersed in chloroform (Figure S2b). which suggested that acidic conditions would facilitate MnO2
Moreover, the FAN was further coated with MnO2 nanoshells. decomposition. In addition, in vitro 19F MRI was also
The TEM image (Figure 1c) indicated that FAN@MnO2 evaluated; as shown in Figure 2c, when the concentration of
nanoprobes were successfully synthesized by in situ coating GSH increased, the 19F MRI of the nanoprobes enhanced. The
19
MnO2 nanoshells. Meanwhile, the DLS size of FAN@MnO2 F MRI SNR was proportional to the concentration of GSH
nanoprobes (114.8 ± 37.7 nm) was larger than that of the (Figure 2d), with a linear regression equation of SNR = 0.710
FAN (78.3 ± 25.8 nm) (Figure 1d). Due to the significant C + 6.70 (R2 = 0.958). Herein, the C represented the
15681 https://dx.doi.org/10.1021/acs.analchem.0c04301
Anal. Chem. 2020, 92, 15679−15684
Analytical Chemistry pubs.acs.org/ac Article

Figure 3. Intracellular imaging of HeLa cells with different treatments.


Cells incubated with FAN@MnO2 nanoprobes for (a1−a3) 0.5 h and
(b1−b3) 6 h; (c1−c3) cells pretreated with NMM (500 μM) for 30
min followed by incubation nanoprobes for 6 h. Scale bar was 20 μm.

Then, nanoprobes were taken up by the cells after 6 h. As a


positive control, FAN nanocomposites were incubated with
HeLa cells and an intense emission was observed (Figure S15),
Figure 2. (a) Time-dependent evolution of 19F NMR signal-to-noise suggesting that MnO2 formation accounts for the initial weak
ratio (SNR) of the FAN@MnO2 nanoprobes in the presence of emission signals of FAN@MnO2 nanoprobe-treated HeLa
different concentrations of GSH at pH = 6.0. (b) SNR recovery of cells. To further investigate whether FAN@MnO2 was able to
FAN@MnO2 nanoprobes containing different concentrations of GSH respond to intracellular GSH in living cells, the cells were
under different pH values (7.4 and 6.0). (c) 19F MRI of FAN@MnO2 pretreated with the GSH scavenger (N-methylmaleimide,2
nanoprobes incubated with different concentrations of GSH at pH =
6.0. (d) Corresponding calibration plot of 19F MRI SNR versus the
NMM) for 30 min and then incubated with the cells for 6 h
concentration of GSH. (e) Fluorescence imaging of FAN@MnO2 before cell imaging. It was worth noting that the fluorescence
nanoprobes incubated with different concentrations of GSH at pH = of cells treated with NMM was significantly decreased (Figure
6.0 (λex = 500 nm; λem = 640 nm). 3c1−c3) compared with that of untreated cells (Figure 3b1−
b3), suggesting the specificity of FAN@MnO2 toward GSH. In
concentration of GSH (mM). Given that the broad absorption addition, a 4T1 xenograft tumor model was used to evaluate
of MnO2 nanoshells overlaps with the fluorescence excitation the ability of intratumoral glutathione for lighting up FAN@
of the FAN, the fluorescence intensities of FAN@MnO2 MnO2. As shown in Figure S16, strong fluorescence intensity
nanoprobes experienced a clear turn-on response with the was observed at the tumor site, suggesting the tumor
increment of GSH under weakly acidic conditions due to the microenvironment responsiveness of the as-prepared nanop-
decomposition of MnO2 (Figures S10a, S10b and Figure 2e). robes.
Meanwhile, a series of other species that commonly occurred Encouraged by these promising results, we explored the in
in biological systems including L-cysteine, H2S, KSCN, NO, vivo utilization possibility of these nanoprobes. As shown in
H2O2, and NaNO2 were tested (Figures S11 and S12), which Figure 4, the same amounts of FAN@MnO2 nanoprobe
indicated that FAN@MnO2 displayed high specificity toward colloidal solution were injected into the tumor and the healthy
GSH under biological conditions. tissues of the same mouse, respectively. As indicated in Figure
Prior to imaging, the cell cytotoxicity of FAN@MnO2 4b,c, the stronger 19F MRI signals were observed at the tumor
nanoprobes was evaluated. As shown in Figure S13a, FAN@ site 1 h post injection than those in normal tissues, implying
MnO2 nanoprobes were incubated with HUVEC cells for 24 tumor microenvironment activatable features of FAN@MnO2.
and 48 h, respectively. When the concentration of the FAN@ These results demonstrated that the FAN@MnO2 nanoprobes
MnO2 nanoprobes reached 250 μg/mL, 84% viability was could indeed be turned on under tumor microenvironments.
acquired after 48 h, indicating that nanoprobes had good
biocompatibility. However, the viability of HeLa cells (Figure
S13b) was about 55% after 48 h, and the viability of 4T1 cells
(Figure S13c) was about 50%, which might be attributed to the
elevated GSH level in cancer cells than that in normal cells,39
implying that the FAN@MnO2 nanoprobes may serve as
theranostic platforms for tumors.
In vitro imaging was initially performed by incubating
nanoprobes with HeLa cells for different time periods (0.5, 2, Figure 4. (a) In vivo 1H MRI, (b) 19F MRI, and (c) merged images
4, and 6 h, Figure 3 and Figure S14). As shown in Figure S14, for tumor-bearing mouse by injecting the same amounts of FAN@
the fluorescence signal was almost undetectable at 0.5 h, and MnO2 nanoprobes at tumor sites and the contralateral site,
red fluorescence could be seen on the cell membrane at 4 h. respectively.

15682 https://dx.doi.org/10.1021/acs.analchem.0c04301
Anal. Chem. 2020, 92, 15679−15684
Analytical Chemistry


pubs.acs.org/ac Article

CONCLUSIONS Matter Science and Engineering, Beijing University of


In conclusion, we designed tumor microenvironment activated Chemical Technology, Beijing 100029, P. R. China
FAN@MnO2 nanoprobes for simultaneous fluorescence and Yunhe Yan − State Key Laboratory of Chemical Resource
19
F MRI turn-on imaging. The nanoprobes were fabricated by Engineering, Beijing Advanced Innovation Center for Soft
Matter Science and Engineering, Beijing University of
coating the 19F moiety and fluorescence AgInS2 QD core with
Chemical Technology, Beijing 100029, P. R. China
in situ formed MnO2 nanoshells, where the fluorescence
Chang Guo − State Key Laboratory of Chemical Resource
intensity and the 19F MRI signal were quenched by adjacent
Engineering, Beijing Advanced Innovation Center for Soft
MnO2 shells through the energy transfer process and PRE
Matter Science and Engineering, Beijing University of
effect, respectively. Under slightly acidic conditions with high
Chemical Technology, Beijing 100029, P. R. China
GSH levels, the MnO2 shells were decomposed, which resulted
in the shielded signals turn-on. Both the in vitro and in vivo Complete contact information is available at:
imaging results demonstrated that these nanoprobes were https://pubs.acs.org/10.1021/acs.analchem.0c04301
sensitive to tumor microenvironments, thus possessing great
potential for precisely lighting up the tumor site without Notes
inducing off-targeting signals. The authors declare no competing financial interest.



ASSOCIATED CONTENT
* Supporting Information
■ ACKNOWLEDGMENTS
The authors gratefully acknowledge the financial support from
The Supporting Information is available free of charge at the National Natural Science Foundation of China (21675009,
https://pubs.acs.org/doi/10.1021/acs.analchem.0c04301. 21725501, and 21874007) and the Fundamental Research
Funds for the Central Universities (XK1901).


XRD pattern of AgInS2 QDs (Figure S1); DLS, TEM
image, excitation and emission spectra of FAN (Figure REFERENCES
S2); excitation spectrum of FAN and absorption
spectrum of MnO2 (Figure S3); 19F NMR signal (1) Chen, Y. J.; Wu, S. C.; Chen, C. Y.; Tzou, S. C.; Cheng, T. L.;
Huang, Y. F.; Yuan, S. S.; Wang, Y. M. Biomaterials 2014, 35, 304−
intensities, fluorescence intensities, and zeta potentials 315.
of FAN@MnO2 (Figure S4); XRD, EDS, and XPS of (2) Deng, R.; Xie, X.; Vendrell, M.; Chang, Y.-T.; Liu, X. J. Am.
FAN@MnO2 (Figure S5); stability investigation of Chem. Soc. 2011, 133, 20168−20171.
FAN@MnO2 nanoprobes (Figures S6 and S7); proper- (3) Qin, Y. T.; Peng, H.; He, X. W.; Li, W. Y.; Zhang, Y. K. Anal.
ties of FAN@MnO2 nanoprobes with and without GSH Chem. 2019, 91, 12696−12703.
(Table S1 and Figure S8); DLS and TEM image of (4) Xu, S.; Cui, J.; Wang, L. TrAC, Trends Anal. Chem. 2016, 80,
FAN@MnO2 nanoprobes after incubation with GSH 149−155.
(Figure S9); responses of FAN@MnO2 nanoprobes (5) Cui, J.; Jiang, R.; Guo, C.; Bai, X.; Xu, S.; Wang, L. J. Am. Chem.
toward GSH (Figure S10) and species in biological Soc. 2018, 140, 5890−5894.
systems (Figures S11 and S12); cell viability tests of (6) Guo, C.; Xu, M.; Xu, S.; Wang, L. Nanoscale 2017, 9, 7163−
7168.
nanoprobes (Figure S13); fluorescence imaging of (7) Hu, G.; Li, N.; Tang, J.; Xu, S.; Wang, L. ACS Appl. Mater.
FAN@MnO2 nanoprobes in HeLa cells (Figure S14 Interfaces 2016, 8, 22830−22838.
and S15); and in vivo fluorescence imaging (Figure S16) (8) Chen, H.; Song, M.; Tang, J.; Hu, G.; Xu, S.; Guo, Z.; Li, N.;
(PDF) Cui, J.; Zhang, X.; Chen, X.; Wang, L. ACS Nano 2016, 10, 1355−


1362.
(9) Zhang, H.; Peng, S.; Xu, S.; Chen, Z. RSC Adv. 2016, 6,
AUTHOR INFORMATION 104731−104734.
Corresponding Authors (10) Yan, K.; Li, H.; Li, P. H.; Zhu, H. E.; Shen, J.; Yi, C. F.; Wu, S.
Suying Xu − State Key Laboratory of Chemical Resource L.; Yeung, K. W. K.; Xu, Z. S.; Xu, H. B.; Chu, P. K. Biomaterials
Engineering, Beijing Advanced Innovation Center for Soft 2014, 35, 344−355.
(11) Peterson, K.; Srivastava, K.; Pierre, V. C. Front. Chem. 2018, 6,
Matter Science and Engineering, Beijing University of 2296−2646.
Chemical Technology, Beijing 100029, P. R. China; (12) Jahromi, A. H.; Wang, C.; Adams, S. R.; Zhu, W.; Narsinh, K.;
orcid.org/0000-0001-6638-0040; Email: syxu@ Xu, H.; Gray, D. L.; Tsien, R. Y.; Ahrens, E. T. ACS Nano 2019, 13,
mail.buct.edu.cn; Fax: (86)10-64427869 143−151.
Leyu Wang − State Key Laboratory of Chemical Resource (13) Hu, G.; Tang, J.; Bai, X.; Xu, S.; Wang, L. Nano Res. 2016, 9,
Engineering, Beijing Advanced Innovation Center for Soft 1630−1638.
Matter Science and Engineering, Beijing University of (14) Munkhbat, O.; Canakci, M.; Zheng, S.; Hu, W.; Osborne, B.;
Chemical Technology, Beijing 100029, P. R. China; Bogdanov, A. A.; Thayumanavan, S. Biomacromolecules 2019, 20,
orcid.org/0000-0002-5961-7764; Email: lywang@ 790−800.
mail.buct.edu.cn (15) Zheng, M. M.; Wang, Y. Q.; Shi, H.; Hu, Y. X.; Feng, L. D.;
Luo, Z. L.; Zhou, M.; He, J.; Zhou, Z. Y.; Zhang, Y.; Ye, D. J. ACS
Authors Nano 2016, 10, 10075−10085.
(16) Akazawa, K.; Sugihara, F.; Minoshima, M.; Mizukami, S.;
Yangyang Zhang − State Key Laboratory of Chemical Kikuchi, K. Chem. Commun. 2018, 54, 11785−11788.
Resource Engineering, Beijing Advanced Innovation Center (17) Yuan, Y.; Sun, H. B.; Ge, S. C.; Wang, M. J.; Zhao, H. X.; Wang,
for Soft Matter Science and Engineering, Beijing University of L.; An, L. N.; Zhang, J.; Zhang, H. F.; Hu, B.; Wang, J. F.; Liang, G. L.
Chemical Technology, Beijing 100029, P. R. China ACS Nano 2015, 9, 761−768.
Qian Ma − State Key Laboratory of Chemical Resource (18) Akazawa, K.; Sugihara, F.; Nakamura, T.; Mizukami, S.;
Engineering, Beijing Advanced Innovation Center for Soft Kikuchi, K. Bioconjugate Chem. 2018, 29, 1720−1728.

15683 https://dx.doi.org/10.1021/acs.analchem.0c04301
Anal. Chem. 2020, 92, 15679−15684
Analytical Chemistry pubs.acs.org/ac Article

(19) Huang, J.; Xu, Y.; Xiao, H.; Xiao, Z.; Guo, Y.; Cheng, D.; Shuai,
X. ACS Nano 2019, 13, 7036−7049.
(20) Song, G.; Zheng, X.; Wang, Y.; Xia, X.; Chu, S.; Rao, J. ACS
Nano 2019, 13, 7750−7758.
(21) Shi, H.; Lei, Y.; Ge, J.; He, X.; Cui, W.; Ye, X.; Liu, J.; Wang, K.
Anal. Chem. 2019, 91, 9154−9160.
(22) Guo, C.; Zhang, Y.; Li, Y.; Xu, S.; Wang, L. Anal. Chem. 2019,
91, 8147−8153.
(23) Liu, D.; Chen, B.; Mo, Y.; Wang, Z.; Qi, T.; Zhang, Q.; Wang,
Y. Nano Lett. 2019, 19, 6964−6976.
(24) Hu, X. C.; Xu, Z. L.; Hu, J. W.; Dong, C. Y.; Lu, Y. L.; Wu, X.
W.; Wumaier, M.; Yao, T. M.; Shi, S. Inorg. Chem. Front. 2019, 6,
2865−2872.
(25) Zhang, T. T.; Xu, C. H.; Zhao, W.; Gu, Y.; Li, X. L.; Xu, J. J.;
Chen, H. Y. Chem. Sci. 2018, 12, 6741−6271.
(26) Pinto, S. M.; Tomé, V.; Calvete, M. J. F.; Castro, M. M. C. A.;
Tóth, E.; Geraldes, C. F. G. C. Coord. Chem. Rev. 2019, 390, 1−31.
(27) Cao, N.; Zeng, P.; Zhao, F.; Zeng, B. Talanta 2019, 204, 402−
408.
(28) Wu, R.; Ge, H.; Liu, C.; Zhang, S. H.; Hao, L.; Zhang, Q.;
Song, J.; Tian, G. H.; Lv, J. G. Talanta 2019, 196, 191−196.
(29) Dong, W.; Wa, R.; Gong, X.; Dong, C. Anal. Bioanal. Chem.
2019, 411, 6687−6695.
(30) Gao, Y.; Wu, K.; Li, H.; Chen, W.; Fu, M.; Yue, K.; Zhu, X.;
Liu, Q. Sens. Actuators B Chem. 2018, 273, 1635−1639.
(31) Fu, C. K.; Tang, J.; Pye, A.; Liu, T. Q.; Zhang, C.; Tan, X.; Han,
F.; Peng, H.; Whittaker, A. K. Biomacromolecules 2019, 20, 2043−
2050.
(32) Huang, P. S.; Guo, W. S.; Yang, G.; Song, H. J.; Wang, Y. Q.;
Wang, C.; Kong, D. L.; Wang, W. W. ACS Appl. Mater. Interfaces
2018, 10, 18532−18542.
(33) Nakamura, T.; Matsushita, H.; Sugihara, F.; Yoshioka, Y.;
Mizukami, S.; Kikuchi, K. Angew. Chem., Int. Ed. 2015, 54, 1007−
1010.
(34) Li, Y.; Zhang, H.; Guo, C.; Hu, G.; Wang, L. Anal. Chem. 2020,
92, 11739−11746.
(35) Gu, S.; Guo, C.; Wang, H.; Tian, G.; Xu, S.; Wang, L. Front.
Chem. 2019, 7, 734.
(36) Xiang, W.; Xie, C.; Wang, J.; Zhong, J.; Liang, X.; Yang, H.;
Luo, L.; Chen, Z. J. Alloys Compd. 2014, 588, 114−121.
(37) Min, H.; Wang, J.; Qi, Y.; Zhang, Y. L.; Han, X. X.; Xu, Y.; Xu,
J. C.; Li, Y.; Chen, L.; Cheng, K. M.; Liu, G. N.; Yang, N.; Li, Y. Y.;
Nie, G. J. Adv. Mater. 2019, 31, 1808200.
(38) Fan, H.; Yan, G.; Zhao, Z.; Hu, X. X.; Zhang, W. H.; Liu, H.;
Fu, X. Y.; Fu, T.; Zhang, X. B.; Tan, W. H. Angew. Chem., Int. Ed.
2016, 55, 5477−5482.
(39) Zhou, F.; Zheng, T.; Abdel-Halim, E. S.; Jiang, L.; Zhu, J.-J. J.
Mater. Chem. B 2016, 4, 2887−2894.

15684 https://dx.doi.org/10.1021/acs.analchem.0c04301
Anal. Chem. 2020, 92, 15679−15684

You might also like