Telechargé par saadsoumia79

Article Exposé TD stress Oxydatif

publicité
Journal of Functional Foods 90 (2022) 104954
Contents lists available at ScienceDirect
Journal of Functional Foods
journal homepage: www.elsevier.com/locate/jff
Fisetin represses oxidative stress and mitochondrial dysfunction in NAFLD
through suppressing GRP78-mediated endoplasmic reticulum (ER) stress
Xianling Dai a, b, 1, Qin Kuang a, b, 1, Yan Sun a, b, 1, Minxuan Xu a, b, c, Liancai Zhu a, *,
Chenxu Ge a, b, c, *, Jun Tan b, c, *, Bochu Wang a, *
a
Key Laboratory of Biorheological Science and Technology (Chongqing University), Ministry of Education, College of Bioengineering, Chongqing University, Chongqing
400030, PR China
Chongqing Key Laboratory of Medicinal Resources in the Three Gorges Reservoir Region, School of Biological and Chemical Engineering, Chongqing University of
Education, Chongqing 400067, PR China
c
Research Center of Brain Intellectual Promotion and Development for Children Aged 0-6 Years, Chongqing University of Education, Chongqing 400067, PR China
b
A R T I C L E I N F O
A B S T R A C T
Keywords:
NAFLD
Fisetin
ROS
Mitochondrial impairment
ER stress
Fisetin (FisT) is a bioactive flavonoid polyphenol with antioxidant, anti-inflammatory and anti-tumor activities.
Although the effects of FisT to meliorate non-alcoholic fatty liver disease (NAFLD) have been investigated, the
underlying mechanisms are not fully understood. In the present study, we found that FisT remarkably suppressed
cellular and mitochondrial reactive oxygen species (ROS) generation in human and murine hepatocytes with
palmitate (PA) stimulation. Additionally, mitochondrial impairment and dysfunction induced by PA were
considerably abrogated in hepatocytes with FisT co-incubation. Furthermore, Cyto-c releases and mitochondrial
apoptosis were detected in PA-treated hepatocytes, while being greatly repressed by FisT. PA-induced inflam­
mation and lipid deposition were also strongly reduced by FisT in hepatocytes. Importantly, our in vitro exper­
iments showed that promoting ROS by nuclear factor erythroid 2-related factor 2 (Nrf2) deletion significantly
abolished the function of FisT to meliorate apoptosis, inflammation and lipid accumulation in PA-incubated
hepatocytes. What’s more, ER stress was strongly induced by PA via increasing 78-kDa glucose-regulated pro­
tein (GRP78) and C/EBP-homologous protein (CHOP), which were, however, dramatically repressed after FisT
co-exposure. Intriguingly, we found that strengthening ER stress by GRP78 over-expression considerably abol­
ished the capacity of FisT to retard ROS generation and mitochondrial impairment in PA-stimulated hepatocytes,
but GRP78 knockdown exhibited totally opposite effects. Thus, ER stress blockage was required for FisT to
ameliorate NAFLD development in vitro. Consistently, our in vivo studies using high fat diet (HFD)-fed mice
confirmed that FisT administration exerted inhibitory and therapeutic potential on fatty liver progression via the
same mechanisms monitored in vitro. Collectively, all our findings disclosed that FisT can efficiently attenuate
NAFLD through restraining ROS generation and mitochondrial impairment mediated by ER stress.
1. Introduction
Excessive fat deposition contributes to non-alcoholic fatty liver dis­
ease (NAFLD) progression. Although advances in NAFLD treatment have
been achieved, a large number of patients still has worse clinical
outcome (Im et al., 2021; Zhou et al., 2020). The consequential surplus
of lipids in hepatocytes leads to oxidative stress and lipotoxicity and
promotes mitochondrial dysfunction through different molecular
mechanisms (Friedman et al., 2018). Mitochondria play irreplaceable
role in adenosine triphosphate (ATP) production and modulate various
Abbreviations: HO-1, heme oxygenase-1; NQO1, NAD(P)H dehydrogenase (quinone 1); GCLM, glutamate-cysteine-ligase modifier; NOX1, NADPH oxidase 1;
NOX2, NADPH oxidase 2; NOX4, NADPH oxidase 4; TNF-α, tumor necrosis factor-α; IL-1β, interleukin-1β; IL-6, interleukin-6; FASN, fatty acid synthase; SCD1,
stearoyl-CoA desaturase 1; PPARγ, peroxisome proliferator-activated receptor-γ.
* Corresponding authors at: Chongqing Key Laboratory of Medicinal Resources in the Three Gorges Reservoir Region, School of Biological and Chemical Engi­
neering, Chongqing University of Education, Chongqing 400067, PR China (C. Ge and J. Tan).
E-mail addresses: [email protected] (L. Zhu), [email protected] (C. Ge), [email protected], [email protected] (J. Tan), [email protected]
(B. Wang).
1
These authors contributed equally to this work.
https://doi.org/10.1016/j.jff.2022.104954
Received 20 October 2021; Received in revised form 3 January 2022; Accepted 10 January 2022
Available online 3 February 2022
1756-4646/© 2022 Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
X. Dai et al.
Journal of Functional Foods 90 (2022) 104954
cellular processes in both physiological conditions and pathological
stress insults (Fanta et al., 2021). Mitochondrial impairment can induce
hepatocyte death and associated abnormal cellular events, which is a
key factor for NAFLD development and hepatic dysfunction (Eduardo
et al., 2021; Sanda et al., 2021). Emerging evidence has suggested that
defects in mitochondrial hemostasis are common characteristics of fatty
liver and maintaining healthy mitochondria is pivotal for the survival of
hepatocytes, contributing to NAFLD treatment (Lee, Park, & Roh, 2019).
Therefore, methods to suppress oxidative stress and mitochondrial
dysfunction may be effective for NAFLD management.
Recently, endoplasmic reticulum (ER) stress has attracted increasing
attention. ER is a crucial organelle where secreted and transmembrane
proteins are synthesized and correctly folded into three-dimensional
conformations. These above dynamic events need highly precise mod­
ulation to sustain normal protein homeostasis (Marciniak & Ron, 2006;
Schröder, 2008). Numerous stimuli including HFD can break the above
equilibrium, subsequently inducing the unfolded protein response
(UPR) and ER stress occurrence (Ghemrawi, Battaglia-Hsu, & Arnold,
2018; Iurlaro & Muñoz-Pinedo, 2016). Upon ER stress initiation, GRP78,
as a key ER stress sensor, is released. Severe ER stress can facilitate the
expression of pro-apoptotic proteins, such as CHOP. Finally, cells un­
dergo apoptotic cell death under stimulated conditions (Lee, 2005). In
addition, ER stress may result in the activation of multiple different
intracellular stress pathways, which can induce or accelerate insulin
resistance, dyslipidemia, inflammation and, in some cases, culminate in
hepatocyte cell death. All these processes are involved in the patho­
genesis of fatty liver (Zhang et al., 2014). Herein, finding therapeutic
approaches to reduce ER stress may be effective for NAFLD treatment.
Fisetin (FisT; 3,3′ ,4′ ,7-tetrahydroxyflavone) is an important dietary
flavonoid and presents in numerous fruits and vegetables. FisT has
antioxidant, anti-inflammatory, anti-cancer and neuroprotective prop­
erties (Grynkiewicz & Demchuk, 2019; Khan et al., 2013; Zhang et al.,
2018). We previously showed that FisT could ameliorate HFD-induced
fatty liver in mice through suppressing hepatic inflammation and lipid
deposition by depressing TNF-α/RIPK3 signaling pathway (Xu, Ge, &
Qin et al., 2019; Xu, Sun, & Dai et al., 2019). Recently, FisT was
demonstrated to prohibit ROS/ER stress-mediated inflammatory
response, contributing to the amelioration of metabolic stimuli-induced
cardiac injury both in vivo and in vitro (Ge et al., 2019). Besides, FisT can
ameliorate oxidative stress, inflammation and apoptosis in HFD- or
streptozotocin (STZ)-induced diabetic cardiomyopathy, exerting car­
dioprotective action (Althunibat et al., 2019; Hu et al., 2020). Moreover,
the potential of FisT to regulate mitochondrial function and apoptosis
was shown to suppress tumor growth and protect against myocardial
ischemia–reperfusion (I/R) injury (Pal et al., 2013; Sabarwal, Agarwal,
& Singh, 2017; Shanmugam et al., 2021). Although these protective
effects of FisT against metabolic stresses-triggered liver and heart injury
have been reported, the underlying molecular mechanisms and partic­
ularly the possible crosstalk between them are still not fully understood.
Here in the present study, we performed in vitro and in vivo experi­
ments to deeply explore and confirm the regulatory role of FisT on PAand HFD-induced hepatic injury, respectively. We newly found that FisT
could restrain cellular and mitochondrial ROS generation, contributing
to the inhibition of inflammatory response and lipid accumulation in PAincubated hepatocytes. More importantly, all these effects mediated by
FisT were largely attributed to the suppression of GRP78-mediated ER
stress. The function of FisT to depress ER stress was validated in HFDchallenged mice, consequently ameliorating NAFLD progression.
(ATCC; Manassas, VA, USA), respectively. All cells were cultured in
Dulbecco’s Modified Eagle Medium (DMEM) (#22320030, Gibco®)
supplemented with 10% fetal bovine serum (FBS; #10100147, Gibco®)
and 1% penicillin-streptomycin in a humidified incubator (Thermo
Fisher Scientific; USA) with 5% CO2 at 37 ◦ C. To imitate the in vivo liver
lipid deposition, cells were incubated with PA (Cat#P9767) obtained
from Sigma-Aldrich (St. Louis, USA).
2.2. Cell viability
Cell viability was measured using a Cell Counting Kit-8 assay kit
according to the manufacturer’s instructions (#C0039, Beyotime,
Shanghai, China). Briefly, cells were planted into 96-well plates. After
treatments, 10 μL of Cell Counting Kit-8 reagent was added into each
well. After incubation for 4 h at 37 ◦ C, the absorbance of each well at
450 nm was measured using a microplate reader (SpectraMax iD3,
Molecular Devices, USA) to examine the number of viable cells.
2.3. Transfection in vitro
Nrf2 si-RNAs (si-Nrf2), GRP78 si-RNAs (si-GRP78) and the corre­
sponding negative control si-RNAs (si-Con) were obtained from Generay
Biotechnology (Shanghai, China). The GRP78 plasmid based on
pcDNA3.1 vector and empty vector (EV) were constructed by GeneChem
Technology (Shanghai, China) using the standard molecular biology
techniques. Transfection for gene knockdown or over-expression was
performed using Lipofectamine 3000 reagent (Invitrogen Life Technol­
ogies, Carlsbad, USA) according to the provider’s instructions.
2.4. Animals and treatments
The animal study protocols were permitted and authorized by the
Institutional Animal Care and Use Committee in College of Bioengi­
neering, Chongqing University (Chongqing, China). The procedures
used in this study were in accordance with the Regulations of Experi­
mental Animal Administration issued by the Ministry of Science and
Technology of the People’s Republic of China. All male, C57BL/6N mice
(6- to 8-week-old; 22–25 g body weight) used in the study were pur­
chased from Beijing Vital River Laboratory Animal Technology Co., Ltd.
(Beijing, China). Mice were allowed to adapt to their living environment
for 1 week before all experiment proper starts. All animals were fed in a
constant temperature (23 ± 25 ◦ C), humidity (50–60%) and pathogenfree-controlled environment cage with a standard 12 h light/12 h dark
cycle, plenty of water and food (pathogen-free) in their cages.
Animal experimental design 1: Mice were then randomly divided
into 4 groups, including (i) the normal chow diet/Vehicle group (NCD/
Veh); (ii) NCD/FisT group; (iii) HFD/Veh group; and (iv) HFD/FisT
group. At the beginning of the experiments, HFD (60 kcal% fat; D12492;
Research Diets, New Brunswick, USA) was subjected to experimental
mice to induce NAFLD murine model. Mice fed with NCD (10.2% fat and
71.5% carbohydrates; D12450B; Research Diets) were served as the
normal controls. For pharmacological analysis of FisT, FisT (80 mg/kg)
as drug (HPLC purity ≥ 98%, Xi’an Ruiying Biotechnology Co. Ltd,
Xi’an, China) was given to mice by gavage daily with HFD for 16 weeks,
continuously (Xu, Ge, & Qin et al., 2019; Xu, Sun, & Dai et al., 2019).
Vehicle group of mice received the same volume of saline. Body weights
were measured during HFD and FisT treatments weekly.
Animal experimental design 2: Male C57BL/6N mice (6–8 weeks
old; 22–25 g) were fed HFD (D12492, Research Diets) for 16 weeks to
induce NASH. And after 8 weeks of HFD, mice were randomly assigned
to four experimental groups: (i) NCD/Veh; (ii) NCD/FisT group; (iii)
HFD/Veh group; and (iv) HFD/FisT group. FisT (80 mg/kg) was
administrated to mice every day by gavage for 8 weeks. Vehicle group of
mice received the same volume of saline.
After 16 weeks, blood was obtained through retro-orbital bleeding,
and the serum was separated. Liver samples from each mouse were
2. Materials and methods
2.1. Cells and culture
Human hepatocyte cell line L02 and murine hepatocyte AML12 were
purchased from the Type Culture Collection of the Chinese Academy of
Sciences (Shanghai, China) and American Type Culture Collection
2
X. Dai et al.
Journal of Functional Foods 90 (2022) 104954
removed, weighed and then embedded in paraffin for histological
analysis or stored at − 80 ◦ C for further analysis.
manufacturer’s protocols. Then, total RNA extraction was reverse
transcribed using the M-MLV-RT system (Invitrogen). Subsequently,
PCR were performed with SYBR Green (Bio-Rad) on an ABI PRISM
7900HT system (Applied Biosystems, USA). The specific primer se­
quences were produced by Invitrogen or Generay Biotech (Shanghai,
China), which were listed in Supplementary Table 1. Fold induction
values were evaluated according the 2(− ΔΔCt) expression. ΔCt represents
the differences in cycle threshold number between the target gene and
GAPDH, and ΔΔCt represents the relative change in the differences
between the control and treatment groups.
2.5. Liver function and lipid contents analysis
The concentration of serum alanine transaminase (ALT) (#MAK052,
Sigma-Aldrich), aspartate aminotransferase (AST) (#MAK055, SigmaAldrich), serum or hepatic triglyceride (TG) (#MAK266, SigmaAldrich) and total cholesterol (TC) (#ab65359, Abcam) were
measured using commercially available detection kits according to the
manufacturers’ protocols.
2.10. Western blotting analysis
2.6. ROS measurements
Cells or liver samples were homogenized using RIPA Lysis and
Extraction Buffer (#89900, Thermo Fisher Scientific) to yield a ho­
mogenate for total protein extraction. The Nuclear and Cytoplasmic
Protein Extraction Kit (#P0028; Beyotime) was used to extract nuclear
and cytosol protein from cells according to the manufacturer’s in­
structions. Protein concentrations were measured using a BCA Protein
Assay Kit (#P0012; Beyotime) with BSA as a standard following the
supplier’s protocols. Equal amounts of the obtained protein (20–50 μg)
were subjected to 10–12% sodium dodecyl sulfate polyacrylamide gel
electrophoresis (SDS-PAGE) system and then transferred to PVDF
membranes (GE Healthcare Life Science, Germany). Next, the mem­
branes were blocked using 5% skim milk (DifcoTM Skim Milk, USA) for
1 h and incubated with the primary antibodies (Supplementary Table 2;
diluted at 1:1000 or 1:500) at 4 ◦ C overnight. Then, the PVDF mem­
branes were washed and incubated with horseradish peroxidase (HRP)conjugated anti-rabbit (#ab6721) or anti-mouse (#ab6789) secondary
antibodies (Abcam, dilution 1:5000). Immunoblotting bands were
visualized with a PierceTM ECL Plus Western Blotting Substrate
(#32134, Thermo Fisher Scientific) and exposed to Kodak (Eastman
Kodak Company, USA) Xray film. Protein expression levels were finally
determined as grey value (Version 1.52v, Image J) and standardized to
housekeeping gene (GAPDH) and expressed as a fold of control.
After treatments, cellular ROS was stained using dichlorofluorescin
diacetate (DCFH-DA) dye (#S0033M; Beyotime). The viable cells were
then incubated with 10 μM of DCFH-DA for 20 min at 37 ◦ C. Mito­
chondrial ROS generation in cells following treatments was examined
using a MitoSOX Red Mitochondrial Superoxide Indicator Kit (YEASEN
Biotechnology, Shanghai, China) according to the provider’s in­
structions. Images indicating cellular and mitochondrial ROS production
were captured under a fluorescence microscope. Total ROS levels in the
liver sections were detected using the fluorescent probe DHE (dihy­
droethidium). In brief, cryosections from snap-frozen liver (5 μm thick)
tissues were prepared. In situ ROS was measured using DHE (#D23107;
Invitrogen) as the provider described. Cryosections were stained with
5 μM DHE for 30 min, mounted by anti-fluorescence quenching sealing
tablets in dark at room temperature and monitored under a fluorescence
microscope.
2.7. Mitochondrial function assays
JC-1 dye (#C2006; Beyotime) accumulates in mitochondria of cells
with mitochondrial depolarization to form monomers and emits green
fluorescence and was used to measure mitochondrial membrane po­
tential (MMP, Δψm) (Perelman et al., 2012). All cells were planted into
six-well plates and were subjected to drug treatments. Cells were then
incubated with 500 μL of JC-1 dye following the provider’s instructions.
The JC-1 red/green images were captured under a fluorescent micro­
scope. After different treatments, cells were incubated with 20 nM tet­
ramethyl rhodamine methyl ester (TMRM) (#T5428; Sigma Aldrich) in
Tyrode’s buffer for 45 min in dark for MMP assessment. After incuba­
tion, the cells were visualized with a fluorescence microscope. Fluo­
rescence intensity was measured with Image-J software (Version 1.52v,
Image J, National Institutes of Health, USA)).
2.11. Histological analysis
To explore histopathologic changes, the liver tissues were fixed with
10% neutral formalin, embedded in paraffin, and then sectioned trans­
versely (5-µm-thick). The liver tissue sections were then stained with
hematoxylin and eosin (H&E) to visualize the pattern of hepatic
inflammation. NAS score following H&E analysis was quantified as
previously described (McPherson et al., 2015). To further indicate lipid
accumulation in livers, the sections were stained with Oil Red O Stain Kit
(#ab150678, Abcam). After washing with 60% isopropyl alcohol, the
liver sections were re-stained with haematoxylin. Images were captured
under a light microscope.
2.8. Biochemical parameters determination
Assay kits for the measurements of malondialdehyde (MDA; #A0031-2), superoxide dismutase (SOD; #A001-3-2) and glutathione (GSH;
#A006-2-1) were obtained from Nanjing Jiancheng Bioengineering
Institute (Nanjing, China). ATP Assay Kit (#S0026, Beyotime) was used
to examine ATP contents in accordance with the provider’s instructions.
Caspase-3 Activity Assay Kit (#ab252897, Abcam) and Caspase 9 Ac­
tivity Assay Kit (#1158, Beyotime) were used to examine Caspase-3 and
Casaspe-9 activities, respectively, following the providers’ instructions.
The contents of cytokines in serum were measured using corresponding
commercial enzyme-linked immuno sorbent assay (ELISA) kits,
including the mouse TNF-α (#MTA00B), IL-1β (#MLB00C) and IL-6
(#M6000B) ELISA kits that were all from R&D system (USA) accord­
ing to the manufacturer’s protocols.
2.12. Immunofluorescence (IF) staining
For IF analysis, the cells after treatments were washed with PBS, and
were then blocked in 10% goat serum (#C0265, Beyotime) containing
0.3% Triton X-100 (#ST797, Beyotime) for 1 h at room temperature and
incubated with primary antibodies against Cyto-c (#PA5-19462, dilu­
tion 1:200) at 4 ◦ C overnight. Cells were then washed, and anti-rabbit
IgG H&L (Alexa Fluor® 594) (#ab150080) secondary fluorescent anti­
body (Abcam, dilution 1:300) were prepared for cell incubation at room
temperature in dark. After washing, 2-(4-Amidinophenyl)-6-indole­
carbamidine dihydrochloride solution (DAPI; #C1006, Beyotime) was
added to the cells for nuclei staining. Images were visualized and
captured under a fluorescence microscopy.
2.9. Real time-quantitative PCR (RT-qPCR)
2.13. Apoptosis analysis
Total RNA in cells or liver tissues were extracted using TRIzol™ re­
agent (#15596-018, Thermo Fisher Scientific) following the
Apoptosis of cells in the treated cells or the prepared liver sections
3
X. Dai et al.
Journal of Functional Foods 90 (2022) 104954
was measured by terminal-deoxynucleotidyl transferase-mediated nick
end labeling (TUNEL) staining (Beyotime) according to the manufac­
turer’s instructions. Representative images were captured under a
fluorescent microscope. The apoptosis was calculated as TUNEL-positive
cells (red)/DAPI (blue), and the number of TUNEL-positive staining cells
was quantified.
dysfunction in PA-incubated hepatocytes.
3.3. Fisetin mitigates mitochondrial apoptosis, inflammation and lipid
deposition in PA-incubated hepatocytes
Mitochondrial impairment contributes to apoptotic cell death under
different stimuli (Maharjan et al., 2014). Thereafter, apoptosis in PAtreated hepatocytes was examined. The Cyto-c liberation from mito­
chondria to cytoplasm/nucleus is reported as the molecular feature of
mitochondrial apoptosis (Chang, Xing, & Yu, 2014). As shown in Fig. 3A
and B, IF staining showed that nuclear Cyto-c was markedly promoted in
response to PA, whereas being significantly meliorated by FisT cotreatment. Furthermore, Caspase-9 and Caspase-3 activities were
considerably down-regulated by FisT in PA-stimulated L02 and AML12
cells (Fig. 3C and D). Western blotting results confirmed that Caspase-3
cleavage induced by PA was highly abrogated by FisT (Fig. 3E). TUNEL
staining subsequently indicated that PA stimulation clearly led to
apoptosis in hepatocytes, while being dramatically ameliorated after
FisT incubation (Fig. 3F and G). Additionally, RT-qPCR results indicated
that FisT treatment remarkably decreased the gene expression of proinflammatory cytokines including TNF-α, IL-1β and IL-6 in PA-treated
cells (Fig. 3H). ORO staining showed that PA-caused increases in lipid
accumulation and cellular TG contents were remarkably abolished upon
FisT co-incubation (Fig. 3I and J). Consistently, the mRNA expression
levels of fatty acid synthesis markers including FASN, SCD1 and PPARγ
were considerably upregulated by PA, which were, however, strongly
depressed by FisT (Fig. 3K). Together, these data suggested that FisT
could meliorate mitochondrial apoptosis, inflammation and lipid
deposition in PA-treated hepatocytes, contributing to NAFLD
amelioration.
2.14. Statistical analysis
Data represented as mean ± standard error of the mean (SEM) unless
otherwise indicated. All analysis were repeated independently with
similar results at least three times. Statistical analysis was conducted
using GraphPad Prism 8.0 (San Diego, CA, USA). Differences between
two groups were analyzed by Student’s t test. One-way analysis of
variance (ANOVA) with Tukey’s post hoc tests were performed for
comparisons between multiple groups. P value < 0.05 was considered
indicative of statistical significance.
3. Results
3.1. Fisetin restrains ROS production and improves Nrf2 activation in PAincubated hepatocytes
Oxidative stress is crucial for NAFLD progression, and the regulatory
role of FisT on ROS production has been reported though without fully
understanding (Friedman et al., 2018; Ge et al., 2019), and thus was
further explored. First, CCK-8 analysis suggested that FisT treatments at
different concentrations and time points showed no significant influence
on the changes of L02 and AML12 cell viability (Supplementary Fig. 1AC), indicating the non-cytotoxicity of FisT. DCF-DA staining subse­
quently indicated that FisT treatments significantly reduced ROS pro­
duction in human and murine hepatocyte lines L02 and AML12 via a
dose-dependent manner (Fig. 1A and B). Given that 20 μM of FisT
exhibited the most efficient capacity to restrain ROS generation, and
thereafter was chosen for subsequent in vitro analysis. Additionally, we
found that 20 μM of FisT treatment for 24 or 48 h did not influence the
morphology of L02 and AML12 cells (Supplementary Fig. 1D), con­
firming the safe use of FisT. The contents of oxidative stress marker MDA
were highly increased in PA-incubated L02 and AML12 cells, while
being greatly ameliorated upon FisT exposure (Fig. 1C). On the contrary,
the activities of antioxidant enzyme including SOD, CAT and GSH were
significantly downregulated in L02 and AML12 cells after PA stimula­
tion, which were, however, remarkably rescued in hepatocytes cotreated with FisT (Fig. 1D-F). NADPH oxidases (NOXs) including
NOX1, NOX2 and NOX4 that contribute to ROS production were
strongly upregulated in PA-incubated cells, whereas being efficiently
abolished by FisT (Fig. 1G). However, FisT remarkably restored the
expression of antioxidants HO-1, NQO1 and GCLM in L02 and AML12
cells after PA stimulation (Fig. 1H). As expected, both nuclear and total
Nrf2 protein expression levels were evidently decreased in PA-treated
hepatocytes, while being considerably restored by FisT (Fig. 1I and J).
Together, these findings confirmed that FisT exerted antioxidant
bioactivity in PA-treated hepatocytes.
3.4. Fisetin represses apoptosis, lipid accumulation and inflammatory
response via ROS suppression in PA-stimulated hepatocytes
Due to the crucial role of ROS production in mediating cell death,
inflammatory response and lipid metabolism (Oyinloye, Adenowo, &
Kappo, 2015), deeper molecular mechanism by which FisT regulated
these cellular events was then explored. Nrf2 increase plays a key role in
the suppression of ROS generation (De Vries et al., 2008), and was then
knocked down in L02 and AML12 cells by transfecting with the con­
structed si-Nrf2. RT-qPCR and western blotting results confirmed the
successful Nrf2 knockdown in cells, particularly si-Nrf2-2# (Supple­
mentary Fig. 2A and B), which was thus selected for subsequent analysis.
As expected, DCF-DA staining showed that FisT-restrained ROS gener­
ation was significantly diminished upon Nrf2 deletion in PA-treated
hepatocytes (Fig. 4A). TUNEL staining demonstrated that Nrf2 silence
significantly abolished the function of FisT to reduce apoptotic cell
death in PA-stimulated cells (Fig. 4B). Similarly, lipid accumulation and
cellular TG levels restrained by FisT in PA-exposed cells were markedly
abrogated when Nrf2 was deleted (Fig. 4C and D). What’s more,
Caspase-3 activity, the gene expression levels of fatty acid synthesis
markers and pro-inflammatory cytokines downregulated by FisT were
markedly restored upon Nrf2 knockdown in PA-exposed hepatocytes
(Fig. 4E-G). Together, these results indicated that FisT-meliorated
apoptosis, lipid deposition and inflammatory response induced by PA
were largely associated with the suppression of ROS regulated by Nrf2.
3.2. Fisetin improves mitochondrial dysfunction and impairment in PAtreated hepatocytes
3.5. Fisetin-mediated cellular events is mainly through ER stress inhibition
In this regard, MitoSOX staining further showed that besides cellular
ROS, FisT incubation also remarkably decreased the generation of
mitochondrial ROS in PA-stimulated L02 and AML12 cells (Fig. 2A).
Both TMRM and JC-1 staining suggested that PA exposure evidently
resulted in MMP impairments in hepatocytes, and these events were
significantly mitigated upon FisT co-incubation (Fig. 2B-D). Consis­
tently, ATP levels downregulated by PA were also strongly increased in
L02 and AML12 cells after FisT treatment (Fig. 2E). These findings
illustrated that FisT could meliorate mitochondrial impairment and
In this part, we found that the mRNA and protein expression levels of
ER stress hallmarks GRP78 and CHOP were significantly promoted in
PA-exposed hepatocytes, while being greatly meliorated upon FisT coincubation (Fig. 5A and B). ER stress plays an essential role in regu­
lating several cellular events, including ROS generation, cell death and
inflammatory response under different stresses (Chaudhari et al., 2014).
To further explore the underlying mechanisms by which FisT mediated
4
X. Dai et al.
Journal of Functional Foods 90 (2022) 104954
Fig. 1. Fisetin restrains ROS production and improves Nrf2 activation in PA-incubated hepatocytes. (A&B) DCF-DA staining to examine ROS production in PA
(250 μM)-incubated L02 and AML12 cells in the presence or absence of FisT at the shown concentrations for 24 h. Scale bar was 75 μm. (C-J) L02 and AML12 cells
were treated with PA (250 μM) for 24 h with or without FisT (20 μM). Then, all cells were collected for studies as follows. Quantification for (C) MDA levels, (D) SOD
activities, (E) CAT activities and (F) GSH levels. (G) RT-qPCR results for oxidative stress markers including NOX1, NOX2, and NOX4. (H) RT-qPCR analysis for the
gene expression levels of antioxidants HO-1, NQO1 and GCLM in cells. (I&J) Western blotting results for nuclear and total Nrf2 protein expression levels. Data were
expressed as the mean ± SEM (n = 4 per group). *p < 0.05, **p < 0.01 and ***p < 0.001 versus the Con/Veh group; +p < 0.05 and ++p < 0.01 versus the PA/
Veh group.
5
X. Dai et al.
Journal of Functional Foods 90 (2022) 104954
Fig. 2. Fisetin improves mitochondrial dysfunction and impairment in PA-treated hepatocytes. After incubation with PA (250 μM) for 24 h in the absence or
presence of FisT (20 μM), L02 and AML12 cells were harvested for the following studies. (A) Mitochondrial ROS production was measured using MitoSOX staining.
Scale bar was 75 μm. (B) TMRM staining of cells to monitor mitochondrial function. Scale bar was 75 μm. (C&D) MMP examination using JC-1 staining. Scale bar was
75 μm. (E) Assessments of cellular ATP levels. Data were expressed as the mean ± SEM (n = 4 per group). *p < 0.05 and **p < 0.01 versus the Con/Veh group;
+
p < 0.05 and ++p < 0.01 versus the PA/Veh group.
NAFLD in vitro, GRP78 was then silenced or promoted by transfection
with GRP78 siRNAs or plasmids, respectively. Transfection efficacy was
validated by RT-qPCR analysis (Supplementary Fig. 3A and B). Because
the most inhibitory effect of si-GRP78-1# on GRP78 expression, and
thus was used for the following studies. Firstly, DCF-DA and MitoSOX
staining indicated that GRP78 knockdown markedly reduced cellular
and mitochondrial ROS generation in PA-incubated hepatocytes.
Surprisingly, FisT-restrained ROS production was significantly restored
by GRP78 over-expression in PA-treated cells (Fig. 5C-F). Consistently,
ER stress suppression by si-GRP78 markedly improved MMP in L02 and
AML12 cells with PA stimulation; however, facilitating GRP78 expres­
sion remarkably abrogated the function of FisT to promote MMP in PAincubated hepatocytes (Fig. 5G). The effects of ER stress on the media­
tion of lipid metabolism have been increasingly demonstrated (Basseri &
6
X. Dai et al.
Journal of Functional Foods 90 (2022) 104954
Fig. 3. Fisetin mitigates mitochondrial apoptosis, inflammation and lipid deposition in PA-incubated hepatocytes. L02 and AML12 cells were exposed to PA
(250 μM) for 24 h with or without FisT (20 μM) treatment and were then harvested for studies as follows. (A&B) IF staining for Cyto-c expression in cells. Scale bar
was 25 μm. Measurements of (C) Caspase-9 and (D) Caspase-3 activity in cells. (E) Western blotting results for cleaved Caspase-3 protein expression levels in cells.
(F&G) TUNEL staining for apoptotic cell death in cells. Scale bar was 50 μm. (H) RT-qPCR analysis for inflammatory cytokines including TNF-α, IL-1β and IL-6 in
cells. (I) ORO staining to monitor lipid deposition in cells. (J) Cellular TG contents were examined. (K) RT-qPCR analysis for FASN, SCD1 and PPARγ mRNA
expression levels in cells. Data were expressed as the mean ± SEM (n = 4 per group). **p < 0.01 and ***p < 0.001 versus the Con/Veh group; +p < 0.05 and
++
p < 0.01 versus the PA/Veh group.
Austin, 2012; Lebeaupin et al., 2018). We then attempted to explore
whether GRP78-regulated ER stress was involved in FisT-ameliorated
lipid deposition under metabolic stresses. As expected, both FisT and
GRP78 knockdown significantly reduced lipid accumulation in PAstimulated L02 and AML12 cells; however, promoting GRP78 expres­
sion completely abolished the function of FisT to restrain cellular lipid
deposition after PA exposure (Supplementary Fig. 4A-C). Consistently,
in FisT-treated hepatocytes, the mRNA expression levels of FASN, SCD1
and PPARγ were almost re-strengthened by GRP78 over-expression after
PA stimulation (Supplementary Fig. 4D). Collectively, these results
revealed that FisT could repress PA-induced ER stress through inhibiting
GRP78/CHOP signaling, contributing to the suppression of ROS pro­
duction, mitochondrial dysfunction and lipid deposition.
3.6. Effects of fisetin on the crosstalk between Nrf2 and GRP78 signaling
in PA-stimulated hepatocytes
Cross talk between oxidative stress and ER stress by Nrf2 and GRP78
signaling pathways under physiological and pathological conditions has
been widely demonstrated (Ge et al., 2019; Xu et al., 2020; Zhang et al.,
2019). As shown in Fig. 6A and B, we found that FisT-reduced gene and
protein expression levels of GRP78 a CHOP were significantly abrogated
upon Nrf2 silence both in L02 and AML12 cells under PA stimulation.
However, under normal status, Nrf2 knockdown showed no significant
influence on GRP78 and CHOP expression changes. These findings
suggested that Nrf2 might not directly interact with GRP78, but could
indeed affect GRP78-mediated ER stress in FisT-treated hepatocytes
under metabolic pressure. Furthermore, we found that in hepatocytes
without PA treatment, GRP78 knockdown or over-expression had no
significant influences on Nrf2 activation, confirming that there might
not be a direct correlation between Nrf2 and GRP78. Under PA7
X. Dai et al.
Journal of Functional Foods 90 (2022) 104954
Fig. 4. Fisetin represses apoptosis, lipid accumulation and inflammatory response via ROS suppression in PA-stimulated hepatocytes. L02 and AML12 cells
were transfected with si-Nrf2 for 24 h, and were then incubated with PA (250 μM) for another 24 h in the presence or absence of FisT (20 μM). Subsequently, all cells
were collected for the following studies. (A) ROS production was examined using DCF-DA staining. Scale bar was 75 μm. (B) TUNEL staining was used to monitor
apoptosis. Scale bar was 50 μm. (C) ORO staining to investigate lipid accumulation. Scale bar was 40 μm. (D) Quantification for cellular TG contents. (E) Caspase-3
activity was measured. The gene expression levels of (F) FASN, SCD1, PPARγ, (G) TNF-α, IL-1β and IL-6 in cells were quantified by RT-qPCR analysis. Data were
expressed as the mean ± SEM (n = 4 per group). **p < 0.01 and ***p < 0.001 versus the Con/Veh group; +p < 0.05 and ++p < 0.01 versus the PA/Veh group;
#
p < 0.05 and ##p < 0.01 versus the PA/FisT group.
stimulated status, similar with FisT biological function, si-GRP78
significantly rescued nuclei and total Nrf2 expression levels; however,
FisT-improved Nrf2 expression was almost abolished upon GRP78 overexpression in PA-stimulated hepatocytes (Fig. 6C). These findings
further suggested that under metabolic stresses, there may be a crosstalk
between oxidative stress and ER stress mediated by Nrf2 and GRP78
signaling pathways, respectively, which is a key metabolism for FisT to
perform its protective function against NAFLD.
3.7. Fisetin ameliorates NAFLD development in HFD-fed mice
Animal studies were then performed using a 16-week HFD-induced
8
X. Dai et al.
Journal of Functional Foods 90 (2022) 104954
Fig. 5. Fisetin-mediated cellular events is mainly through ER stress inhibition. (A) Rt-qPCR and (B) western blotting analysis for GRP78 and CHOP gene and
protein expression levels in L02 and AML12 cells after 24 h incubation with PA (250 μM) in the presence or absence of FisT (20 μM). (C-G) L02 and AML12 cells were
transfected with si-GRP78 or GRP78 plasmids for 24 h to reduce or promote GRP78 expression, respectively, and were then exposed to PA (250 μM) for another 24 h
with or without FisT (20 μM) co-incubation. Finally, all cells were collected for studies as follows. (C) DCF-DA and (D) MitoSOX staining for cellular and mito­
chondrial ROS production, respectively. Scale bar was 75 μm. Quantification for (E) cellular ROS and (F) mitochondrial ROS was shown. (G) MMP of cells was
assessed using JC-1 staining. Scale bar was 75 μm. Data were expressed as the mean ± SEM (n = 4 per group). **p < 0.01 and ***p < 0.001 versus the Con/Veh group;
++
p < 0.01 versus the PA/Veh group; #p < 0.05 versus the PA/FisT group.
9
X. Dai et al.
Journal of Functional Foods 90 (2022) 104954
Fig. 6. Effects of fisetin on the crosstalk between Nrf2 and GRP78 signaling in PA-stimulated hepatocytes. (A&B) L02 and AML12 cells were transfected with
si-Nrf2 for 24 h, and were then exposed to PA (250 μM) for an additional 24 h with or without FisT (20 μM). Next, all cells were harvested for RT-qPCR and western
blotting assays of GRP78 and CHOP. (C) L02 and AML12 cells were transfected with si-GRP78 or GRP78 plasmids. After 24 h, cells were subjected to PA (250 μM) for
another 24 h with or without FisT (20 μM) co-incubation, followed by western blotting analysis of nuclei and total Nrf2 protein expression levels. Data were
expressed as the mean ± SEM (n = 4 per group). **p < 0.01 and ***p < 0.001 versus the Con/Veh group; +p < 0.05 and ++p < 0.01 versus the PA/Veh group;
#
p < 0.05 and ##p < 0.01 versus the PA/FisT group; ns, no significant difference.
NAFLD, and FisT was subjected to mice for a total of 16 weeks to confirm
its inhibitory potential on fatty liver. At first, we found that HFD feedingincreased body weights of mice were remarkably decreased by FisT
administration, along with significantly reduced fasting blood glucose
and insulin levels (Fig. 6A-C), indicating the effects of FisT to improve
metabolic disorders in vivo. Additionally, higher LW/BW was observed
in HFD-challenged mice, while being greatly meliorated by FisT
(Fig. 6D). Hepatic dysfunction triggered by HFD was also considerably
alleviated by FisT, as evidenced by the decreased serum ALT and AST
levels (Fig. 6E). Histological staining by H&E and ORO showed that FisT
treatment markedly reduced the NAFLD characteristics such as inflam­
matory cell infiltration and lipid deposition in HFD-fed mice, along with
decreased NAS scores (Fig. 6F and G). Moreover, TC and TG contents in
serum and liver were strongly upregulated by HFD feeding, while being
significantly meliorated after FisT treatment (Fig. 6H and I). We also
found that serum and hepatic pro-inflammatory cytokines including
TNF-α, IL-1β and IL-6 enhanced by HFD were dramatically restrained by
FisT co-administration (Fig. 6J and K). As expected, hepatic FASN, SCD1
and PPARγ gene expression levels boosted by HFD feeding were highly
diminished in mice co-treated with FisT (Fig. 6L). These data demon­
strated that FisT could ameliorate NAFLD characteristics in HFD-fed
mice.
10
X. Dai et al.
Journal of Functional Foods 90 (2022) 104954
3.8. Fisetin suppresses oxidative stress and mitochondrial apoptosis in
liver of HFD-challenged NAFLD mice
However, antioxidants HO-1, NQO1 and GCLM mRNA levels were
considerably rescued in liver of HFD-challenged mice with FisT
administration (Fig. 7G). Consistent with in vitro results, nuclear and
total Nrf2 protein expression levels were significantly decreased in liver
of HFD-fed mice, while being dramatically restored by FisT (Fig. 7H).
TUNEL staining indicated that FisT administration remarkably
decreased the number of apoptotic cells in liver of HFD-challenged mice
(Fig. 7I and J), accompanied by reduced expression of cleaved Caspase-9
and Caspase-3 (Fig. 7K). These data disclosed that FisT could restrain
hepatic ROS and apoptosis in HFD-fed mice.
In this section, DHE staining in Fig. 7A and B showed that HFD
feeding led to ROS production in liver sections of mice, while being
significantly ameliorated by FisT. Hepatic MDA levels elevated by HFD
were also abolished in mice co-treated with FisT (Fig. 7C). On the con­
trary, HFD-reduced SOD activity and GSH levels in liver tissues were
markedly upregulated after FisT administration (Fig. 7D and E). As ex­
pected, liver NOX1, NOX2 and NOX4 gene expression levels enhanced
by HFD were strongly downregulated by FisT treatment (Fig. 7F).
Fig. 7. Fisetin ameliorates NAFLD development in HFD-fed mice. (A) Initial and final body weights of mice. (B) Fasting blood glucose and (C) insulin levels were
measured. (D) Quantification for LW/BW. (E) Serum ALT and AST levels were examined. (F) H&E and ORO staining of liver sections from the shown groups of mice.
Scale bar was 200 μm. (G) NAS score was quantified. (H) Serum TC and TG levels were tested. (I) Liver TG and TC contents were examined. (J) Serum TNF-α, IL-1β
and IL-6 levels were examined using ELISA analysis. RT-qPCR analysis for (K) TNF-α, IL-1β, IL-6, (L) FASN, SCD1 and PPARγ gene expression levels in liver samples.
Data were expressed as the mean ± SEM (n = 5 or 8 per group). **p < 0.01 and ***p < 0.001 versus the NCD/Veh group; +p < 0.05 and ++p < 0.01 versus the HFD/
Veh group.
11
X. Dai et al.
Journal of Functional Foods 90 (2022) 104954
3.9. Fisetin prohibits ER stress in liver of HFD-fed mice
hepatocyte damage and immune cell activation (Ghemrawi, BattagliaHsu, & Arnold, 2018; Zhang et al., 2014). Approaches to suppress
mitochondrial dysfunction and ER stress are likely to be effective for
NAFLD treatment. FisT is a dietary flavonoid and can be found in many
fruits and vegetables. FisT could prohibit inflammatory response to
ameliorate fatty liver and nephropathy progression induced by HFD
through various signaling pathways (Chenxu et al., 2021; Xu, Ge, & Qin
et al., 2019; Xu, Sun, & Dai et al., 2019). The regulatory effects of FisT to
mediate ER stress and mitochondrial dysfunction have also been re­
ported to control the progression of cardiovascular diseases and
different types of tumors (Althunibat et al., 2019; Sabarwal, Agarwal, &
Singh, 2017). However, whether these mechanisms, especially the
possible interrelation between them can be regulated by FisT to effec­
tively ameliorate NAFLD, it’s not been fully investigated.
In the present study, we provided deeper insights into the thera­
peutic potential of FisT on NAFLD progression, particularly the under­
lying mechanisms. Our in vitro studies showed that FisT markedly
reduced cellular ROS production in PA-incubated hepatocytes through
increasing Nrf2 expression levels. In addition, mitochondrial ROS,
impairment and dysfunction triggered by PA were also strongly miti­
gated by FisT. Furthermore, PA exposure led to mitochondrial apoptosis,
inflammatory response and lipid deposition in L02 and AML12 cells;
however, these events were remarkably ameliorated after FisT treat­
ment. Importantly, we found that promoting ROS by Nrf2 knockdown
strongly abolished the anti-apoptotic, anti-inflammatory and antilipogenic biological functions of FisT in PA-treated hepatocytes. Addi­
tionally, FisT significantly suppressed ER stress induced by PA in he­
patocytes by decreasing GRP78 and CHOP expression. Notably, we
showed that the capacity of FisT to prohibit ROS production, mito­
chondrial damage and lipid deposition were dramatically abrogated
upon ER stress facilitation induced by GRP78 over-expression. We
further found that under PA-stimulated status, Nrf2 silence abolished
the inhibitory effects of FisT on GRP78 expression. Meanwhile, pro­
moting GRP78 markedly eliminated the function of FisT to improve Nrf2
in PA-treated hepatocytes. These results demonstrated that crosstalk
between oxidative stress and ER stress mediated by Nrf2 and GRP78
signaling pathways, respectively, might be involved in FisT-ameliorated
fatty liver progression (Fig. 9C). Our in vivo experiments finally sup­
ported that FisT administration both for 16 weeks and 8 weeks could
efficiently mitigate NAFLD pathologies in HFD-fed mice through the
molecular mechanisms as we observed in vitro. Collectively, all our
studies confirmed that FisT exerted suppressive effects on GRP78mediated ER stress, contributing to the inhibition of ROS production
and mitochondrial dysfunction, which ameliorated NAFLD development
ultimately.
Oxidative stress is a key factor involved in NAFLD pathogenesis.
•
During oxidative stress, excessive ROS production, including H2O2, O2 −
−
and ONOO , leads to damage to virtually any cellular component, and
elevated ROS has been detected in patients with NAFLD (Jianhua et al.,
2021). NOXs are major source of ROS production and have the precise
•
role of directly catalyzing O2 − production from O2 by transferring
electrons from NADPH across biological membranes. NOX-produced
ROS in particular are toxic in non-neoplastic conditions, since they are
associated with the establishment of an inflammatory vicious cycle
determining and sustaining the transition of NAFLD to NASH (Chen
et al., 2020; Gabbia, Cannella, & De Martin, 2021). Nrf2 is served as a
master redox switch that turns on the expression of endogenous anti­
oxidant genes and ultimately promotes the cellular redox potential. Cells
adapt to ROS through this potent feedback mechanism, strengthening
the protection against oxidative damage (Oyinloye, Adenowo, & Kappo,
2015). The effects of FisT to restrain oxidative stress and ROS production
through improving Nrf2 signaling pathway have been demonstrated in
metabolic stress-induced cardiac injury (Ge et al., 2019). Here, we
confirmed that PA caused aberrant ROS production in hepatocytes by
decreasing Nrf2 nuclear expression, along with markedly increased
expression of NOX1, NOX2 and NOX4. Meanwhile, antioxidants such as
Finally, ER stress stimulated by HFD was also detected in liver of
mice, as proved by the markedly increased expression of GRP78 and
CHOP both from mRNA and protein levels; however, these effects were
considerably abolished after FisT administration (Fig. 8A and B). These
findings indicated the suppressive effects of FisT on ER stress, contrib­
uting to the inhibition of NAFLD development.
3.10. Therapeutic efficiency examination of 8-week fisetin administration
in NAFLD mice after a 16-week HFD feeding
To further examine the therapeutic potential of FisT on NAFLD
context, another animal experiment design was performed in 16 weeks
of HFD-challenged mice with FisT supplementation for 8 weeks (Sup­
plementary Fig. 5A). We found that HFD-increased body weights of mice
were also significantly reduced after an 8-week FisT therapy (Supple­
mentary Fig. 5B). Serum fasting blood glucose and insulin levels induced
by HFD were strongly ameliorated by an 8-week FisT supplementation
(Supplementary Fig. 5C and D). HFD-triggered higher LW/BW was also
significantly alleviated after FisT treatment for 8 weeks (Supplementary
Fig. 5E). Moreover, HFD-induced mice suffered higher levels of serum
ALT and AST, and treatment of 8-week FisT remarkably decreased ALT
and AST contents in serum (Supplementary Fig. 5F). H&E and ORO
staining indicated that 8-week FisT effectively alleviated NAFLD
symptoms, reflected by the descent of lipid accumulation and inflam­
matory cells infiltration in hepatic sections, accompanied by the
decreased NAS scores (Supplementary Fig. 5G and H). After HFD
feeding, higher TC and TG contents both in serum and liver were
significantly mitigated after FisT administration for 8 weeks (Supple­
mentary Fig. 5I and J). Furthermore, FisT treatment remarkably
decreased sera and hepatic TNF-α, IL-1β and IL-6 levels in HFD-fed mice
(Supplementary Fig. 5K and L). Liver FASN, SCD1 and PPARγ gene
expression levels were also downregulated by 8-week FisT consumption
in HFD-treated mice (Supplementary Fig. 5M). These findings suggested
that an 8-week FisT therapeutic approach could ameliorate NAFLD
progression in HFD-fed mice.
DHE staining further showed that FisT administration for 8 weeks
had antioxidant functions in liver of NAFLD mice, proved by the
evidently weakened fluorescent intensity, which was along with the
decreased MDA contents (Supplementary Fig. 6A-C). However, mark­
edly rescued SOD and GSH activities were detected in liver of HFD-fed
mice with FisT treatment for 8 weeks (Supplementary Fig. 6D and E).
Hepatic NOX1, NOX2 and NOX4 mRNA expression levels provoked by
HFD were highly diminished in mice after an 8-week FisT administration
(Supplementary Fig. 6F), accompanied with improved nuclei and total
Nrf2 protein expression levels (Supplementary Fig. 6G). Western blot­
ting results confirmed that 8-week FisT supplementation remarkably
downregulated Caspase-9 and Caspase-3 cleavage in liver of NAFLD
mice (Supplementary Fig. 6H). Consistently, GRP78 and CHOP expres­
sion levels induced by HFD were strongly decreased in HFD-fed mice
after FisT treatment for 8 weeks (Supplementary Fig. 6I and J). These
findings supported that FisT exerted therapeutic potential against HFDtriggered NAFLD by suppressing oxidative stress and ER stress.
4. Discussion
NAFLD includes a broad spectrum of hepatic dysfunctions, and it is
predicted to become the primary cause of liver failure and even hepa­
tocellular carcinoma (HCC) (Im et al., 2021; Zhou et al., 2020). Mito­
chondria is a type of highly dynamic organelles associated with multiple
metabolic/bioenergetic pathways in liver (Fanta et al., 2021). During
NAFLD progression, mitochondrial dysfunction may precede insulin
resistance and participate to fat accumulation (Eduardo et al., 2021; Lee,
Park, & Roh, 2019; Sanda et al., 2021). ER stress is also attributed to
multi-hits of NAFLD development in various different aspects such as
12
X. Dai et al.
Journal of Functional Foods 90 (2022) 104954
Fig. 8. Fisetin suppresses oxidative stress and mitochondrial apoptosis in liver of HFD-challenged NAFLD mice. (A&B) DHE staining for ROS production in
liver samples. Scale bar was 200 μm. (C) MDA levels, (D) SOD activity and (E) GSH contents in liver of mice were measured. (F) RT-qPCR analysis for NOX1, NOX2
and NOX4 gene expression levels in liver tissues of mice. (G) RT-qPCR results for HO-1, NQO1 and GCLM mRNA levels in liver samples. (H) Nuclear and total Nrf2
protein expression levels measured by western blotting analysis. (I&J) TUNEL staining for apoptosis in hepatic sections. Scale bar was 200 μm. (K) Cleaved Caspase-9
and Caspase-3 protein expression levels in liver samples by western blotting assays. Data were expressed as the mean ± SEM (n = 5 per group). *p < 0.05, **p < 0.01
and ***p < 0.001 versus the NCD/Veh group; +p < 0.05 and ++p < 0.01 versus the HFD/Veh group.
13
X. Dai et al.
Journal of Functional Foods 90 (2022) 104954
Fig. 9. Fisetin prohibits ER stress in liver of HFD-fed mice. (A) RT-qPCR and (B) western blotting analysis for hepatic GRP78 and CHOP gene and protein
expression levels, respectively. (C) Working model illustrating the regulatory role of FisT on NAFLD progression through the suppression of GRP78-mediated ER
stress to subsequently inhibit ROS accumulation and mitochondrial dysfunction that contributed to inflammation, lipid deposition and apoptosis. Data were
expressed as the mean ± SEM (n = 5 per group). ***p < 0.001 versus the NCD/Veh group; ++p < 0.001 and +++p < 0.001 versus the HFD/Veh group.
HO-1, NQO1 and GCLM were also found to be decreased in PAincubated hepatocytes, causing oxidative damage to cells. However,
these effects were consistently mitigated by FisT through enforcing Nrf2
nuclear translocation. Excessive ROS production can cause lipid accu­
mulation, inflammation, and cell apoptosis (Oyinloye, Adenowo, &
Kappo, 2015; Turillazzi et al., 2016). FisT has been shown to restrain
inflammatory response and dyslipidemia under metabolic stresses (Ge
et al., 2019; Xu, Ge, & Qin et al., 2019; Xu, Sun, & Dai et al., 2019). In
our present study, we confirmed that FisT exposure could suppress
inflammation and lipid deposition in PA-stimulated hepatocytes.
Intriguingly, we found that promoting ROS by Nrf2 deletion signifi­
cantly abolished the effects of FisT to suppress PA-induced inflammation
and lipid accumulation in L02 and AML12 cells. Therefore, we
concluded that the inhibitory effects of FisT on inflammation and lipid
deposition caused by metabolic stress were at least in part attributed to
ROS suppression. These events mediated by FisT were validated in liver
of NAFLD mice induced by HFD.
Mitochondria are considered as primary contributors to oxidative
stress since the mitochondrial single-electron transport chain can
generate most of ROS (Fanta et al., 2021). Fatty liver occurs when fatty
acid availability exceeds the demand for ATP in hepatocytes (Zhang
et al., 2011). In addition, the reduction of mitochondrial ATPsynthesizing respiration has been observed in murine model with sim­
ple steatosis, indicating that fat accumulates in liver because the mito­
chondria fail to oxidize enough fatty acids (Wang et al., 2009).
Mitochondrial dysfunction leads to a vicious cycle of hepatocyte damage
with ROS over-production, releases of inflammatory factors and cell
death (Eduardo et al., 2021; Lee, Park, & Roh, 2019; Sanda et al., 2021;
Wei et al., 2008). Consistently, here in our study, we also found that PA
stimulation led to mitochondrial ROS generation and impaired MMP,
contributing to the dysfunction of mitochondria in hepatocytes. More­
over, TUNEL staining indicated that PA exposure caused severe
apoptosis in hepatocytes. Cyto-c locates in mitochondria under normal
conditions; however, in response to stimuli such as PA, Cyto-c will
release from mitochondria into nuclear or cytoplasm, thereafter,
inducing Caspase-9 and Caspase-3 cleavage to trigger mitochondrial
apoptosis (Brentnall et al., 2013). In our study, we further found that PA
treatment markedly increased nuclear Cyto-c expression, along with
elevated activation of Caspase-9 and Casapse-3. Notably, all these pro­
cesses including mitochondrial ROS production, dysfunction and
apoptosis were dramatically abrogated by FisT. Our in vivo experiments
confirmed that FisT administration restrained hepatic apoptosis and
Caspase-9/-3 activation in HFD-induced NAFLD mice. Therefore, we
concluded that FisT could ameliorate fatty liver progression partially
through improving mitochondrial function.
The ER is an intracellular membranous organelle and responsible for
numerous critical cellular events such as protein synthesis, lipid syn­
thesis, carbohydrate metabolism, oxidative damage and cell death
(Ghemrawi, Battaglia-Hsu, & Arnold, 2018; Iurlaro & Muñoz-Pinedo,
2016; Marciniak & Ron, 2006; Zhang & Kaufman, 2008). ER in hepa­
tocytes has an important capacity to adapt to extracellular and intra­
cellular alterations, which ensure that normal and crucial hepatic
metabolic functions are preserved. ER stress is clinically associated with
NAFLD progression (Gonzalez-Rodriguez et al., 2014; Lebeaupin et al.,
2015). It has been demonstrated that experimental ER stress inducers
affect hepatocyte survival and the intrinsic functions of hepatocytes,
including lipogenesis and inflammatory response (Parafati et al., 2018).
GRP78, a molecular Chaperone, monitors ER protein activities. Under
resting status, GRP78 binds to the luminal part of ER membrane pro­
teins. However, under stimulated conditions, GRP78 will be released
from ER membrane proteins, enabling the homo-oligomerization of
protein kinase RNA-like ER kinase (PERK), inositol-requiring enzyme 1
(IRE1) or both and resulting in the autophosphorylation of these en­
zymes and the activation of their substrates, such as CHOP (Lee, 2005).
CHOP plays an essential role in ER stress-induced apoptosis, including
mitochondrial apoptosis (Hu et al., 2019). In a study with a small cohort
14
X. Dai et al.
Journal of Functional Foods 90 (2022) 104954
of patients, GRP78 and CHOP mRNA and protein levels were markedly
upregulated in liver of patients with steatohepatitis (Gonzalez-Rodri­
guez et al., 2014; Lebeaupin et al., 2015; Toriguchi et al., 2014). Simi­
larly, in our present study, we also found that PA exposure led to
significant increases in GRP78 and CHOP expression both from mRNA
and protein levels, which were, however, remarkably downregulated in
hepatocytes with FisT treatment, indicating that FisT could suppress ER
stress caused by metabolic stimuli. Intriguingly, we found that pro­
moting ER stress by GRP78 over-expression remarkably abolished the
capacities of FisT to suppress ROS production and mitochondrial
impairment in PA-treated hepatocytes. Increasing studies have intro­
duced the crosstalk between ER stress and oxidative stress under
numerous physiological conditions (Xu et al., 2020; Zhang et al., 2019).
Here, our in vitro studies showed that GRP78 knockdown upregulated
Nrf2 nuclei expression in PA-incubated hepatocytes, while its overexpression significantly diminished the capacity of FisT to improve
Nrf2 signaling. Nrf2, as a central role orchestrating the antioxidant
response, is a direct substrate of PERK and an effector of PERKdependent cell survival (Cullinan et al., 2003; Lebeaupin et al., 2018),
which explained why the GRP78 knockdown cells expressed a large
amount of nuclear Nrf2, and over-expression hepatocytes showed
evidently decreased Nrf2 expression. There are also studies suggesting
that Nrf2-mediated antioxidant signaling may contribute to the inhibi­
tion of ER stress by restraining GRP78 expression, exerting protective
effects against cell death and metabolic disorders (Niazpour et al., 2021;
Xu et al., 2020). Similarly, here we also found that Nrf2-silent hepato­
cytes abrogated the inhibitory effect of FisT on GRP78 signaling,
exhibiting significantly elevated GRP78 and CHOP expression after PA
stimulation. However, under normal status without PA exposure,
neither GRP78 nor Nrf2 knockdown had no significant influences on the
expression changes of Nrf2 and GRP78, respectively, revealing that
there may be no direct interaction between the two signals. These in vitro
findings suggested that the crosstalk between Nrf2-regulated oxidative
stress and GRP78-mediated ER stress under pathological conditions was
involved in FisT-ameliorated fatty liver context (Fig. 9C). Finally, the
inhibitory and therapeutic potential of FisT to prohibit ER stress in liver
were confirmed in HFD-fed mice with NAFLD, contributing to the
amelioration of fatty liver.
In summary, our results demonstrated that FisT significantly
restrained cellular and mitochondrial ROS production through
improving Nrf2 in PA-incubated hepatocytes, subsequently contributing
to the inhibition of inflammatory response, lipid deposition and
apoptotic cell death. Moreover, ER stress stimulated by PA was also
highly prohibited by FisT in hepatocytes. Notably, FisT-inhibited ROS
generation and mitochondrial dysfunction were largely attributed to
GRP78-mediated ER stress inhibition. These cellular events mediated by
FisT were also detected in HFD-induced murine model with NAFLD
(Fig. 9C). Therefore, we provided deeper and further insights for the
molecular mechanisms of FisT in NAFLD treatment. Herein, FisT is ex­
pected to become a promising drug for the management of NAFLD.
CRediT authorship contribution statement
Xianling Dai: Conceptualization, Methodology, Software, Visuali­
zation, Investigation. Qin Kuang: Conceptualization, Methodology,
Software, Data curation, Writing – original draft, Visualization, Inves­
tigation. Yan Sun: Conceptualization, Methodology, Software, Data
curation, Writing – original draft, Visualization, Investigation. Minxuan
Xu: Data curation, Writing – original draft. Liancai Zhu: Data curation,
Writing – original draft, Supervision, Software, Validation, Writing –
review & editing. Chenxu Ge: Conceptualization, Methodology, Soft­
ware, Data curation, Writing – original draft, Supervision, Software,
Validation, Writing – review & editing. Jun Tan: Conceptualization,
Methodology, Software, Supervision, Software, Validation, Writing –
review & editing. Bochu Wang: Conceptualization, Methodology,
Software, Supervision, Software, Validation, Writing – review & editing.
Declaration of Competing Interest
The authors declare that they have no known competing financial
interests or personal relationships that could have appeared to influence
the work reported in this paper.
Acknowledgement
This work was supported by (1) Graduate Student Research Inno­
vation Project of Chongqing University (grant number: CYS21070); (2)
Chinese Medicine Technology Innovation and Application Development
Project (grant number: 2020ZY013802) funded by Chongqing Munic­
ipal Health Commission and Chongqing Science and Technology Bu­
reau; (3) Chongqing Professional Talents Plan for Innovation and
Entrepreneurship
Demonstration
Team
(grant
number:
CQCY201903258, cstc2021ycjh-bgzxm0202); (4) Advanced Programs
of Post-doctor of Chongqing (grant number: 2017LY39); (5) Supported
by Youth Project of Science and Technology Research Program of
Chongqing Education Commission of China (grant number:
KJQN201901606) and (6) Chongqing Research Program of Basic
Research and Frontier Technology(grant number: cstc2018jcyjAX0393).
Appendix A. Supplementary material
Supplementary data to this article can be found online at https://doi.
org/10.1016/j.jff.2022.104954.
References
Althunibat, O. Y., Al Hroob, A. M., Abukhalil, M. H., et al. (2019). Fisetin ameliorates
oxidative stress, inflammation and apoptosis in diabetic cardiomyopathy. Life
Sciences, 221, 83–92.
Basseri, S., & Austin, R. C. (2012). Endoplasmic reticulum stress and lipid metabolism:
Mechanisms and therapeutic potential. Biochemistry Research International, 2012,
841362.
Brentnall, M., Rodriguez-Menocal, L., De Guevara, R. L., et al. (2013). Caspase-9,
caspase-3 and caspase-7 have distinct roles during intrinsic apoptosis. BMC Cell
Biology, 14(1), 1–9.
Chang, Z., Xing, J., & Yu, X. (2014). Curcumin induces osteosarcoma MG63 cells
apoptosis via ROS/Cyto-C/Caspase-3 pathway. Tumor Biology, 35(1), 753–758.
Chaudhari, N., Talwar, P., Parimisetty, A., et al. (2014). A molecular web: Endoplasmic
reticulum stress, inflammation, and oxidative stress. Frontiers in Cellular
Neuroscience, 8, 213.
Chen, Z., Tian, R., She, Z., et al. (2020). Role of oxidative stress in the pathogenesis of
nonalcoholic fatty liver disease. Free Radical Biology and Medicine, 152, 116–141.
Chenxu, G., Xianling, D., Qin, K., et al. (2021). Fisetin protects against high fat dietinduced nephropathy by inhibiting inflammation and oxidative stress via the
blockage of iRhom2/NF-κB signaling. International Immunopharmacology, 92,
107353.
Cullinan, S. B., Zhang, D., Hannink, M., et al. (2003). Nrf2 is a direct PERK substrate and
effector of PERK-dependent cell survival. Molecular and Cellular Biology, 23(20),
7198–7209.
De Vries, H. E., Witte, M., Hondius, D., et al. (2008). Nrf2-induced antioxidant
protection: A promising target to counteract ROS-mediated damage in
neurodegenerative disease? Free Radical Biology and Medicine, 45(10), 1375–1383.
Ethics statement
All research procedures associating with mice were approved by the
Institutional Animal Care and Use Committee in Chongqing Key Labo­
ratory of Medicinal Resources in the Three Gorges Reservoir Region,
School of Biological and Chemical Engineering, Chongqing University of
Education, and were performed in accordance with the Guide for the
Care and Use of Laboratory Animals, issued by the National Institutes of
Health in 1996. The protocols used in this study were in accordance with
the Regulations of Experimental Animal Administration issued by the
Ministry of Science and Technology of the People’s Republic of China
(http://www.most.gov.cn).
15
X. Dai et al.
Journal of Functional Foods 90 (2022) 104954
Pal, H. C., Sharma, S., Elmets, C. A., et al. (2013). Fisetin inhibits growth, induces G 2/M
arrest and apoptosis of human epidermoid carcinoma A 431 cells: Role of
mitochondrial membrane potential disruption and consequent caspases activation.
Experimental Dermatology, 22(7), 470–475.
Parafati, M., Kirby, R. J., Khorasanizadeh, S., et al. (2018). A nonalcoholic fatty liver
disease model in human induced pluripotent stem cell-derived hepatocytes, created
by endoplasmic reticulum stress-induced steatosis. Disease Models & Mechanisms, 11
(9), dmm033530.
Perelman, A., Wachtel, C., Cohen, M., et al. (2012). JC-1: Alternative excitation
wavelengths facilitate mitochondrial membrane potential cytometry. Cell Death &
Disease, 3(11), e430.
Sabarwal, A., Agarwal, R., & Singh, R. P. (2017). Fisetin inhibits cellular proliferation
and induces mitochondria-dependent apoptosis in human gastric cancer cells.
Molecular Carcinogenesis, 56(2), 499–514.
Sanda, W., Min Robert, W. M., Jun, Z., et al. (2021). Hepatic mitochondrial SAB deletion
or knockdown alleviates diet induced metabolic syndrome, steatohepatitis and
hepatic fibrosis. Hepatology. https://doi.org/10.1002/hep.32083
Schröder, M. (2008). Endoplasmic reticulum stress responses. Cellular and Molecular Life
Sciences, 65(6), 862–894.
Shanmugam, K., Prem, P. N., Boovarahan, S. R., et al. (2021). FIsetin Preserves
Interfibrillar Mitochondria to Protect Against Myocardial Ischemia-Reperfusion
Injury. Cell Biochemistry and Biophysics, 1–15.
Toriguchi, K., Hatano, E., Tanabe, K., et al. (2014). Attenuation of steatohepatitis,
fibrosis, and carcinogenesis in mice fed a methionine-choline deficient diet by
CCAAT/enhancer-binding protein homologous protein deficiency. Journal of
Gastroenterology and Hepatology, 29(5), 1109–1118.
Turillazzi, E., Neri, M., Cerretani, D., et al. (2016). Lipid peroxidation and apoptotic
response in rat brain areas induced by long-term administration of nandrolone: The
mutual crosstalk between ROS and NF-kB. Journal of Cellular and Molecular Medicine,
20(4), 601–612.
Wang, Q., Jiang, L., Wang, J., et al. (2009). Abrogation of hepatic ATP-citrate lyase
protects against fatty liver and ameliorates hyperglycemia in leptin receptordeficient mice. Hepatology, 49(4), 1166–1175.
Wei, Y., Rector, R. S., Thyfault, J. P., et al. (2008). Nonalcoholic fatty liver disease and
mitochondrial dysfunction. World Journal of Gastroenterology: WJG, 14(2), 193.
Xu, M., Ge, C., Qin, Y., et al. (2019a). Activated TNF-α/RIPK3 signaling is involved in
prolonged high fat diet-stimulated hepatic inflammation and lipid accumulation:
Inhibition by dietary fisetin intervention. Food & Function, 10(3), 1302–1316.
Xu, B., Qin, Y., Li, D., et al. (2020). Inhibition of PDE4 protects neurons against oxygenglucose deprivation-induced endoplasmic reticulum stress through activation of the
Nrf-2/HO-1 pathway. Redox Biology, 28, 101342.
Xu, M., Sun, Y., Dai, X., et al. (2019b). Fisetin attenuates high fat diet-triggered hepatic
lipid accumulation: A mechanism involving liver inflammation overload associated
TACE/TNF-α pathway. Journal of Functional Foods, 53, 7–21.
Zhang, Y., Jiang, L., Hu, W., et al. (2011). Mitochondrial dysfunction during in vitro
hepatocyte steatosis is reversed by omega-3 fatty acid–induced up-regulation of
mitofusin 2. Metabolism, 60(6), 767–775.
Zhang, K., & Kaufman, R. J. (2008). From endoplasmic-reticulum stress to the
inflammatory response. Nature, 454(7203), 455–462.
Zhang, B., Li, M., Zou, Y., et al. (2019). NFκB/Orai1 facilitates endoplasmic reticulum
stress by oxidative stress in the pathogenesis of non-alcoholic fatty liver disease.
Frontiers in Cell and Developmental Biology, 7, 202.
Zhang, L., Wang, H., Zhou, Y., et al. (2018). Fisetin alleviates oxidative stress after
traumatic brain injury via the Nrf2-ARE pathway. Neurochemistry International, 118,
304–313.
Zhang, X. Q., Xu, C. F., Yu, C. H., et al. (2014). Role of endoplasmic reticulum stress in
the pathogenesis of nonalcoholic fatty liver disease. World Journal of
Gastroenterology: WJG, 20(7), 1768.
Zhou, J., Zhou, F., Wang, W., et al. (2020). Epidemiological features of NAFLD from 1999
to 2018 in China. Hepatology, 71(5), 1851–1864.
Eduardo, V.-G., Nephew Lauren, D., Raj, V., et al. (2021). High quality diet, physical
activity and college education are associated with low risk of NAFLD among the U.S.
population. Hepatology. https://doi.org/10.1002/hep.32207
Fanta, B., Saad, K., Haiguang, W., et al. (2021). The Emerging Role of B Cells in the
Pathogenesis of NAFLD. Hepatology. https://doi.org/10.1002/hep.31889
Friedman, S. L., Neuschwander-Tetri, B. A., Rinella, M., et al. (2018). Mechanisms of
NAFLD development and therapeutic strategies. Nature Medicine, 24(7), 908–922.
Gabbia, D., Cannella, L., & De Martin, S. (2021). The Role of Oxidative Stress in NAFLDNASH-HCC Transition-Focus on NADPH Oxidases. Biomedicines, 9(6), 687.
Ge, C. X., Xu, M. X., Qin, Y. T., et al. (2019). Endoplasmic reticulum stress-induced
iRhom2 up-regulation promotes macrophage-regulated cardiac inflammation and
lipid deposition in high fat diet (HFD)-challenged mice: Intervention of fisetin and
metformin. Free Radical Biology and Medicine, 141, 67–83.
Ghemrawi, R., Battaglia-Hsu, S. F., & Arnold, C. (2018). Endoplasmic reticulum stress in
metabolic disorders. Cells, 7(6), 63.
Gonzalez-Rodriguez, A., Mayoral, R., Agra, N., et al. (2014). Impaired autophagic flux is
associated with increased endoplasmic reticulum stress during the development of
NAFLD. Cell Death & Disease, 5(4), e1179.
Grynkiewicz, G., & Demchuk, O. M. (2019). New perspectives for fisetin. Frontiers in
Chemistry, 7, 697.
Hu, L. F., Feng, J., Dai, X., et al. (2020). Oral flavonoid fisetin treatment protects against
prolonged high-fat-diet-induced cardiac dysfunction by regulation of multicombined
signaling. The Journal of Nutritional Biochemistry, 77, 108253.
Hu, H., Tian, M., Ding, C., et al. (2019). The C/EBP homologous protein (CHOP)
transcription factor functions in endoplasmic reticulum stress-induced apoptosis and
microbial infection. Frontiers in Immunology, 9, 3083.
Im, Y. R., Hunter, H., de Gracia Hahn, D., et al. (2021). A Systematic Review of Animal
Models of NAFLD Finds High-Fat, High-Fructose Diets Most Closely Resemble
Human NAFLD. Hepatology. https://doi.org/10.1002/hep.31897
Iurlaro, R., & Muñoz-Pinedo, C. (2016). Cell death induced by endoplasmic reticulum
stress. The FEBS Journal, 283(14), 2640–2652.
Jianhua, R., Jiannan, Q., Ming, N.i., et al. (2021). Macrophage Nrf2 deficiency promotes
innate immune activation by Timp3-mediated RhoA/ROCK pathway in the ischemic
liver. Hepatology. https://doi.org/10.1002/hep.32184
Khan, N., Syed, D. N., Ahmad, N., et al. (2013). Fisetin: A dietary antioxidant for health
promotion. Antioxidants & Redox Signaling, 19(2), 151–162.
Lebeaupin, C., Proics, E., De Bieville, C. H. D., et al. (2015). ER stress induces NLRP3
inflammasome activation and hepatocyte death. Cell Death & Disease, 6(9), e1879.
Lebeaupin, C., Vallée, D., Hazari, Y., et al. (2018). Endoplasmic reticulum stress
signalling and the pathogenesis of non-alcoholic fatty liver disease. Journal of
Hepatology, 69(4), 927–947.
Lee, A. S. (2005). The ER chaperone and signaling regulator GRP78/BiP as a monitor of
endoplasmic reticulum stress. Methods, 35(4), 373–381.
Lee, J., Park, J. S., & Roh, Y. S. (2019). Molecular insights into the role of mitochondria
in non-alcoholic fatty liver disease. Archives of Pharmacal Research, 42(11), 935–946.
Maharjan, S., Oku, M., Tsuda, M., et al. (2014). Mitochondrial impairment triggers
cytosolic oxidative stress and cell death following proteasome inhibition. Scientific
Reports, 4(1), 1–11.
Marciniak, S. J., & Ron, D. (2006). Endoplasmic reticulum stress signaling in disease.
Physiological Reviews, 86(4), 1133–1149.
McPherson, S., Hardy, T., Henderson, E., et al. (2015). Evidence of NAFLD progression
from steatosis to fibrosing-steatohepatitis using paired biopsies: Implications for
prognosis and clinical management. Journal of Hepatology, 62(5), 1148–1155.
Niazpour, F., Hoseini, H., Seyyedebrahimi, S. S., et al. (2021). The relationship between
Nrf2/Keap1 system and endoplasmic reticulum stress and inflammatory markers in
peripheral blood mononuclear cells of type 2 diabetic subjects. Clinical Diabetology,
10(5), 394–402.
Oyinloye, B. E., Adenowo, A. F., & Kappo, A. P. (2015). Reactive oxygen species,
apoptosis, antimicrobial peptides and human inflammatory diseases.
Pharmaceuticals, 8(2), 151–175.
16
Téléchargement