Next Article in Journal
The HIFα-Stabilizing Drug Roxadustat Increases the Number of Renal Epo-Producing Sca-1+ Cells
Next Article in Special Issue
Post-Translational Modifications of ATG4B in the Regulation of Autophagy
Previous Article in Journal
Free Fatty Acid Receptors (FFARs) in Adipose: Physiological Role and Therapeutic Outlook
Previous Article in Special Issue
Transcriptional Regulation of Hepatic Autophagy by Nuclear Receptors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Basal Autophagy Is Necessary for A Pharmacologic PPARα Transactivation

1
Department of Biochemistry and Cell Biology, Cell and Matrix Research Institute, School of Medicine, Kyungpook National University, Daegu 41944, Korea
2
BK21 FOUR KNU Biomedical Convergence Program, Department of Biomedical Science, The Graduate School, Kyungpook National University, Daegu 41944, Korea
*
Author to whom correspondence should be addressed.
Cells 2022, 11(4), 754; https://doi.org/10.3390/cells11040754
Submission received: 7 January 2022 / Revised: 5 February 2022 / Accepted: 17 February 2022 / Published: 21 February 2022

Abstract

:
Autophagy is a conserved cellular process of catabolism leading to nutrient recycling upon starvation and maintaining tissue and energy homeostasis. Tissue-specific loss of core-autophagy-related genes often triggers diverse diseases, including cancer, neurodegeneration, inflammatory disease, metabolic disorder, and muscle disease. The nutrient-sensing nuclear receptors peroxisome proliferator-activated receptor α (PPARα) plays a key role in fasting-associated metabolisms such as autophagy, fatty acid oxidation, and ketogenesis. Here we show that autophagy defects impede the transactivation of PPARα. Liver-specific ablation of the Atg7 gene in mice showed reduced expression levels of PPARα target genes in response to its synthetic agonist ligands. Since NRF2, an antioxidant transcription factor, is activated in autophagy-deficient mice due to p62/SQSTM1 accumulation and its subsequent interaction with KEAP1, an E3 ubiquitin ligase. We hypothesize that the nuclear accumulation of NRF2 by autophagy defects blunts the transactivation of PPARα. Consistent with this idea, we find that NRF2 activation is sufficient to inhibit the pharmacologic transactivation of PPARα, which is dependent on the Nrf2 gene. These results reveal an unrecognized requirement of basal autophagy for the transactivation of PPARα by preventing NRF2 from a nuclear translocation and suggest a clinical significance of basal autophagy to expect a pharmacologic efficacy of synthetic PPARα ligands.

1. Introduction

Autophagy is an evolutionarily conserved catabolic and adaptive process induced by diverse conditions, including the deprivation of nutrient or growth factors, pathogen infection, hypoxia, and even exercise [1,2,3,4,5,6,7]. This leads to nutrient recycling, prevents cellular damage, and maintains tissues homeostasis in response to energy shortage and constant external and internal harmful insults. Therefore, autophagy has inevitable cytoprotective roles and needs to be precisely and tightly controlled to properly respond to various stimuli. It has been shown that autophagy dysfunctions are intimately associated with a wide range of human pathogenesis, including neurodegeneration, inflammatory and metabolic diseases, and cancers [3,8,9]. Recent studies reveal that autophagy used to be considered a nonselective process is able to selectively eliminate unwanted, potentially cytotoxic molecules, such as damaged organelles or aggregated proteins, serving as a primary cytoprotective function [10,11]. This leads to intensive research to modulate autophagy activity due to its therapeutic potential in many human diseases [12].
Macroautophagy is the major type of autophagy and has been most intensively studied. However, it is important to note that there are two other types of autophagy: microautophagy and chaperone-mediated autophagy. The initial genetic screen in yeasts identified many core autophagy-related genes. Subsequent studies have found that they are highly conserved in mammals, including mice and humans. For example, Atg1 and Atg6 in yeast are homologs of mammalian ULK1 and Beclin1, respectively [13].
In particular, tissue-specific Atg7 ablation using a Cre-loxp system in mice has revealed that core autophagy-related genes encoding macroautophagy machinery proteins are essential for maintaining almost every tissue homeostasis [14,15,16,17,18,19,20]. It has also been shown that liver-specific Atg7 (Atg7LKO) knockout mice showed hepatomegaly phenotypes and marked accumulations of p62, an autophagy adaptor protein [17,21]. Increased protein levels of p62 are capable of interacting with KEAP1, an E3 ubiquitin ligase, which liberates NRF2 from the ubiquitin-mediated proteasomal degradation system. As a potent antioxidant transcription factor, the nuclear NRF2 binds to the antioxidant response element (ARE) of many cytoprotective genes such as Nqo1, Hmox1, Gsta1, and so forth and upregulates their expressions to remove reactive oxygen species [22,23].
Among 48 members of nuclear receptors in the human genome, PPARα is the most notable nuclear receptor activated during fasting and regulates the expression of genes involved in fasting-mediated metabolism, including fatty acid oxidation and ketogenesis [24,25,26,27]. Consistent with its essential roles during fasting, Lee et al. showed that PPARα is necessary for fasting-induced autophagy and that pharmacological activation of PPARα is sufficient for induction of autophagy even in fed state [28,29]. Particularly, PPARα plays a key role in the induction of lipophagy, a selective autophagy for lipid droplets. This may provide peroxisomes and mitochondria with free fatty acids for β-oxidation and ketogenesis [28,29].
In this study, we want to define whether autophagy activity affects a pharmacological activation of nuclear receptor PPARα. A previous study has already suggested that both Atg7 knockdown in AML12 cells and Atg7LKO mice showed reduced concentrations of ketone body in response to synthetic PPARα agonists [28]. This result allows us to further investigate whether basal autophagy is essential for the pharmacologic activation of PPARα. We found that autophagy inhibition in vivo and in cultured cells blunts pharmacological induction of many PPARα target genes. Furthermore, as it has been reported that there are crosstalks between NRF2 and PPARγ, the closest homolog of PPARα in various tissues [30,31,32,33,34], NRF2 activation is often observed in conditions of autophagic defects might interfere with the pharmacological activation of PPARα in the nucleus.

2. Materials and Methods

2.1. Chemicals and Reagents

Wild-type C57BL/6J mice were purchased from Japan SLC, Inc. (Hamamatsu, Shizuoka, Japan) (C57BL/6JJmsSlc); Atg7F/F (RBRC02759) [17] mice were purchased from RIKEN BioResource Research Center (Tsukuba, Ibaraki, Japan); Alb-Cre mice (strain #003574) [35] and Nrf2F/F mice (strain #025433) [29,36] were purchased from Jackson laboratory (Bar Harbor, ME, USA); Keap1F/F mice [37] were obtained from Masayuki Yamamoto laboratory. AML12 cells were purchased from ATCC (Manassas, VA, USA) (CRL-2254); GW7647 (Cat. #10008613), Wy-14,643 (Cat. #70703), and L-sulforaphane (SFN, Cat. #14797) from Cayman; dimethylfumarate (DMF, Cat. #242926), butylated hydroxyanisole (BHA, Cat. #B1253-500G), dexamethasone (Cat. #D4902), polyethylene glycol 400 (PEG 400, Cat. #P3265-1KG), 3-methyladenine (Cat. #M9281) and Tween 80 (Cat. #P1754-500ML) from Sigma-Aldrich (St. Louis, MO, USA); Bafilomycin A1 (BafA1, Cat. #BML-CM110) from Enzo Life Science (Farmingdale, NY, USA); HyClone DMEM/F12 (Cat. #SH30023.01) and fetal bovine serum (FBS, Cat. #SH20084.03) from HyClone (Logan, UT, USA); Trizol Reagent (Cat. #15596018) from Invitrogen (Waltham, MA, USA); RbTaqTM qPCR 2X PreMIX (SYBR Green with high ROX, Cat. #RT531M) from Enzynomics (Daejeon, Korea); PrimeScriptTM 1st strand cDNA Synthesis kit (Cat. #6110A) from TaKaRa Bio Inc. (Kusatsu, Shiga, Japan); cOmpleteTM, Mini, EDTA-free Protease Inhibitor Cocktail (Cat. #11836170001), PhosSTOPTM (Cat. #04906837001), and Insulin-Transferrin-Sodium Selenite Supplement (ITS, Cat. #11074547001) from Roche (Basel, Switzerland); penicillin-streptomycin (Cat. #15140122), PierceTM BCA protein assay kit (Cat. #23225) and Pierce ECL Western Blotting Substrate (Cat. #32106) from Thermo Scientific (Waltham, MA, USA); BSA (Cat. #0332-100G) from VWR (Radnor, PA, USA); Immun-Blot PVDF Membrane (Cat. #1620177) from Bio-Rad Laboratories (Hercules, CA, USA); Rabbit monoclonal anti-β-actin (13E5) antibody (Cat. #5125, 1:3000 dilution), Rabbit monoclonal anti-Atg7 (D12B11) antibody (Cat. #8558S, 1:1000 dilution), Rabbit SQSTM1/p62 antibody (Cat. #5114S, 1:1000 dilution) and anti-rabbit IgG HRP-linked antibody (Cat. #7074S, 1:3000 dilution) from Cell Signaling Technology (Danvers, MA, USA); Dimethyl sulfoxide (DMSO, Cat. #sc-358801) from Santa Cruz Biotechnology, Inc. (Dallas, TX, USA); Atg5−/− (RCB 2711) and Atg7−/− (RCB 3707) MEFs from Riken BRC. Information for other reagents not shown here is described in the relevant methods and references.

2.2. Animal Studies

All animal studies and procedures were approved by the Institutional Animal Care and Use Committee of the Kyungpook National University (KNU-2020-036). Wild-type mice were C57BL/6J mice. Male Alb-Cre/+ mice were first bred with either female Atg7F/F, Keap1F/F, or Nrf2F/F mice to generate male Alb-Cre; Atg7F/+, Alb-Cre; Keap1F/+, or Alb-Cre; Nrf2F/+ mice. These mice were then further bred with female Atg7F/F, Keap1F/F, or Nrf2F/F mice to generate Alb-Cre; Atg7F/F (Atg7LKO), Alb-Cre; Keap1F/F (Keap1LKO), or Alb-Cre; Nrf2F/F mice (Nrf2LKO) that showed a hepatocyte-specific ablation of either Atg7, Keap1, or Nrf2 gene, respectively. All experiments were performed in ad libitum fed male mice unless otherwise indicated. To activate PPARα in the liver, 8–9 week-old male wild-type C57BL/6J, Atg7F/F, Atg7LKO, Alb-Cre/+, Keap1LKO, Nrf2F/F, or Nrf2LKO mice were intraperitoneally injected with vehicle (0.1% dimethylsulfoxide (DMSO) in 90:5:5 of saline, PEG-400, and Tween 80, respectively) or GW7647 (10 mg/kg BW) twice a day (first injection at 00:00 and second injection at 12:00). A total of 5 h after the second injection, mice were sacrificed to collect livers (Figure 1, Figure S1, Figures 4 and 5). To activate NRF2, 8–9 week-old male wild-type C57BL/6J, Nrf2F/F, or Nrf2LKO mice were orally gavaged with either vehicle or butylated hydroxyanisole (BHA, 200 mg/kg BW) once a day at 12:00 for 3 days (Figures 4 and 5). To avoid circadian issues, all mice were sacrificed at 17:00–18:00.

2.3. Cell Culture and Drug Treatments

AML12 cells were maintained in the complete media: DMEM/F12 high glucose supplemented with 10% FBS, 1% ITS, 1% penicillin-streptomycin antibiotics, and 40 ng/mL dexamethasone. About 70–80% of confluent cells were seeded in a 6- or 12-well plate in a 1:5 ratio 48 h before drug treatments. To assess a pharmacological activity of PPARα in an autophagy-inhibited condition (Figure 1c–e and Figure S2c–e), these cells were treated with either bafilomycin A1 (BafA1) or 3-methyladenine (3-MA) in a dose-dependent manner (BafA1 and 3-MA, 0.1, 0.5, or 1 μM). Simultaneously, they were also treated with synthetic PPARα agonists 1 μM GW7647 or 100 μM Wy-14,643. To determine the pharmacological activity of PPARα in NRF2-activated conditions, AML12 cells were treated with either sulforaphane (SFN, 12.5, 25, or 50 μM) or dimethylfumarate (DMF, 50, 75, or 100 μM) in a dose-dependent manner. Likewise, these cells were also dose-dependently treated with GW7647 (0.01, 0.1, or 1 μM) or Wy-14,643 (1, 10, or 100 μM). On the other hand, AML12 cells were treated with either sulforaphane (25 μM SFN, 6, 12, or 18 h) or dimethylfumarate (75 or 100 μM DMF, 6, 12, or 18 h) in a time-dependent manner. These cells were simultaneously treated with 1 μM GW7647 or 100 μM Wy-14,643 for 24 h. The vehicle is 0.1% DMSO. Mouse embryonic fibroblasts derived from wild-type, Atg5−/−, or Atg7−/− mouse embryos were maintained in DMEM high glucose supplemented with 10% FBS, 1% ITS, and 1% penicillin-streptomycin antibiotics.

2.4. RNA Purification, cDNA Synthesis, and qPCR Analysis

Total RNA was purified from the snap-frozen mouse liver or AML12 cells using Trizol Reagent according to the manufacturer’s instructions (Invitrogen: Waltham, MA, USA). RNA concentration and quality were determined by Nanodrop. A total of 1 μg of total RNA was converted to cDNA using PrimeScriptTM 1st strand cDNA Synthesis kit (Takara Bio Inc.: Kusatsu, Shiga, Japan). qPCR was performed with RbTaqTM qPCR 2X PreMIX (SYBR Green with high ROX) (Enzynomics: Daejeon, Korea) on StepOnePlus Real-Time PCR systems (Applied Biosystems: Waltham, MA, USA). All reactions were carried out in either duplicate or triplicate, and Ct values were obtained. mRNA levels were normalized to the expression levels of the housekeeping gene 36b4 (also known as Rplp0) using the standard curve method. Information of mouse primer sequences for qPCR analysis is listed in Table S1.

2.5. Immunoblot Analysis

~100 mg of frozen liver tissue was homogenized in 1 mL of ice-cold RIPA buffer (25 mM Tris-HCl, pH 7.6, 150 mM NaCl, 1% NP-40, 1% sodium deoxycholate, 1 mM EDTA) supplemented with cOmplete protease and PhosSTOP inhibitors (Roche: Basel, Switzerland), followed by brief sonication in ice. The lysates were centrifuged at 13,000 rpm for 15 min at 4 °C to collect supernatants, of which protein concentration was determined by the Pierce BCA protein assay kit (ThermoFisher Scientific: Waltham, MA, USA). A total of 25 μg of protein lysates were loaded on a 10% SDS-PAGE gel to be resolved, and the proteins were transferred onto an Immuno-Blot PVDF membrane (Bio-Rad Laboratories: Hercules, CA, USA). The membrane was blocked with 5% BSA/TBST (Tris-buffered saline with 0.1% Tween-20) solution for 1 h at room temperature. Primary antibodies were diluted in 5% BSA/TBST solution, which was applied onto the membranes at 4 °C for 16 h. The membrane was washed three times at room temperature (each wash for 15 min; the first and second washes with 5% BSA/TBST solution; the third wash with TBST solution). After washes, the membrane was incubated with an anti-rabbit IgG HRP-linked antibody (Cell Signaling Technology (Danvers, MA, USA)) at room temperature for 1 h, followed by three washes as shown above. The bands were visualized using the Pierce ECL Western Blotting Substrate (ThermoFisher Scientific: Waltham, MA, USA).

2.6. Statistical Analysis

All values are shown as mean ± s.e.m. and error bars were derived from biological replicates rather than technical replicates. Significant differences between the two groups were evaluated using a two-tailed, unpaired Student t-test, which was found to be appropriate as groups displayed a normal distribution, and comparable variance of p < 0.05 was considered statistically significant.

3. Results

3.1. Autophagy Inhibition Impairs a Pharmacologic PPARα Transactivation

To examine any crosstalk between autophagy and nuclear receptor signaling, liver-specific Atg7 knockout (Atg7LKO) mice were generated by crossing male Alb-Cre mice with female Atg7F/F mice. Consistent with previous reports, hepatomegaly phenotypes were observed in Atg7LKO mice compared with Atg7F/F control littermates (Figure S1a). As expected, hepatic expression and protein levels of the Atg7 gene were also markedly reduced in Atg7LKO mice (Figure S1b,c).
A previous study has shown that increased serum levels of β-hydroxybutyrate, one species of ketone bodies in Atg7F/F control littermates upon treatment of a synthetic PPARα agonist GW7647 [38], were blunted in Atg7LKO mice. These results allowed us to set up a hypothesis that core autophagy-related genes are essential for the normal function of PPARα. To test this, hepatic expression levels of PPARα target genes (Acox1, Acot3, Fgf21, Pdk4, Acot1, and Cidec) were determined in fed or fasted Atg7F/F control littermates and Atg7LKO mice in response to intraperitoneal treatment of GW7647 (Figure 1a). Consistent with a previous report, pharmacologic inductions of hepatic PPARα target genes were markedly increased in Atg7F/F mice compared with those of vehicle-treated Atg7F/F mice. However, these inductions were largely decreased in GW7647-treated Atg7LKO mice (Figure 1b). To investigate whether these in vivo results recapitulate in cultured cells, we used mouse embryonic fibroblasts (MEFs) derived from wild-type, Atg5−/−, or Atg7−/− mice (Figure S2a) [17,39]. As expected, expression levels of Atg5 or Atg7 were almost completely absent in either Atg5−/− or Atg7−/− MEFs. Treatment of Wy-14,643, a synthetic PPARα agonist in wild-type MEFs robustly increased expressions of PPARα target genes Acox1, Pdk4, Acot3, and Ucp2, but these responses were markedly blunted in either Atg5−/− or Atg7−/− MEFs (Figure S2b,c). These results suggest that core-autophagy genes are required for the pharmacological PPARα activation.
To examine whether pharmacologic inhibition of autophagy also shows similar results, mouse hepatocyte cell line AML12 cells were dose-dependently treated with 3-methyladenine or bafilomycin A1, well-known autophagy inhibitors with or without synthetic PPARα agonist ligands GW7647 or Wy-14,643 (Figure S2d and Figure 1c) [40,41]. Consistent with in vivo mouse results shown in Figure 1b, inductions of most of PPARα target genes upon GW7647 or Wy-14,643 treatment were dramatically blunted by a dose-dependent treatment of 3-methyladenine or bafilomycin A1 (Figure S2f and Figure 1d,e). All these in vivo and cell culture results indicate that autophagy inhibition can impair a pharmacologic PPARα activity.
Previous reports have also demonstrated that genetic ablations of core autophagy-related genes, including Atg7, lead to the accumulation of p62, an autophagy adaptor protein [42]. We could also confirm marked accumulations of p62 in the livers of Atg7LKO mice (Figure S1d). This causes interactions with KEAP1, an E3 ubiquitin ligase, which liberates NRF2 from the ubiquitin-mediated proteasomal degradation system. Free cytoplasmic NRF2 is then translocated into the nucleus, where it acts as an antioxidant transcription factor to remove reactive oxygen species [21]. As expected, we also observed dramatically increased expression levels of the Nqo1 gene, an NRF2 target gene in the livers of Atg7LKO mice compared with those of Atg7F/F mice (Figure S1e). Atg5−/− and Atg7−/− MEFs also increased Nqo1 and Hmox1 expressions (Figure S2a). Based on these data, we concluded that our Atg7LKO mice also increased hepatic NRF2 activation. It is of interest to note that the pharmacologic PPARα activation by GW7647 treatment does not significantly alter hepatic Nqo1 expression compared with vehicle-treated counterparts (Figure S1e). We also found that a dose-dependent treatment of bafilomycin A1 in AML12 cells markedly increased expression levels of the Hmox1 gene, another known NRF2 target gene. However, expression levels of the Nqo1 gene showed opposite results, which might reflect different expression kinetics of two known NRF2 target genes in response to autophagy inhibition (Figure S1f,g). Treatment of 3-methyladenine in AML12 cells showed similar results (Figure S2d,e).

3.2. SFN-Mediated NRF2 Activation in a Dose-Dependent Manner Impairs a Pharmacologic PPARα Transactivation

To examine whether NRF2 activation directly affects PPARα activity, AML12 cells were dose-dependently treated with sulforaphane [43] or dimethylfumarate [44], potent NRF2 activators in the presence or absence of synthetic PPARα agonist ligands GW7647 or Wy-14,643 (Figure 2a and Figure S2a). As expected, sulforaphane or dimethylfumarate dramatically increases the expression of NRF2 target genes Nqo1 and Hmox1, which are not altered by the treatment of synthetic PPARα agonists (Figure 2b,c and Figure S2b,c). These data indicate that a pharmacologic PPARα activation may not affect sulforaphane or dimethylfumarate-mediated NRF2 activation.
Intriguingly, however, completely opposite results were observed in terms of a pharmacologic PPARα activation in the presence of NRF2 activators. We found that dose-dependent treatments of sulforaphane or dimethylfumarate markedly downregulated expression of PPARα target genes Pdk4, Acot3, and Ucp2 in response to GW7647 or Wy-14,643 treatment (Figure 2d,e and Figure S2d,e). All these data suggest that NRF2 activation might mediate the suppression of a pharmacologic PPARα activation.

3.3. SFN-Mediated NRF2 Activation in a Time-Dependent Manner Impairs a Pharmacologic PPARα Transactivation

To further thoroughly investigate whether NRF2 activation directly affects PPARα activity, AML12 cells were time-dependently treated with sulforaphane or dimethylfumarate in the presence or absence of GW7647 or Wy-14,643 (Figure 3a and Figure S3a). As expected, these NRF2 activators dramatically increase expression of the Nqo1 gene in a time-dependent manner (Figure 3b and Figure S3b). However, as mentioned above, we again observed different expression kinetics of two known NRF2 target genes, Nqo1 and Hmox1. A total of 6 h treatment of NRF2 activators showed the highest expression levels of Hmox1 gene whose mRNA levels were gradually decreased by further increasing treatment time of NRF2 activators (Figure 3b and Figure S3b). Nevertheless, time-dependent treatments of sulforaphane or dimethylfumarate in AML12 cells dramatically suppress the expression of PPARα target genes Pdk4, Acot3, and Ucp2 in response to GW7647 or Wy-14,643 (Figure 3c,d and Figure S3c,d). All these data support that acute NRF2 activation might mediate the suppression of the pharmacologic PPARα activation.

3.4. NRF2 Activation Is Necessary and Sufficient for Suppressing a Pharmacologic PPARα Transactivation

To further confirm these results in vivo, C57BL/6J wild-type mice were orally gavaged with butylated hydroxyanisole (BHA), an NRF2 activator, for 3 days [45]. On day 3, these mice were intraperitoneally treated with vehicle or GW7647, as shown in Figure 4a. After that, hepatic expression levels of PPARα target genes Pdk4, Ucp2, Acot3, Acox1, and Fgf21 were determined. As expected, GW7647 treatment markedly increased mRNA levels of these PPARα target genes, but these responses were significantly blunted in the livers of BHA-treated mice (Figure 4b). These data indicate that BHA-mediated NRF2 activation is able to suppress a pharmacologic PPARα activation.
To genetically validate these results, we generated liver-specific Keap1 knockout (Keap1LKO) mice by crossing male Alb-Cre mice with female Keap1F/F mice. Since floxed Keap1 alleles in Keap1F/F mice turned out to be hypomorphic, leading to the spontaneous activation of NRF2, we used Alb-Cre control littermates as control mice instead of Keap1F/F mice (Figure 4c) [46]. As expected, Keap1LKO mice showed almost an absence of hepatic Keap1 expressions but marked increased mRNA levels of Nqo1 and Gasta1, indicating that a potent NRF2 activation occurs in the livers of Keap1LKO mice. Once again, hepatic expression levels of PPARα target genes Pdk4, Ucp2, Cidec, Acot1 to 3, and Acox1 were significantly downregulated in GW7647-treated Keap1LKO mice (Figure 4d). These data suggest that genetic NRF2 activation leads to the inhibition of the pharmacological PPARα activation.
To genetically validate whether these blunted pharmacological PPARα activations mediated by NRF2 activators depend on NRF2 proteins, we first generated liver-specific Nrf2 knockout (Nrf2LKO) mice by crossing male Alb-Cre mice with female Nrf2F/F mice. Nrf2F/F control littermates and Nrf2LKO mice were then orally gavaged with BHA for 3 days. On day 3, these mice were intraperitoneally treated with vehicle or GW7647, as shown in Figure 5a. As expected, hepatic Nrf2 mRNA levels were markedly reduced in Nrf2LKO mice, confirming a genetic ablation of the Nrf2 gene in the liver (Figure 5b). Like the results obtained in C57BL/6J mice (Figure 4), BHA treatment also robustly increased hepatic expression levels of NRF2 target genes Nqo1 and Gasta1 in Nrf2F/F control littermates, but these responses were significantly blunted in Nrf2LKO mice (Figure 5b).
These results demonstrate that induction of Nqo1 and Gsta1 genes by BHA treatment largely depends on NRF2. Moreover, GW7647 treatment in Nrf2F/F control littermates also robustly induced hepatic expressions of PPAR target genes Pdk4, Ucp2, Cidec, and Acot1 to 3, but these inductions were significantly blunted in BHA-treated Nrf2F/F counterparts. However, these suppressed expression levels of PPAR target genes were almost completely lost in BHA-treated Nrf2LKO mice, demonstrating that blunted pharmacological transactivation of PPARα mediated by BHA treatment indeed requires NRF2 proteins (Figure 5c).

4. Discussion

In this study, we showed that inhibition of basal autophagy impairs a pharmacological activation of the fasting-activated nuclear receptor PPARα. Moreover, our data also suggested that NRF2 known to be activated in autophagy-defective conditions might mediate this effect. In the clinical aspects, this phenomenon could be critical, in particular for hyperlipidemia patients taking fibrate derivatives. Simultaneous activation of both NRF2 and PPARα transcription factors might diminish the beneficial effects of fibrate medications.
Similar results have also been reported by the Komatsu and Panasyuk laboratories in regards to the impairment of PPARα transactivation by autophagy defects, although their molecular mechanisms are distinct from ours [47,48]. They have found that NCoR1 corepressor proteins are remarkedly accumulated in autophagy-defective hepatocytes obtained from Atg5, Atg7, or Vps15 conditional knockout mice. These increased levels of NCoR1 tend to be recruited to PPARα, which leads to suppression of its target gene expression. In terms of the detailed molecular mechanism by which autophagy defects impede PPARα signaling, we speculate that there are many more mechanisms yet to be discovered.
Similar to our findings, sulforaphane treatment in 3T3-L1 cells inhibits adipocyte differentiation by reducing the expression of PPARγ, a master regulator of adipogenesis, indicating that NRF2 activation might impair functionalities of other PPARs [30]. Conversely, NRF2 activation also increases the expression of PPARs. Treatment of either BHA or sulforaphane has been shown to induce hepatic PPARγ or PPARβ/δ expression in livers of wild-type mice [31,32]. Consistently, in our experiments in Figure 4a, we could confirm that a 3-day treatment of BHA increases expression of both Pparγ and Pparβ/δ genes (data not shown). Moreover, cooperative actions between PPARγ and NRF2 have also been reported. In oxidative stress conditions, both transcription factors reciprocally induce their expression as an upstream regulator, resulting in increased expressions of many cytoprotective genes encoding HO-1, CAT, SOD, and GST. These enzymes have been known to alleviate reactive oxygen species in cells and tissues [33,34].
It has been known that heme oxygenase-1 (HO-1) encoded by the Hmox1 gene, a known NRF2 target gene, produces biliverdin from heme, and biliverdin reductase then converts biliverdin to bilirubin [49]. Interestingly, the Hinds laboratory has shown that bilirubin might act as an endogenous agonist ligand for PPARα, although its physiological relevance to the fasting-activated PPARα signaling should be confirmed [50,51]. According to this scenario, PPARα target genes might be induced rather than reduced via an HO-1/bilirubin/PPARα axis. However, we found that treatment of sulforaphane, dimethylfumarate, or BHA, or liver-specific Keap1 deletion did not increase basal expression levels of most PPARα target genes at least we examined, indicating that NRF2 activation might not affect a basal activity of PPARα.
In the future, the generation of liver-specific Atg7 and Nrf2 double knockout (Atg7LKO; Nrf2LKO) mice may provide a valuable resource to address whether NRF2 activation in Atg7LKO mice is genetically required for suppressing the pharmacological activation of the nuclear receptor PPARα. Moreover, the generation of liver-specific Nrf2 and Keap1 double knockout (Nrf2LKO; Keap1LKO) mice may also give us important insights to examine whether the blunted pharmacological activation of PPARα in Keap1LKO mice is genetically dependent on NRF2.
Further investigations are also necessary to define detailed molecular mechanisms by which nuclear NRF2 proteins interfere with the functions of the nuclear receptor PPARα. This could be achieved in a variety of mechanisms, including a direct protein-protein interaction, squelching coactivators, interfering with the recruitment of coactivators, promoting recruitment of corepressors, inhibiting nuclear receptor binding to its response element of target genes, and so on.

5. Conclusions

In summary, we propose an essential role of basal autophagy activity for the pharmacological functions of the nuclear receptor PPARα. Autophagic defects lead to NRF2 translocation from the cytoplasm to the nucleus, where it acts as an antioxidant transcription factor by regulating many cytoprotective genes. We speculate that these nuclear NRF2 transcription factor proteins could interfere with the activities of many other transcription factors, including PPARα, via a variety of mechanisms (Figure 6). This study establishes the unexpected role of basal autophagy for the therapeutic effects of pharmacologically targeting the nuclear receptor PPARα, which may be useful for the treatment of diverse metabolic diseases.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cells11040754/s1, Figure S1: Upregulation of NRF2 target genes in response to autophagy inhibition; Figure S2: Both genetic and pharmacological inhibitions of autophagy suppress the pharmacological PPARα activation. Figure S3: DMF-mediated NRF2 activation in a dose-dependent manner impairs a pharmacologic PPARα transactivation; Figure S4: DMF-mediated NRF2 activation in a time-dependent manner impairs a pharmacologic PPARα transactivation; Table S1: Mouse primer sequences for qPCR analysis.

Author Contributions

Conceptualization, J.M.L.; methodology, E.Y.K. and J.M.L.; validation, E.Y.K. and J.M.L.; formal analysis, E.Y.K. and J.M.L.; investigation, E.Y.K. and J.M.L.; writing—original draft preparation, E.Y.K.; writing—review and editing, J.M.L.; visualization, E.Y.K.; supervision, J.M.L.; project administration, E.Y.K. and J.M.L.; funding acquisition, J.M.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIT): 2016R1C1B2009587, 2019R1H1A1101394.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

The authors would like to thank Noboru Mizushima for Atg5−/− MEFs, Masaaki Komatsu for Atg7−/− MEFs and Atg7F/F mice, Masayuki Yamamoto for Keap1F/F mice, and Youngtae Jeong for a rederivation of Keap1F/F mice. Furthermore, we would like to thank David D. Moore, Kang Ho Kim, Martin Wagner, and all members of the Nuclear Receptor and Disease Laboratory for critical comments, helpful suggestions, and additional support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kim, S.; Eun, H.S.; Jo, E.K. Roles of Autophagy-Related Genes in the Pathogenesis of Inflammatory Bowel Disease. Cells 2019, 8, 77. [Google Scholar] [CrossRef]
  2. Klionsky, D.J.; Abdelmohsen, K.; Abe, A.; Abedin, M.J.; Abeliovich, H.; Acevedo Arozena, A.; Adachi, H.; Adams, C.M.; Adams, P.D.; Adeli, K.; et al. Guidelines for the use and interpretation of assays for monitoring autophagy (3rd edition). Autophagy 2016, 12, 1–222. [Google Scholar] [CrossRef]
  3. Levine, B.; Kroemer, G. Autophagy in the pathogenesis of disease. Cell 2008, 132, 27–42. [Google Scholar] [CrossRef]
  4. Mizushima, N.; Levine, B.; Cuervo, A.M.; Klionsky, D.J. Autophagy fights disease through cellular self-digestion. Nature 2008, 451, 1069–1075. [Google Scholar] [CrossRef]
  5. Rabinowitz, J.D.; White, E. Autophagy and metabolism. Science 2010, 330, 1344–1348. [Google Scholar] [CrossRef]
  6. Silwal, P.; Paik, S.; Jeon, S.M.; Jo, E.K. Nuclear Receptors as Autophagy-Based Antimicrobial Therapeutics. Cells 2020, 9, 1979. [Google Scholar] [CrossRef]
  7. Yazdani, H.O.; Huang, H.; Tsung, A. Autophagy: Dual Response in the Development of Hepatocellular Carcinoma. Cells 2019, 8, 91. [Google Scholar] [CrossRef]
  8. Klionsky, D.J.; Petroni, G.; Amaravadi, R.K.; Baehrecke, E.H.; Ballabio, A.; Boya, P.; Bravo-San Pedro, J.M.; Cadwell, K.; Cecconi, F.; Choi, A.M.K.; et al. Autophagy in major human diseases. EMBO J. 2021, 40, e108863. [Google Scholar] [CrossRef] [PubMed]
  9. Takamura, A.; Komatsu, M.; Hara, T.; Sakamoto, A.; Kishi, C.; Waguri, S.; Eishi, Y.; Hino, O.; Tanaka, K.; Mizushima, N. Autophagy-deficient mice develop multiple liver tumors. Genes Dev. 2011, 25, 795–800. [Google Scholar] [CrossRef] [PubMed]
  10. Gatica, D.; Lahiri, V.; Klionsky, D.J. Cargo recognition and degradation by selective autophagy. Nat. Cell Biol. 2018, 20, 233–242. [Google Scholar] [CrossRef] [PubMed]
  11. Green, D.R.; Levine, B. To be or not to be? How selective autophagy and cell death govern cell fate. Cell 2014, 157, 65–75. [Google Scholar] [CrossRef] [PubMed]
  12. Fleming, A.; Noda, T.; Yoshimori, T.; Rubinsztein, D.C. Chemical modulators of autophagy as biological probes and potential therapeutics. Nat. Chem. Biol. 2011, 7, 9–17. [Google Scholar] [CrossRef] [PubMed]
  13. Nakatogawa, H.; Suzuki, K.; Kamada, Y.; Ohsumi, Y. Dynamics and diversity in autophagy mechanisms: Lessons from yeast. Nat. Rev. Mol. Cell Biol. 2009, 10, 458–467. [Google Scholar] [CrossRef]
  14. Ebato, C.; Uchida, T.; Arakawa, M.; Komatsu, M.; Ueno, T.; Komiya, K.; Azuma, K.; Hirose, T.; Tanaka, K.; Kominami, E.; et al. Autophagy is important in islet homeostasis and compensatory increase of beta cell mass in response to high-fat diet. Cell Metab. 2008, 8, 325–332. [Google Scholar] [CrossRef] [PubMed]
  15. Jung, H.S.; Chung, K.W.; Won Kim, J.; Kim, J.; Komatsu, M.; Tanaka, K.; Nguyen, Y.H.; Kang, T.M.; Yoon, K.H.; Kim, J.W.; et al. Loss of autophagy diminishes pancreatic beta cell mass and function with resultant hyperglycemia. Cell Metab. 2008, 8, 318–324. [Google Scholar] [CrossRef] [PubMed]
  16. Komatsu, M.; Waguri, S.; Chiba, T.; Murata, S.; Iwata, J.; Tanida, I.; Ueno, T.; Koike, M.; Uchiyama, Y.; Kominami, E.; et al. Loss of autophagy in the central nervous system causes neurodegeneration in mice. Nature 2006, 441, 880–884. [Google Scholar] [CrossRef] [PubMed]
  17. Komatsu, M.; Waguri, S.; Ueno, T.; Iwata, J.; Murata, S.; Tanida, I.; Ezaki, J.; Mizushima, N.; Ohsumi, Y.; Uchiyama, Y.; et al. Impairment of starvation-induced and constitutive autophagy in Atg7-deficient mice. J. Cell Biol. 2005, 169, 425–434. [Google Scholar] [CrossRef]
  18. Masiero, E.; Agatea, L.; Mammucari, C.; Blaauw, B.; Loro, E.; Komatsu, M.; Metzger, D.; Reggiani, C.; Schiaffino, S.; Sandri, M. Autophagy is required to maintain muscle mass. Cell Metab. 2009, 10, 507–515. [Google Scholar] [CrossRef]
  19. Zhang, Y.; Goldman, S.; Baerga, R.; Zhao, Y.; Komatsu, M.; Jin, S. Adipose-specific deletion of autophagy-related gene 7 (atg7) in mice reveals a role in adipogenesis. Proc. Natl. Acad. Sci. USA 2009, 106, 19860–19865. [Google Scholar] [CrossRef]
  20. Singh, R.; Xiang, Y.; Wang, Y.; Baikati, K.; Cuervo, A.M.; Luu, Y.K.; Tang, Y.; Pessin, J.E.; Schwartz, G.J.; Czaja, M.J. Autophagy regulates adipose mass and differentiation in mice. J. Clin. Investig. 2009, 119, 3329–3339. [Google Scholar] [CrossRef]
  21. Komatsu, M.; Kurokawa, H.; Waguri, S.; Taguchi, K.; Kobayashi, A.; Ichimura, Y.; Sou, Y.S.; Ueno, I.; Sakamoto, A.; Tong, K.I.; et al. The selective autophagy substrate p62 activates the stress responsive transcription factor Nrf2 through inactivation of Keap1. Nat. Cell Biol. 2010, 12, 213–223. [Google Scholar] [CrossRef]
  22. Yamamoto, M.; Kensler, T.W.; Motohashi, H. The KEAP1-NRF2 System: A Thiol-Based Sensor-Effector Apparatus for Maintaining Redox Homeostasis. Physiol. Rev. 2018, 98, 1169–1203. [Google Scholar] [CrossRef]
  23. Baird, L.; Yamamoto, M. The Molecular Mechanisms Regulating the KEAP1-NRF2 Pathway. Mol. Cell. Biol. 2020, 40, e00099-20. [Google Scholar] [CrossRef]
  24. Evans, R.M.; Barish, G.D.; Wang, Y.X. PPARs and the complex journey to obesity. Nat. Med. 2004, 10, 355–361. [Google Scholar] [CrossRef] [PubMed]
  25. Issemann, I.; Green, S. Activation of a member of the steroid hormone receptor superfamily by peroxisome proliferators. Nature 1990, 347, 645–650. [Google Scholar] [CrossRef] [PubMed]
  26. Kersten, S.; Desvergne, B.; Wahli, W. Roles of PPARs in health and disease. Nature 2000, 405, 421–424. [Google Scholar] [CrossRef] [PubMed]
  27. Bensinger, S.J.; Tontonoz, P. Integration of metabolism and inflammation by lipid-activated nuclear receptors. Nature 2008, 454, 470–477. [Google Scholar] [CrossRef] [PubMed]
  28. Lee, J.M.; Wagner, M.; Xiao, R.; Kim, K.H.; Feng, D.; Lazar, M.A.; Moore, D.D. Nutrient-Sensing nuclear receptors coordinate autophagy. Nature 2014, 516, 112–115. [Google Scholar] [CrossRef]
  29. Lee, J.M. Transcriptional coordination of hepatic autophagy by nutrient-sensing nuclear receptor PPARα and FXR. Ann. Pediatr. Endocrinol. Metab. 2016, 21, 193–198. [Google Scholar] [CrossRef] [PubMed]
  30. Choi, K.M.; Lee, Y.S.; Sin, D.M.; Lee, S.; Lee, M.K.; Lee, Y.M.; Hong, J.T.; Yun, Y.P.; Yoo, H.S. Sulforaphane inhibits mitotic clonal expansion during adipogenesis through cell cycle arrest. Obesity 2012, 20, 1365–1371. [Google Scholar] [CrossRef] [PubMed]
  31. Huang, J.; Tabbi-Anneni, I.; Gunda, V.; Wang, L. Transcription factor Nrf2 regulates SHP and lipogenic gene expression in hepatic lipid metabolism. Am. J. Physiol. Gastrointest. Liver Physiol. 2010, 299, G1211–G1221. [Google Scholar] [CrossRef] [PubMed]
  32. Hu, R.; Xu, C.; Shen, G.; Jain, M.R.; Khor, T.O.; Gopalkrishnan, A.; Lin, W.; Reddy, B.; Chan, J.Y.; Kong, A.N. Gene expression profiles induced by cancer chemopreventive isothiocyanate sulforaphane in the liver of C57BL/6J mice and C57BL/6J/Nrf2 (−/−) mice. Cancer Lett. 2006, 243, 170–192. [Google Scholar] [CrossRef] [PubMed]
  33. Lee, C. Collaborative Power of Nrf2 and PPARgamma Activators against Metabolic and Drug-Induced Oxidative Injury. Oxid. Med. Cell Longev. 2017, 2017, 1378175. [Google Scholar] [CrossRef]
  34. Dovinova, I.; Kvandova, M.; Balis, P.; Gresova, L.; Majzunova, M.; Horakova, L.; Chan, J.Y.; Barancik, M. The role of Nrf2 and PPARgamma in the improvement of oxidative stress in hypertension and cardiovascular diseases. Physiol. Res. 2020, 69, S541–S553. [Google Scholar] [CrossRef]
  35. Postic, C.; Shiota, M.; Niswender, K.D.; Jetton, T.L.; Chen, Y.; Moates, J.M.; Shelton, K.D.; Lindner, J.; Cherrington, A.D.; Magnuson, M.A. Dual roles for glucokinase in glucose homeostasis as determined by liver and pancreatic beta cell-specific gene knock-outs using Cre recombinase. J. Biol. Chem. 1999, 274, 305–315. [Google Scholar] [CrossRef]
  36. Kong, X.; Thimmulappa, R.; Craciun, F.; Harvey, C.; Singh, A.; Kombairaju, P.; Reddy, S.P.; Remick, D.; Biswal, S. Enhancing Nrf2 pathway by disruption of Keap1 in myeloid leukocytes protects against sepsis. Am. J. Respir. Crit. Care Med. 2011, 184, 928–938. [Google Scholar] [CrossRef] [PubMed]
  37. Okawa, H.; Motohashi, H.; Kobayashi, A.; Aburatani, H.; Kensler, T.W.; Yamamoto, M. Hepatocyte-specific deletion of the keap1 gene activates Nrf2 and confers potent resistance against acute drug toxicity. Biochem. Biophys. Res. Commun. 2006, 339, 79–88. [Google Scholar] [CrossRef]
  38. Brown, P.J.; Stuart, L.W.; Hurley, K.P.; Lewis, M.C.; Winegar, D.A.; Wilson, J.G.; Wilkison, W.O.; Ittoop, O.R.; Willson, T.M. Identification of a subtype selective human PPARα agonist through parallel-array synthesis. Bioorg. Med. Chem. Lett. 2001, 11, 1225–1227. [Google Scholar] [CrossRef]
  39. Kuma, A.; Hatano, M.; Matsui, M.; Yamamoto, A.; Nakaya, H.; Yoshimori, T.; Ohsumi, Y.; Tokuhisa, T.; Mizushima, N. The role of autophagy during the early neonatal starvation period. Nature 2004, 432, 1032–1036. [Google Scholar] [CrossRef] [PubMed]
  40. Santilli, A.A.; Scotese, A.C.; Tomarelli, R.M. A potent antihypercholesterolemic agent: (4-chloro-6-(2,3-xylidino)-2-pyrimidinylthio) acetic acid (Wy-14643). Experientia 1974, 30, 1110–1111. [Google Scholar] [CrossRef]
  41. Reddy, J.K.; Moody, D.E.; Azarnoff, D.L.; Tomarelli, R.M. Hepatic effects of some [4-chloro-6-(2,3-xylidino)-2-pyrimidinylthio] acetic acid (WY-14,643) analogs in the mouse. Arch. Int. Pharmacodyn. Ther. 1977, 225, 51–57. [Google Scholar]
  42. Komatsu, M.; Waguri, S.; Koike, M.; Sou, Y.S.; Ueno, T.; Hara, T.; Mizushima, N.; Iwata, J.; Ezaki, J.; Murata, S.; et al. Homeostatic levels of p62 control cytoplasmic inclusion body formation in autophagy-deficient mice. Cell 2007, 131, 1149–1163. [Google Scholar] [CrossRef]
  43. McMahon, M.; Itoh, K.; Yamamoto, M.; Chanas, S.A.; Henderson, C.J.; McLellan, L.I.; Wolf, C.R.; Cavin, C.; Hayes, J.D. The Cap’n’Collar basic leucine zipper transcription factor Nrf2 (NF-E2 p45-related factor 2) controls both constitutive and inducible expression of intestinal detoxification and glutathione biosynthetic enzymes. Cancer Res. 2001, 61, 3299–3307. [Google Scholar] [PubMed]
  44. Wilms, H.; Sievers, J.; Rickert, U.; Rostami-Yazdi, M.; Mrowietz, U.; Lucius, R. Dimethylfumarate inhibits microglial and astrocytic inflammation by suppressing the synthesis of nitric oxide, IL-1beta, TNF-α and IL-6 in an in-vitro model of brain inflammation. J. Neuroinflamm. 2010, 7, 30. [Google Scholar] [CrossRef] [PubMed]
  45. Itoh, K.; Chiba, T.; Takahashi, S.; Ishii, T.; Igarashi, K.; Katoh, Y.; Oyake, T.; Hayashi, N.; Satoh, K.; Hatayama, I.; et al. An Nrf2/small Maf heterodimer mediates the induction of phase II detoxifying enzyme genes through antioxidant response elements. Biochem. Biophys. Res. Commun. 1997, 236, 313–322. [Google Scholar] [CrossRef] [PubMed]
  46. Taguchi, K.; Maher, J.M.; Suzuki, T.; Kawatani, Y.; Motohashi, H.; Yamamoto, M. Genetic analysis of cytoprotective functions supported by graded expression of Keap1. Mol. Cell. Biol. 2010, 30, 3016–3026. [Google Scholar] [CrossRef] [PubMed]
  47. Iershov, A.; Nemazanyy, I.; Alkhoury, C.; Girard, M.; Barth, E.; Cagnard, N.; Montagner, A.; Chretien, D.; Rugarli, E.I.; Guillou, H.; et al. The class 3 PI3K coordinates autophagy and mitochondrial lipid catabolism by controlling nuclear receptor PPARα. Nat. Commun. 2019, 10, 1566. [Google Scholar] [CrossRef] [PubMed]
  48. Saito, T.; Kuma, A.; Sugiura, Y.; Ichimura, Y.; Obata, M.; Kitamura, H.; Okuda, S.; Lee, H.C.; Ikeda, K.; Kanegae, Y.; et al. Autophagy regulates lipid metabolism through selective turnover of NCoR1. Nat. Commun. 2019, 10, 1567. [Google Scholar] [CrossRef]
  49. Chiang, S.K.; Chen, S.E.; Chang, L.C. A Dual Role of Heme Oxygenase-1 in Cancer Cells. Int. J. Mol. Sci. 2018, 20, 39. [Google Scholar] [CrossRef]
  50. Stec, D.E.; John, K.; Trabbic, C.J.; Luniwal, A.; Hankins, M.W.; Baum, J.; Hinds, T.D., Jr. Bilirubin Binding to PPARα Inhibits Lipid Accumulation. PLoS ONE 2016, 11, e0153427. [Google Scholar] [CrossRef]
  51. Gordon, D.M.; Hong, S.H.; Kipp, Z.A.; Hinds, T.D., Jr. Identification of Binding Regions of Bilirubin in the Ligand-Binding Pocket of the Peroxisome Proliferator-Activated Receptor-A (PPARα). Molecules 2021, 26, 2975. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Autophagy inhibition impairs pharmacologic PPAR𝛼 transactivation. (a) A schematic diagram of an experimental procedure in mice. Fed or 24 h fasted Atg7F/F, and Atg7LKO mice were intraperitoneally injected with vehicle (0.1% DMSO in 4:1 ratio of PEG 400 and Tween 80) or GW7647, a synthetic PPAR𝛼 agonist (10 mg/kg BW) twice a day. A total of 5 h after last treatment, all mice were killed to collect livers, of which total RNAs were isolated for qPCR analysis. (b) Hepatic expression levels of PPAR𝛼 target genes Acox1, Acot3, Fgf21, Pdk4, Acot1, and Cidec were determined in Atg7F/F and Atg7LKO mice shown in panel (a) by qPCR analysis. n = 4 per group, ** p < 0.01 vs. Atg7F/F mice treated with vehicle. ## p < 0.01. Data represent mean ± s.e.m. and are plotted as fold change. Each dot indicates an individual mouse. Statistics by two-tailed t-test. (c) A schematic diagram of an experimental procedure in AML12 cells, a mouse hepatocyte-derived cell line. AML12 cells were treated with bafilomycin A1 (BafA1: 0.1, 0.5, or 1 μM), a known autophagy inhibitor in the absence or presence of synthetic PPAR𝛼 agonists (1 μM of GW764 or 100 μM of Wy-14,643) for 24 h. Vehicle is 0.1% DMSO. Total RNAs from these cells were prepared to perform qPCR analysis. (d,e) Expression levels of PPAR𝛼 target genes Pdk4, Acot3, and Ucp2, were determined in AML12 cells shown in panel (c) by qPCR analysis. n = 3 per group, ** p < 0.01 vs. AML12 cells treated with vehicle. ## p < 0.01 vs. AML12 cells treated with GW7647 or Wy-14,643. Data represent mean ± s.e.m. and are plotted as fold change. Each dot indicates an individual sample. Statistics by a two-tailed, unpaired Student t-test. BW, body weight.
Figure 1. Autophagy inhibition impairs pharmacologic PPAR𝛼 transactivation. (a) A schematic diagram of an experimental procedure in mice. Fed or 24 h fasted Atg7F/F, and Atg7LKO mice were intraperitoneally injected with vehicle (0.1% DMSO in 4:1 ratio of PEG 400 and Tween 80) or GW7647, a synthetic PPAR𝛼 agonist (10 mg/kg BW) twice a day. A total of 5 h after last treatment, all mice were killed to collect livers, of which total RNAs were isolated for qPCR analysis. (b) Hepatic expression levels of PPAR𝛼 target genes Acox1, Acot3, Fgf21, Pdk4, Acot1, and Cidec were determined in Atg7F/F and Atg7LKO mice shown in panel (a) by qPCR analysis. n = 4 per group, ** p < 0.01 vs. Atg7F/F mice treated with vehicle. ## p < 0.01. Data represent mean ± s.e.m. and are plotted as fold change. Each dot indicates an individual mouse. Statistics by two-tailed t-test. (c) A schematic diagram of an experimental procedure in AML12 cells, a mouse hepatocyte-derived cell line. AML12 cells were treated with bafilomycin A1 (BafA1: 0.1, 0.5, or 1 μM), a known autophagy inhibitor in the absence or presence of synthetic PPAR𝛼 agonists (1 μM of GW764 or 100 μM of Wy-14,643) for 24 h. Vehicle is 0.1% DMSO. Total RNAs from these cells were prepared to perform qPCR analysis. (d,e) Expression levels of PPAR𝛼 target genes Pdk4, Acot3, and Ucp2, were determined in AML12 cells shown in panel (c) by qPCR analysis. n = 3 per group, ** p < 0.01 vs. AML12 cells treated with vehicle. ## p < 0.01 vs. AML12 cells treated with GW7647 or Wy-14,643. Data represent mean ± s.e.m. and are plotted as fold change. Each dot indicates an individual sample. Statistics by a two-tailed, unpaired Student t-test. BW, body weight.
Cells 11 00754 g001
Figure 2. SFN-mediated NRF2 activation in a dose-dependent manner impairs pharmacologic PPAR𝛼 transactivation. (a) A schematic diagram of an experimental procedure in AML12 cells. AML12 cells were treated with sulforaphane (SFN, 12.5, 25, or 50 μM), a known NRF2 activator in a dose-dependent manner in the absence or presence of synthetic PPAR𝛼 agonists (GW7647, 0.01, 0.1 or 1 μM; Wy-14,643, 1, 10 or 100 μM) for 24 h. Vehicle is 0.1% DMSO. Total RNAs from these cells were prepared to perform qPCR analysis. (b,c) Expression levels of NRF2 target genes Nqo1 and Hmox1 were determined in AML12 cells shown in panel (a) by qPCR analysis. n = 3 per group, ** p < 0.01 vs. AML12 cells treated with vehicle. Data represent mean ± s.e.m. and are plotted as fold change. Each dot indicates an individual well. (d,e) Expression levels of PPAR𝛼 target genes Pdk4, Acot3, and Ucp2, were determined in AML12 cells shown in panel (a) by qPCR analysis. n = 3 per group, ** p < 0.01 vs. AML12 cells treated with vehicle. ## p < 0.01 vs. AML12 cells treated with 1 μM GW7647 or 100 μM Wy-14,643. Data represent mean ± s.e.m. and are plotted as fold change. Each dot indicates an individual sample. Statistics by a two-tailed, unpaired Student t-test.
Figure 2. SFN-mediated NRF2 activation in a dose-dependent manner impairs pharmacologic PPAR𝛼 transactivation. (a) A schematic diagram of an experimental procedure in AML12 cells. AML12 cells were treated with sulforaphane (SFN, 12.5, 25, or 50 μM), a known NRF2 activator in a dose-dependent manner in the absence or presence of synthetic PPAR𝛼 agonists (GW7647, 0.01, 0.1 or 1 μM; Wy-14,643, 1, 10 or 100 μM) for 24 h. Vehicle is 0.1% DMSO. Total RNAs from these cells were prepared to perform qPCR analysis. (b,c) Expression levels of NRF2 target genes Nqo1 and Hmox1 were determined in AML12 cells shown in panel (a) by qPCR analysis. n = 3 per group, ** p < 0.01 vs. AML12 cells treated with vehicle. Data represent mean ± s.e.m. and are plotted as fold change. Each dot indicates an individual well. (d,e) Expression levels of PPAR𝛼 target genes Pdk4, Acot3, and Ucp2, were determined in AML12 cells shown in panel (a) by qPCR analysis. n = 3 per group, ** p < 0.01 vs. AML12 cells treated with vehicle. ## p < 0.01 vs. AML12 cells treated with 1 μM GW7647 or 100 μM Wy-14,643. Data represent mean ± s.e.m. and are plotted as fold change. Each dot indicates an individual sample. Statistics by a two-tailed, unpaired Student t-test.
Cells 11 00754 g002
Figure 3. SFN-mediated NRF2 activation in a time-dependent manner impairs pharmacologic PPAR𝛼 transactivation. (a) A schematic diagram of an experimental procedure in AML12 cells. AML12 cells were treated with 25 μM of SFN, a known NRF2 activator in a time-dependent manner (6, 12, or 18 h) in the absence or presence of synthetic PPAR𝛼 agonists (GW7647, 0.01, 0.1 or 1 μM; Wy-14,643, 1, 10 or 100 μM) for 24 h. Vehicle is 0.1% DMSO. Total RNAs from these cells were prepared to perform qPCR analysis. (b,c) Expression levels of NRF2 target genes Nqo1 and Hmox1 were determined in AML12 cells shown in panel (a) by qPCR analysis. (d) Expression levels of PPAR𝛼 target genes Pdk4, Acot3, and Ucp2, were determined in AML12 cells shown in panel (a) by qPCR analysis. n = 3 per group, ** p < 0.01 vs. AML12 cells treated with vehicle. ## p < 0.01 vs. AML12 cells treated with GW7647 or Wy-14,643. Data represent mean ± s.e.m. and are plotted as fold change. Each dot indicates an individual sample. Statistics by a two-tailed, unpaired Student t-test.
Figure 3. SFN-mediated NRF2 activation in a time-dependent manner impairs pharmacologic PPAR𝛼 transactivation. (a) A schematic diagram of an experimental procedure in AML12 cells. AML12 cells were treated with 25 μM of SFN, a known NRF2 activator in a time-dependent manner (6, 12, or 18 h) in the absence or presence of synthetic PPAR𝛼 agonists (GW7647, 0.01, 0.1 or 1 μM; Wy-14,643, 1, 10 or 100 μM) for 24 h. Vehicle is 0.1% DMSO. Total RNAs from these cells were prepared to perform qPCR analysis. (b,c) Expression levels of NRF2 target genes Nqo1 and Hmox1 were determined in AML12 cells shown in panel (a) by qPCR analysis. (d) Expression levels of PPAR𝛼 target genes Pdk4, Acot3, and Ucp2, were determined in AML12 cells shown in panel (a) by qPCR analysis. n = 3 per group, ** p < 0.01 vs. AML12 cells treated with vehicle. ## p < 0.01 vs. AML12 cells treated with GW7647 or Wy-14,643. Data represent mean ± s.e.m. and are plotted as fold change. Each dot indicates an individual sample. Statistics by a two-tailed, unpaired Student t-test.
Cells 11 00754 g003
Figure 4. NRF2 activation impairs pharmacologic PPAR𝛼 transactivation in vivo. (a) A schematic diagram of an experimental procedure in mice. The 8- to 9-week-old male C57BL/6J wild-type mice were orally gavaged with vehicle or butylated hydroxyanisole (BHA, 200 mg/kg BW) once a day for 3 days. During the last 24 h, all mice were intraperitoneally injected with vehicle (0.1% DMSO in 4:1 ratio of PEG 400 and Tween 80) or GW7647, a synthetic PPAR𝛼 agonist (10 mg/kg BW) twice a day (first injection: 00:00 am, second injection: 12:00 pm). A total of 5 h after last treatment, all mice were killed to collect livers of which total RNAs were prepared for qPCR analysis. (b) Hepatic expression levels of PPAR𝛼 target genes Pdk4, Upc2, Acot3, Acox1, and Fgf21 were determined as shown in panel (a) by qPCR analysis. n = 8 per group, * p < 0.05, ** p < 0.01 vs. WT treated with vehicle. (c) A schematic diagram of an experimental procedure in mice. The 8- to 9-week-old male Alb-Cre/+ and Keap1LKO mice were intraperitoneally injected with vehicle (0.1% DMSO in 4:1 ratio of PEG 400 and Tween 80) or GW7647 (10 mg/kg BW) twice a day. A total of 5 h after last treatment, all mice were killed to collect livers, of which total RNAs were isolated for qPCR analysis. (d) Hepatic expression levels of Keap1, NRF2 target genes (Nqo1 and Gsta1), and PPAR𝛼 target genes Pdk4, Upc2, Cidec, Acot1 to 3, and Acox1 were determined shown in panel (c) by qPCR analysis. n = 4 per group, ** p < 0.01 vs. Alb-Cre/+ treated with vehicle. # p < 0.05, ## p < 0.01. Data represent mean ± s.e.m. and are plotted as fold change. Each dot indicates an individual mouse. Statistics by a two-tailed, unpaired Student t-test. WT, wild-type.
Figure 4. NRF2 activation impairs pharmacologic PPAR𝛼 transactivation in vivo. (a) A schematic diagram of an experimental procedure in mice. The 8- to 9-week-old male C57BL/6J wild-type mice were orally gavaged with vehicle or butylated hydroxyanisole (BHA, 200 mg/kg BW) once a day for 3 days. During the last 24 h, all mice were intraperitoneally injected with vehicle (0.1% DMSO in 4:1 ratio of PEG 400 and Tween 80) or GW7647, a synthetic PPAR𝛼 agonist (10 mg/kg BW) twice a day (first injection: 00:00 am, second injection: 12:00 pm). A total of 5 h after last treatment, all mice were killed to collect livers of which total RNAs were prepared for qPCR analysis. (b) Hepatic expression levels of PPAR𝛼 target genes Pdk4, Upc2, Acot3, Acox1, and Fgf21 were determined as shown in panel (a) by qPCR analysis. n = 8 per group, * p < 0.05, ** p < 0.01 vs. WT treated with vehicle. (c) A schematic diagram of an experimental procedure in mice. The 8- to 9-week-old male Alb-Cre/+ and Keap1LKO mice were intraperitoneally injected with vehicle (0.1% DMSO in 4:1 ratio of PEG 400 and Tween 80) or GW7647 (10 mg/kg BW) twice a day. A total of 5 h after last treatment, all mice were killed to collect livers, of which total RNAs were isolated for qPCR analysis. (d) Hepatic expression levels of Keap1, NRF2 target genes (Nqo1 and Gsta1), and PPAR𝛼 target genes Pdk4, Upc2, Cidec, Acot1 to 3, and Acox1 were determined shown in panel (c) by qPCR analysis. n = 4 per group, ** p < 0.01 vs. Alb-Cre/+ treated with vehicle. # p < 0.05, ## p < 0.01. Data represent mean ± s.e.m. and are plotted as fold change. Each dot indicates an individual mouse. Statistics by a two-tailed, unpaired Student t-test. WT, wild-type.
Cells 11 00754 g004
Figure 5. NRF2 activation is necessary and sufficient for suppressing pharmacologic PPAR𝛼 transactivation. (a) A schematic diagram of an experimental procedure in mice. The 8- to 9-week-old male Nrf2F/F and Nrf2LKO mice were orally gavaged with vehicle or butylated hydroxyanisole (BHA, 200 mg/kg BW) once a day for 3 days. During the last 24 h, all mice were intraperitoneally injected with vehicle (0.1% DMSO in 4:1 ratio of PEG 400 and Tween 80) or GW7647, a synthetic PPAR𝛼 agonist (10 mg/kg BW) twice a day (first injection: 00:00 am, second injection: 12:00 pm). A total of 5 h after last treatment, all mice were killed to collect livers of which total RNAs were prepared for qPCR analysis. (b) Hepatic expression levels of Nrf2 and its target genes Nqo1, Gsta1, and Hmox1 were determined in AML12 cells shown in panel (a) by qPCR analysis, * p < 0.05. (c) Hepatic expression levels of PPAR𝛼 target genes Pdk4, Upc2, Cidec, and Acot1 to 3 were determined as shown in panel (a) by qPCR analysis. n = 3–4 per group, ** p < 0.01 vs. Nrf2F/F treated with vehicle. # p < 0.05, ## p < 0.01. Data represent mean ± s.e.m. and are plotted as fold change. Each dot indicates an individual mouse. Statistics by a two-tailed, unpaired Student t-test.
Figure 5. NRF2 activation is necessary and sufficient for suppressing pharmacologic PPAR𝛼 transactivation. (a) A schematic diagram of an experimental procedure in mice. The 8- to 9-week-old male Nrf2F/F and Nrf2LKO mice were orally gavaged with vehicle or butylated hydroxyanisole (BHA, 200 mg/kg BW) once a day for 3 days. During the last 24 h, all mice were intraperitoneally injected with vehicle (0.1% DMSO in 4:1 ratio of PEG 400 and Tween 80) or GW7647, a synthetic PPAR𝛼 agonist (10 mg/kg BW) twice a day (first injection: 00:00 am, second injection: 12:00 pm). A total of 5 h after last treatment, all mice were killed to collect livers of which total RNAs were prepared for qPCR analysis. (b) Hepatic expression levels of Nrf2 and its target genes Nqo1, Gsta1, and Hmox1 were determined in AML12 cells shown in panel (a) by qPCR analysis, * p < 0.05. (c) Hepatic expression levels of PPAR𝛼 target genes Pdk4, Upc2, Cidec, and Acot1 to 3 were determined as shown in panel (a) by qPCR analysis. n = 3–4 per group, ** p < 0.01 vs. Nrf2F/F treated with vehicle. # p < 0.05, ## p < 0.01. Data represent mean ± s.e.m. and are plotted as fold change. Each dot indicates an individual mouse. Statistics by a two-tailed, unpaired Student t-test.
Cells 11 00754 g005
Figure 6. Working model of basal autophagy defects or NRF2 activators that impair pharmacological transactivation of the nuclear receptor PPARα. Similar to NRF2 activators, autophagy inhibition by pharmacologic or genetic manipulations liberates NRF2 from KEAP1-mediated proteasomal degradation (1, 2). This event leads to the nuclear translocation of NRF2 from the cytoplasm, which acts as a potent antioxidant transcription factor regulating the expression of many cytoprotective genes (3). On the other hand, the nuclear NRF2 interferes with a pharmacologic nuclear receptor signaling, including PPARα (4). ARE, antioxidant response element; PPRE, peroxisome proliferator response element; DR1, direct repeat 1. This schematic diagram was created in BioRender.com (accessed on 6 January 2022).
Figure 6. Working model of basal autophagy defects or NRF2 activators that impair pharmacological transactivation of the nuclear receptor PPARα. Similar to NRF2 activators, autophagy inhibition by pharmacologic or genetic manipulations liberates NRF2 from KEAP1-mediated proteasomal degradation (1, 2). This event leads to the nuclear translocation of NRF2 from the cytoplasm, which acts as a potent antioxidant transcription factor regulating the expression of many cytoprotective genes (3). On the other hand, the nuclear NRF2 interferes with a pharmacologic nuclear receptor signaling, including PPARα (4). ARE, antioxidant response element; PPRE, peroxisome proliferator response element; DR1, direct repeat 1. This schematic diagram was created in BioRender.com (accessed on 6 January 2022).
Cells 11 00754 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kim, E.Y.; Lee, J.M. Basal Autophagy Is Necessary for A Pharmacologic PPARα Transactivation. Cells 2022, 11, 754. https://doi.org/10.3390/cells11040754

AMA Style

Kim EY, Lee JM. Basal Autophagy Is Necessary for A Pharmacologic PPARα Transactivation. Cells. 2022; 11(4):754. https://doi.org/10.3390/cells11040754

Chicago/Turabian Style

Kim, Eun Young, and Jae Man Lee. 2022. "Basal Autophagy Is Necessary for A Pharmacologic PPARα Transactivation" Cells 11, no. 4: 754. https://doi.org/10.3390/cells11040754

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop