Next Article in Journal
TLR7 Expression Is Associated with M2 Macrophage Subset in Calcific Aortic Valve Stenosis
Previous Article in Journal
Revealing the Proteome of Motor Cortex Derived Extracellular Vesicles Isolated from Amyotrophic Lateral Sclerosis Human Postmortem Tissues
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Peroxisome Proliferator-Activated Receptors and Caloric Restriction—Common Pathways Affecting Metabolism, Health, and Longevity

1
Department of Nutritional Sciences, University of Vienna, 1090 Vienna, Austria
2
Toxalim (Research Centre in Food Toxicology), INRAE, ENVT, INP-Purpan, UMR 1331, UPS, Université de Toulouse, F-31027 Toulouse, France
3
Lee Kong Chian School of Medicine, Nanyang Technological University Singapore, Singapore 308232, Singapore
4
Center for Integrative Genomics, University of Lausanne, Le Génopode, CH-1015 Lausanne, Switzerland
*
Authors to whom correspondence should be addressed.
Cells 2020, 9(7), 1708; https://doi.org/10.3390/cells9071708
Submission received: 24 June 2020 / Revised: 14 July 2020 / Accepted: 14 July 2020 / Published: 16 July 2020
(This article belongs to the Section Cellular Aging)

Abstract

:
Caloric restriction (CR) is a traditional but scientifically verified approach to promoting health and increasing lifespan. CR exerts its effects through multiple molecular pathways that trigger major metabolic adaptations. It influences key nutrient and energy-sensing pathways including mammalian target of rapamycin, Sirtuin 1, AMP-activated protein kinase, and insulin signaling, ultimately resulting in reductions in basic metabolic rate, inflammation, and oxidative stress, as well as increased autophagy and mitochondrial efficiency. CR shares multiple overlapping pathways with peroxisome proliferator-activated receptors (PPARs), particularly in energy metabolism and inflammation. Consequently, several lines of evidence suggest that PPARs might be indispensable for beneficial outcomes related to CR. In this review, we present the available evidence for the interconnection between CR and PPARs, highlighting their shared pathways and analyzing their interaction. We also discuss the possible contributions of PPARs to the effects of CR on whole organism outcomes.

Graphical Abstract

1. Introduction

Caloric restriction (CR) is one of the primary interventions for weight loss and health maintenance. As early as the 16th century, Luigi Cornaro (1484–1566) described the beneficial effects of this approach in his “Discorsa della vita sobria”. Later, at the beginning of the 20th century, the first experimental evidence emerged when Osborne et al. reported that CR slowed the growth of rats but prolonged their lifespan [1]. In rats, a CR of 40% applied from weaning onward has been linked to a lifespan extension of almost two-fold [2]. In fact, CR has been associated with increases in mean and maximum lifespan, regardless of sex, in multiple species, including various rat and mouse strains, yeasts, worms, fruit flies, fishes, hamsters, dogs, cows, and owls [3]. The effects of CR in these organisms include reduced neurodegenerative disease incidence, diminished rates of age-specific mortality, and a lower incidence of cancer, diabetes, atherosclerosis, and cardiovascular disease. CR also is linked to delayed onset of age-related processes, such as immunosenescence, sarcopenia, and atrophy of the brain gray matter [3,4,5,6,7]. In monkeys, CR leads to diabetes suppression and a reduced incidence of neoplasia and cardiovascular diseases by up to 50% [6]. These effects have been attributed to a reduction in major risk factors, including cholesterol, C-reactive protein, blood pressure, and intima-media thickness of the carotid arteries [7,8,9]. Beneficial outcomes of CR have been consistently reported, which supports this approach considering that distinct CR protocols are used in different publications. CR applied in diverse studies ranges from 10% to up to 50% of daily caloric intake. Furthermore, the length of CR varies from a few weeks to life-long treatment. Additionally, some protocols restrict all nutrients, whereas others limit macronutrients only and supplement micronutrients in order to investigate selectively the impact of calorie reduction and prevent malnutrition, making a distinction between “dietary restriction” and “energy restriction” [10]. As expected, the type of CR protocol influences the magnitude of outcomes [11,12]. Moreover, the results obtained from experimental models cannot be directly translated to humans [13]. Therefore, it is important to compile the results of multiple studies to identify common patterns of responses regardless of the type of CR. A comparison of the responses from different species may help to draw a more comprehensive picture of the outcomes of CR.
CR has been tied to a complex network of pathways implicating insulin-like growth factor 1 (IGF-1), sirtuins (SIRTs), adenosine monophosphate (AMP)-activated protein kinase (AMPK), and target of rapamycin (TOR). The sympathetic and neuroendocrine systems, as well as thyroid hormones, adipokines, and ghrelin, also have been associated with the beneficial outcomes of CR [4]. This ensemble of processes associated with CR affects the whole body, manifesting in reduced inflammation, body fat mass, resting metabolic rate, and body temperature and improved insulin sensitivity [14]. As a result of the variety of outcomes related to CR and the complexity of the contributing pathways, the exact mechanisms underlying these health benefits are still not well understood. However, the results of ongoing studies have filled in some parts of this puzzle. This review focuses on one important piece: the role of peroxisome proliferator-activated receptors (PPARs) in outcomes associated with CR.

2. PPARs

PPARs form a subfamily of the ligand-dependent nuclear receptor superfamily, which consists of three isotypes, PPARα (NR1C1), PPARβ/δ (NR1C2), and PPARγ (NR1C3), each coded by a separate gene [15,16,17]. PPARs play major roles in a broad spectrum of biological processes, including cell proliferation and differentiation, fatty acid (FA) and eicosanoid signaling, bone formation, tissue repair and remodeling, insulin sensitivity, and, in particular, glucose and lipid metabolism (Figure 1). They are transcription factors that form heterodimers with retinoid X receptors (RXRs) and bind to specific peroxisome proliferator response elements (PPREs) in the transcription regulatory region of their target genes. A number of coactivators and corepressors modulate PPAR activity, either stimulating or inhibiting receptor function [18]. Two important PPAR corepressors are the nuclear corepressor 1 (NCoR1) and the silencing mediator of retinoic acid and thyroid hormone receptor (SMRT) [19,20,21], which are recruited to PPARs in the absence of ligands and limit PPAR transcriptional activity [22,23]. Coactivators encompass various proteins, including factors with histone acetylase activity [cAMP response element-binding (CREB) protein (CBP)/p300 and steroid receptor coactivator 1–3 complexes], helicases [PPAR A–interacting complex (Pric)285, Pric320/chromodomain helicase DNA binding protein 9], and an ATPase in the SWItch/sucrose non-fermentable (SWI/SNF) complex, and non-enzymatic activators that can be found in the active PPAR transcriptional complex [PPARγ coactivator (PGC)-1α, PGC-δ/PGC-1–related estrogen receptor coactivator, mediator of RNA polymerase II transcription subunit/TRAP220/PPAR-binding protein, PPAR-interacting protein/nuclear receptor coactivator 6, SWI/SNF-related, matrix-associated, actin-dependent regulator of chromatin, subfamily D, member 1] [18]. The characteristic feature of the PPAR ligand-binding cavity is its 3-4-fold larger size compared to other nuclear receptors. Thus, PPARs can accommodate and bind a variety of natural and synthetic lipophilic acids [24,25,26,27]. Synthetic ligands of PPARs are widely used in clinical practice to treat glucose and lipid disorders and in the prevention and treatment of cardiovascular and metabolic diseases [28,29,30]. Synthetic ligands can be specific for each PPAR isotype or activate two (saroglitazar, elafibranor) [31,32] or all three (bezafibrate, lanifibranor) isotypes [33,34]. Natural ligands of PPARs include various FAs, phospholipids, prostaglandins, prostacyclins, and leukotrienes [35,36] linking the activity of PPARs to nutrition, metabolism, and inflammation. In addition to ligands, PPAR transcriptional activity can be modulated by post-translational modifications including phosphorylation, ubiquitination, O-GlcNAcylation, and SUMOylation [29,37,38,39,40,41,42,43].

2.1. PPARα

The first cloned PPAR, now known as PPARα, was initially identified as the molecular target of xenobiotics inducing hepatic peroxisome proliferation in rodents [44]. PPARα is particularly abundant in organs with high demand for FA catabolism, such as liver and brown adipose tissue (BAT) [45]. In these tissues, this transcription factor coordinates several aspects of metabolism by modulating the expression of genes involved in peroxisomal and mitochondrial β-oxidation, FA transport and catabolism, ketogenesis, and gluconeogenesis [35]. In line with this role of PPARα, the L162V Pparα gene polymorphism, which results in a PPARα variant with lower transcriptional activity, is associated with increased serum levels of triglycerides (TGs), total cholesterol, low-density lipoprotein cholesterol, and apolipoprotein (apo) A1 and apoB [46,47,48,49,50], influencing the onset and progress of type 2 diabetes (T2D) [51,52].
PPARα serves as a sensor of nutritional status. In starvation, when the organism mobilizes stored FAs, PPARα in hepatocytes increases FA uptake and β-oxidation and stimulates hepatokine gene expression [53,54,55,56]. Therefore, PPARα adapts the rates of FA catabolism and ketone body synthesis to energy depletion. Furthermore, in the mouse liver, PPARα is a key factor in the metabolic adaptation to sepsis. Deficiency in hepatic PPARα causes an impaired metabolic response, and upon bacterial infection, PPARα-null mice have a higher mortality rate. These observations suggest that activating PPARα may be considered a plausible metabolic intervention for improving sepsis outcomes. In support of this hypothesis, non-surviving critically ill patients have reduced PPARα activity in their livers [57].
The natural ligands of PPARα are FAs and their derivatives [58]. Synthetic fibrate PPARα agonists are generally used to treat patients with primary hypercholesterolemia, mixed dyslipidemia, and hypertriglyceridemia. Natural and synthetic PPARα agonists have been discussed extensively [29,30,59,60], and their effect appears to be based mostly on the stimulation of cellular lipid trafficking and β-oxidation [61,62]. However, the need persists for better drug candidates that selectively target PPARα without the side effects of fibrates. The recent development of selective PPARα modulators may fill this gap [63].

2.2. PPARβ/δ

PPARβ/δ is ubiquitously expressed, with particularly high levels in the skin, gastrointestinal tract, liver, kidney, and various parts of the central nervous system [64,65]. PPARβ/δ is involved in cell proliferation, differentiation, and survival, and it plays a role in tissue repair [66,67,68,69]. It is necessary for placental and gut development and is involved in the control of energy homeostasis [70,71,72,73,74]. Metabolically, PPARβ/δ promotes FA oxidation in adipose tissue and skeletal muscle, leading to improved lipidemia, and it stimulates overall energy expenditure, protecting against diet-induced obesity and insulin resistance [75,76,77,78]. It is probably best known for stimulating energy expenditure in muscles and increasing exercise capacity [79]. Under regular exercise, PPARβ/δ induces a switch to an increased amount of type I muscle fibers, enhancing mitochondrial function and fat oxidation and boosting endurance [80]. For this reason, this receptor also is a target for doping substances, so much so that the World Anti-Doping Agency added the PPARβ/δ ligands GW501516 and other related chemicals to its prohibited list in 2009. These synthetic ligands (GW501516, GW0742, and L-165041) selectively activate PPARβ/δ at very low concentrations and are often applied in both in vivo and in vitro research models [81], but no PPARβ/δ agonist has been used in clinical practice. Nevertheless, PPARβ/δ remains a potential therapeutic target in metabolic diseases. Other synthetic and natural PPARβ/δ agonists recently have been discussed elsewhere [29].

2.3. PPARγ

PPARγ is recognized primarily for its insulin-sensitizing properties and its role as a master regulator of adipogenesis [82,83,84]. It also has been identified as a promising therapeutic target for several types of cancer because it limits tumor progression by decreasing cell proliferation [85,86,87], increasing cell differentiation [85,88], inducing apoptosis [85,86,89,90,91,92,93], and inhibiting angiogenesis [94]. In addition, PPARγ plays a role in multiple processes including inflammation, adipogenesis, FA storage, and lipid and glucose metabolism [82,95,96,97,98,99]. We have described novel roles for intestinal PPARγ in long-chain FA processing in the intestinal epithelium [100] and in the regulation of body adiposity through sympathetic nervous system signaling [101], and we identified it as a major regulator of mucosal defenses upon high-fat diet feeding in mice [102]. However, its effects are not exclusively beneficial. For instance, it has recently been reported that PPARγ has a carcinogenic effect in advanced brain metastases [103] and liver cancer [104].
PPARγ occurs in two isoforms, PPARγ1 and PPARγ2, with the latter distinguished by 30 and 28 extra amino acids at the N-terminus in mouse and human, respectively [105]. PPARγ2 is found at high levels in adipose tissues [106,107], whereas PPARγ1 shows a broader expression pattern. In addition to adipose tissue, PPARγ1 is expressed in the gut, brain, and vascular cells and in specific types of immune and inflammatory cells [65,105,108,109,110]. A plethora of factors induce PPARγ expression and activity, and not all are ligands. They can include FAs and their metabolites [25,111], eicosanoids [25], prostaglandins [112], phytanic acid [113], and multiple nutrients, along with glutamine, curcumin, capsaicin, ginsenosides, and vitamin E, all of which have been reported to exhibit anti-inflammatory properties [114]. Of note, the presence of specific bacterial strains, their metabolites, and bacterial by-products [115,116,117,118,119,120] also stimulate PPARγ expression and/or activity.
Synthetic agonists of PPARγ can be divided into two groups: classical full agonists, which are represented by the thiazolidinediones (TZDs) [121] such as troglitazone, rosiglitazone, and pioglitazone, and partial agonists, which were developed to reduce the side effects of full agonists, including weight gain and bone loss [122,123]. PPARα/γ dual agonists exert positive influences on both lipid and glucose metabolism. They not only have antidiabetic capacity but also are hypolipemic, anti-inflammatory, hypotensive, and antiatherogenic and show anticoagulant action leading to improved endothelial function [30,31,59,124,125,126,127,128,129,130].

2.4. PPARs in CR

The effects of CR on PPAR expression vary among different tissues and organs. PPAR expression either remains unchanged [PPARα, -β/δ, and -γ in the heart, PPARα in white adipose tissue (WAT), PPARγ in the liver], decreases (PPARβ/δ in the liver, PPARα in the spleen, PPARα, -β/δ, and -γ in muscle), or increases (PPARα in the liver, heart, and intestinal epithelium) [131,132,133,134,135,136,137,138,139]. In rat kidneys, PPAR mRNA and protein levels and DNA-binding activity decrease with age, but a 60% CR blunts this reduction [140].
PPARs can contribute to CR-related outcomes by modulating multiple pathways connected with metabolism, insulin sensitivity, inflammation, and oxidative stress, all of which we describe at the molecular level in this review. Most of the published works on the impact of PPARs in CR concern PPARα. CR induces the upregulation of PPARα expression in the heart, where it increases lipid use [139]. In fact, 19% of the hepatic genes involved in lipid metabolism, inflammation, and cell growth and whose expression is changed by CR depend on PPARα, as shown by the absence of regulation of these processes in PPARα-null mice [141]. Therefore, ligands of PPARα have been proposed to serve as mimetics of CR [141]. Although less studied so far, the two other PPAR isotypes likewise modulate multiple pathways in relation to CR, as we discuss below.

3. Major Pathways Affected by CR

3.1. mTOR

The mammalian (m)TOR pathway is a major nutrient sensor signaling pathway known to regulate longevity (Figure 2). TOR is a well-conserved Ser/Thr protein kinase that belongs to the family of phosphatidylinositol 3 (PI3) kinase-related kinases [142,143]. It functions as an essential part of two complexes, mTORC1 and mTORC2, which have some proteins in common and some different proteins between them [142]. mTORC1 comprises the following core subunits: mTOR, mLST8 (mammalian lethal with sec-13 or GβL), DEPTOR (DEP domain-containing mTOR-interacting protein Tti1/Tel2 complex), PRAS40 (proline-rich Akt substrate of 40 kDa), and Raptor (regulatory-associated protein of mammalian target of rapamycin). mTORC2 is composed of mTOR, mLST8, DEPTOR, the Tti1/Tel2 complex, Rictor (rapamycin-insensitive companion of mTOR), mSin1 (mammalian stress-activated MAP kinase-interacting protein 1 or MAPKAP1), and protor1/2 (protein observed with Rictor 1 and 2) [144,145,146,147]. The configuration of each of these two complexes is conserved from yeast to mammals [148]. mTORC1 is sensitive to inhibition by rapamycin and plays essential roles in the regulation of mRNA translation and autophagy. Cellular energy and nutrient status regulate it directly, whereas mTORC2, which is not rapamycin sensitive, functions mainly as an important regulator of the cellular actin cytoskeleton [149,150].
Rheb (Ras homolog enriched in the brain) is a GTPase that in its GTP-bound form directly binds to and activates mTOR [151,152,153,154]. Rheb activity is inhibited by the heterodimer complex of tuberous sclerosis proteins 1 and 2 (TSC1 and TSC2) [155,156,157,158,159,160]. TSC1/2 mediates for mTORC1 many of the upstream signals from growth factors, such as insulin and IGF-1, which stimulate the PI3K and Ras pathways. The effector kinases of these pathways, including Akt (or protein kinase B or PKB), extracellular-signal-regulated kinase 1/2 (ERK1/2), and ribosomal protein S6 kinase β-1 (S6K1), directly phosphorylate and inactivate the TSC1/TSC2 complex, leading to the activation of mTORC1 [154,157,161,162,163,164,165,166,167]. Rheb also can transmit upstream signals from the p38β–PRAK pathway, which is activated upon glucose starvation [168]. Finally, as a core component of mTORC2, mTOR functions as a tyrosine protein kinase that promotes activation of the insulin receptor and IGF-1 receptors [169]. These interactions illustrate the tightly interconnected signaling between mTOR and insulin.
The mTOR pathway integrates inputs from major intracellular and extracellular physiological stimuli (growth factors, stress, energy balance, oxygen, amino acids) and controls many major downstream processes, including macromolecule synthesis, autophagy, cell cycle, growth, and metabolism [142,143,170]. For example, the canonical Wnt pathway, AMPK, some pro-inflammatory cytokines such as tumor necrosis factor-α (TNFα), and the hypoxia-inducible proteins REDD1 and REDD2 modulate mTORC1 activity via TSC1/2 [171,172,173,174,175,176]. In addition to phosphorylating TSC1/2, AMPK phosphorylates Raptor, leading to the allosteric inhibition of mTOR [177]. mTORC1 activity is further regulated by lipid-derived signaling molecules (phosphatidic acid) [178], the redox status of the cell [179], and amino acids, particularly leucine and arginine [180,181]. DNA damage also signals to mTORC1 through multiple mechanisms, all of which require p53-dependent transcription, induction of the expression of TSC2 and phosphatase and tensin homolog deleted on chromosome 10 (PTEN), and AMPK activation [182,183,184].
Downstream signaling of mTORC1 controls autophagy and energy metabolism, including the glycolytic flux, lipid synthesis [185,186,187,188], and cholesterol synthesis via the activation of sterol regulatory element-binding protein (SREBP) 1/2 [185,189,190]. mTORC1 also promotes anabolism in the fed state by controlling lipid metabolism in the liver through the modulation of Srebp1c expression, which is a regulator of lipogenesis and lipid storage [191,192].
Under mTORC1 regulation, mitochondrial DNA content and the expression of genes involved in oxidative metabolism increase. mTORC1 exerts this effect in part by mediating the nuclear association between PPARγ coactivator 1α (PGC-1α) and the transcription factor Yin-Yang 1, which positively regulates mitochondrial biogenesis and oxidative function [193] (see the section on mitochondrial function).
The activation of mTOR also leads to the phosphorylation of many target proteins related to the translational machinery and ribosome biogenesis, such as p70 ribosomal S6 kinase (S6K) and eukaryotic initiation factor 4E-binding protein (4E-BP) [170,194,195,196,197,198,199]. The regulation of protein metabolism also is a much-recognized function of mTOR. Amino acid activation of mTORC1 promotes protein synthesis via the activation of S6K and/or inhibition of 4E-BP, whereas the inactivation of mTORC1 promotes the degradation of damaged proteins and intracellular organelles via autophagy [200,201] (Figure 2).
mTORC2 functions mainly as an important regulator of the actin cytoskeleton through its stimulation of F-actin stress fibers, paxillin, RhoA, Rac1, Cdc42, and protein kinase C (PKC) α [146]. mTORC2 phosphorylates Akt [202,203] and thus affects metabolism and cell survival. mTORC2 also directly activates SGK1, which is a kinase controlling ion transport and growth [204]. Both Akt and SGK1 phosphorylate FoxO1/3a [205,206,207].
As a result of its role as an amino acid sensor, the TOR pathway has been proposed as a mediator of CR. The high activity of mTORC1 is a major driving force of aging, whereas the suppression of mTOR is tied to many of the benefits associated with CR, including lifespan extension [208,209,210,211], as has been demonstrated in yeast [208,212], worms [209], and flies [210]. Rapamycin treatment slightly extends the lifespan in flies subjected to CR [213]. In yeast, CR does not further extend the lifespan in the absence of TOR1, which is one of the two TOR genes in yeast, suggesting that TOR inhibition and CR promote lifespan via a common mechanism [208]. Similarly, in C. elegans, using RNA interference against TOR or autophagy genes in eat-2 mutant worms, which have impaired feeding behavior and are used as a genetic model for CR, does not extend the lifespan [214,215]. Furthermore, the inhibition of one of the principal targets of TOR signaling, S6K, extends the lifespan of eat-2 C. elegans [216]. Of note, mTOR activation in the rat’s brain results in reduced food intake by promoting the expression of the orexigenic neuropeptide Y and agouti-related peptide in the hypothalamus [217,218]. These data suggest that CR and TOR inhibition promote lifespan via overlapping pathways.

3.2. mTOR and PPARα

The various direct or indirect interactions between mTOR and PPARs have diverse effects on FA synthesis, glucose metabolism, oncogenesis, and immune system activation (Figure 3). First, mTORC1 regulates hepatic ketone body production in response to fasting [219]. mTORC1 activity is low during fasting, which correlates with increased PPARα activity. Consequently, mice with a constitutive activation of mTORC1 in the liver and with correspondingly low PPARα cannot induce ketogenesis when fasted [219]. As alluded to above, PPARα is particularly known for its role in fasting-triggered FA oxidation and lipid metabolism in the liver. The intertwined functions of mTORC1 and PPARα lie in the mTORC1 inhibition of PPARα activity during feeding and consequently blocking hepatic ketogenesis.
In the fed state, the insulin-dependent PI3K pathway activates mTORC1, promoting the cytoplasmic and nuclear localization of the PPARα corepressor NCoR1 and leading to the inhibition of PPARα activity and ketogenesis [219]. Fasting leads to decreased mTORC1 and consequently S6 kinase 2 activity, which promotes the cytoplasmic localization of NCoR1. In the absence of nuclear NCoR1, the increased transcriptional activity of PPARα enhances the FA oxidation that generates substrates for ketogenesis and gluconeogenesis [219,220]. In fact, PPARα also stimulates the expression of mitochondrial hydroxymethylglutaryl-CoA synthase (HMGCS), which is a rate-limiting enzyme of ketogenesis that catalyzes the condensation of acetyl coenzyme A (acetyl-CoA) and acetoacetyl-CoA to generate 3-hydroxy-3-methylglutaryl (HMG)-CoA and CoA [37]. The CR-triggered increase in the intestinal HMG-CoA synthase 2 (HMGCS2) expression alters the regeneration and differentiation capacity of intestinal steam cells. The result is a reduction in the differentiation markers of secretory cells with the promotion of enterocytes, changing the functionality of the intestine [221]. The subsequent refeeding of CR mice leads to reduced HMGCS2 levels and an increased expression of Paneth and goblet cell markers [221]. Additionally, in the intestine, the age-related increased activity of mTOR inhibits PPARα, resulting in higher levels of Notum and decreased Wnt signaling, consequently diminishing the regenerative function of stem cells in the Paneth cell niche [222]. Similarly, mTORC1 activity is elevated in the livers of old mice [219], which correlates with reduced PPARα activity and hepatic ketogenesis during aging [223,224,225]. mTORC1 inhibition is sufficient to prevent both the aging-induced loss of PPARα activity and ketone body production [219].
Another type of functional interaction between mTOR and PPARα relates to ligand production for the latter by FA synthase (FAS). In the fed state, mTORC1 mediates insulin-dependent phosphorylation and thus the inhibition of cytoplasmic FAS, limiting ligand generation. During fasting, when mTOR is inhibited, nonphosphorylated active cytoplasmic FAS promotes the synthesis of endogenous PPARα ligands [37].
In several organs, CR induces autophagy, which is a process that integrates mTOR and PPARα. To protect the liver from acute failure, PPARα-mediated induction of autophagy attenuates a lipopolysaccharide (LPS)-induced pro-inflammatory response [226]. Furthermore, agonists of PPARα (GW7647 and WY-14,643) regulate multiple genes involved in autophagy and lysosomal biogenesis and function, such as the transcription factor EB, which is a master gene for lysosomal biogenesis [227]. Of interest, a protein called farnesoid X receptor (FXR) is activated in the fed liver and suppresses autophagy. PPARα, activated in fasted and CR livers, regulates genomic circuits that are complementary to those under FXR control. Moreover, FXR stimulates the hepatic expression of PPARα [228]. These findings highlight how the liver senses nutrient status and how these two nuclear receptors translate this status in autophagy regulation [229,230].

3.3. mTOR and PPARβ/δ

Relatively little evidence connects mTOR and PPARβ/δ functions. In human lung carcinoma cells, nicotine activates PPARβ/δ expression through PI3K/mTOR [231], whereas the PPARβ/δ agonist GW501516 stimulates the growth of these cells through the inhibition of PTEN expression [232], indicating the interplay between the two pathways. Additionally, PPARβ/δ may modulate mTOR activity by mediating the metabolism of FAs and the production of phosphatidic acid, which is a metabolite that directly activates the mTOR complex by increasing its stability and activity. Phosphatidic acid responsiveness has been proposed as a lipid precursor sensing mechanism for the biosynthesis of cell membranes in the context of cell division and cell mass increase [233].

3.4. mTOR and PPARγ

As noted, PPARγ is a master regulator of adipogenesis. In parallel, mTORC1 senses growth factors and nutrients that drive adipose tissue accumulation. The inhibition of mTORC1 impairs adipogenesis and adipocyte maintenance in vitro [187,234,235,236,237], at least in part by modulating PPARγ expression and transcription [187,188,238,239]. mTORC1 may activate PPARγ through SREBP1, which promotes the production of endogenous PPARγ ligands [240,241]. Once activated by its natural or synthetic ligands, PPARγ stimulates mTORC1 and AMPK and upregulates TG-derived FA uptake, lipoprotein lipase activity, and accumulation in subcutaneous WAT and BAT. Chronic mTOR inhibition attenuates these processes, which leads to hyperlipidemia. These observations imply that mTOR regulates the hypolipidemic and lipogenic effects of PPARγ [239], as also suggested by the rapamycin inhibition of adipocyte differentiation [187,234,237]. Additionally, rapamycin reduces the phosphorylation of lipin-1 [242], which is a phosphatidic phosphatase that is involved in phospholipid and TG synthesis as well as the coactivation of many transcription factors linked to lipid metabolism, including PPARγ, PPARα, and PGC-1α [243,244,245]. A model has been proposed for nutrient and insulin signaling during adipogenesis in which the mTOR and PI3K/Akt pathways respectively act in parallel to regulate PPARγ activation [187], revealing pathway synergy in the context of metabolism.
Conversely, PPARγ also affects mTOR activity. It inhibits mTOR signaling with a protective action in rats with status epilepticus [246], and after activation with rosiglitazone, it stimulates mTORC1 and AMPK [239]. The interaction between PPARγ and mTOR has been particularly well demonstrated in mouse models of breast cancer, in which the activation of PPARγ led to rapamycin resistance, whereas rapamycin treatment triggered both the expression and activation of PPARγ [247].
Similar to PPARα, PPARγ regulates autophagy, the main cellular process targeted by mTOR. The activation of PPARγ stimulates autophagy, and its inhibition decreases it [248,249]. In addition to cell cycle arrest, the induction of autophagy is a major effect of PPARγ in cancer cells, as observed in colon [250], breast [251,252], bladder [253], and adrenocortical [254] cancer cell lines. The application of troglitazone and rosiglitazone in MDA-MB-231 breast cancer cells results in the redistribution of microtubule-associated protein 1A/1B-light chain 3 (LC3) and formation of the autophagosome [252]. Liver cells treated with 4-O-carboxymethyl ascochlorin, an agonist of PPARγ, undergo an autophagic response [249]. PPARγ activation in Caco-2 cells and other cell types affects autophagy via positive regulation of PTEN [255,256,257,258,259]. In contrast, the deletion of PTEN results in an increased expression of PPARγ in the liver [260]. In a model of cigarette smoking extract (CSE)-induced inflammation in a human bronchial epithelial cell line (16HBE), PPARγ plays an anti-inflammatory role in an autophagy-dependent manner. In this model, PPARγ activation promotes autophagy by phosphorylating and activating AMPK (see the AMPK and PPARγ section) [261]. Accordingly, the PPARγ agonists TZDs inhibit IGF-1-induced p70S6K phosphorylation at sites specifically phosphorylated by mTOR as well as dephosphorylating a downstream factor, S6K, through AMPK activation [262]. Similarly, PPARγ induces the expression of autophagy-related proteins but reduces inflammatory molecules in aged human dental pulp cells, suggesting a mechanism by which it might maintain pulp homeostasis [263].
Conversely, PPARγ exerts anti-autophagic effects in neurons [264,265] and macrophages [266]. Rosiglitazone downregulates autophagy-related protein expression and improves locomotor function after spinal cord injury (SCI), whereas GW9662, a PPARγ inhibitor, abolishes the protective effect in SCI [265]. The PPARγ agonist 15-PGJ2 counteracts the increased expression of the autophagy-related protein in the ischemic cortex following cerebral ischemia–reperfusion (I/R) injury. Moreover, cerebral I/R injury increases levels of LC3 and cathepsin-B in neurons, whereas 15-PGJ2 decreases them. Thus, PPARγ exerts neuroprotection by inhibiting neuronal autophagy after cerebral I/R injury [264]. Finally, in human monocyte-derived macrophages, PPARγ induces cathepsin L, which is a cysteine protease that has been found to trigger apoptosis and decrease autophagy [266].

4. AMPK

CR decreases energy input, which leads to the activation of a signaling cascade to generate fuel and increase longevity. Decreased glucose intake reduces carbon flow through the glycolytic pathway and slows the conversion of ADP to ATP. As a principal cellular energy sensor, AMPK monitors the AMP:ATP and ADP:ATP ratios. Functionally, AMPK is a serine/threonine kinase comprising one catalytic subunit, α, and two regulatory subunits, β and γ. Each of the subunits occurs as different isoforms (α1, α2, β1, β2, γ1, γ2, γ3) allowing for different versions of AMPK in various tissues [267,268]. From nematodes to humans, the kinase activity of AMPK is rapidly increased by the binding of AMP or ADP to the AMPKγ subunit [269]. This binding promotes allosteric activation and the phosphorylation of AMPK by the upstream AMPK kinase and thus also inhibits its dephosphorylation [270]. An alternative activating pathway triggers AMPK in response to increases in cellular Ca2+ and involves the Ca2+/calmodulin-dependent protein kinase kinase β (CaMKKβ) [271]. Once activated, AMPK promotes ATP preservation by repressing energy-consuming biosynthetic pathways while enhancing the expression or activity of proteins involved in catabolism. This process results in the mobilization of deposited energy to restore the ATP supply [272]. Several downstream factors including CREB-regulated transcriptional coactivator-2 (CRTC2) [273], TBC1D1/AS160 [274,275], PGC-1α [276], and histone deacetylase (HDAC) 5 [277] mediate the impact of AMPK on metabolism. Functionally, AMPK phosphorylates acetyl-CoA carboxylase 1 (ACC1) and ACC2 [278,279], SREBP1c [280], glycerol phosphate acyl-transferase, [281], and HMG-CoA reductase [282], resulting in the inhibition of FA, cholesterol, and TG synthesis while activating FA uptake and β-oxidation. Additionally, AMPK prevents protein biosynthesis by inhibiting mTOR and TIF-IA/RRN3, which is a transcription factor for RNA polymerase I that is responsible for ribosomal RNA synthesis [283]. AMPK also influences glucose metabolism by stimulating both nutrient-induced insulin secretion from pancreatic β-cells [284] and glucose uptake by phosphorylating Rab-GTPase-activating protein TBC1D1, which ultimately induces the fusion of glucose transporter (GLUT)4 vesicles with the plasma membrane in skeletal muscle [285]. AMPK stimulates glycolysis by the phosphorylation of 6-phosphofructo-2-kinase (fructose-2,6-bisphosphatase 2) [286], and in parallel, it inhibits glycogen synthesis through the phosphorylation of glycogen synthase [287]. In the liver, AMPK inhibits gluconeogenesis by inhibiting transcription factors including hepatocyte nuclear factor 4 and CRTC2 [288,289,290]. AMPK also affects the energy balance by regulating circadian metabolic activities and promoting feeding through its action in the hypothalamus [291,292]. It promotes mitochondrial biogenesis via PGC-1α [276] (see the section on mitochondria) and activates antioxidant defenses. AMPK plays a major role in metabolism but is also involved in inflammation, cell growth, autophagy, and apoptosis [293]. Therefore, reducing AMPK signaling exerts a cytostatic and tumor-suppressing effect [294,295].
In C. elegans, the lifespan extension effect of CR depends on AMPK [296,297]. Similarly, in Drosophila, pathways mediating increased lifespan include AMPK activation [298]. In addition, tissue-specific overexpression of AMPK in muscle and body fat extends the lifespan in Drosophila, whereas AMPK RNA interference shortens the lifespan [299]. The link between AMPK and PPARs and their interaction in metabolism regulation in response to CR have been well documented and are discussed below.

4.1. AMPK and PPARα

AMPK and PPARα both act as sensors of intracellular energy status and adjust metabolism in response to changes. As noted, AMPK responds to intracellular ATP depletion, whereas PPARα induces the expression of genes encoding enzymes and proteins involved in increasing cellular ATP yields. Moreover, AMPK and PPARα serve as critical regulators of short-term and long-term FA oxidation, respectively, and their activity thus needs to be coordinated. Accordingly, during prolonged fasting, when glucose levels drop and FA levels rise, high intracellular AMP concentrations induce AMPK, resulting in increased mitochondrial FA uptake for β-oxidation. In parallel, the activation of PPARα elevates the maximal FA-oxidizing capacity in the liver [35,37,300,301].
Similar to AMPK, phosphorylation affects the activity of PPARα. Several kinases, including p38, ERK, protein kinase A, and PKC, and AMPK itself can phosphorylate PPARα, which modifies (mainly increasing) its transcriptional activity [302]. However, the activation of p38, which AMPK may execute [303,304], induces the activation of PPARα in some cells while reducing it in others. Additionally, the phosphorylation of PPARα by glycogen synthase kinase, also regulated by AMPK [305], leads to the degradation of PPARα [302,306].
The activation of PPARα by AMPK has been shown in multiple experimental models. In myocytes, either 5-aminoimidazole-4-carboxamide ribonucleotide (AICAR), a synthetic activator of AMPK, or adiponectin, an insulin-sensitizing adipokine, increase FA oxidation gene expression via AMPK-dependent PPARα activation [307,308]. Therefore, the reduced serum levels of adiponectin in people with obesity and T2D may contribute to the observed impairment in PPARα activity [309]. Of note, in muscles, PPARα does not directly interact with AMPK [310]. Similarly, in the left atrial appendage of mixed-breed dogs, the AMPK/PPARα/VLCAD (very long-chain acyl-CoA dehydrogenase) pathway mediates the metformin-triggered reduction of lipid accumulation and increases the β-oxidation of FA [311]. In pancreatic β-cells, glucose represses PPARα gene expression via AMPK inactivation [312,313]. The mechanism of the direct interaction between AMPK and PPARα has been uncovered in hepatocytes. In this pathway, activated AMPKα subunits bind to and activate PPARα, which occurs independently of AMPK activity and is not associated with increased AMP concentration. Instead, the interaction is stimulated by increased MgATP levels. Surprisingly, treatment with AICAR decreases PPARα activity in rat hepatocytes, which is associated with translocation of the AMPKα2 isoform out of the nucleus and is independent of the kinase activity of AMPKα [314]. The contradictory information concerning the interaction between PPARα and the ligands of AMPK likely reflects tissue- and context-specific situations.
One publication has reported that AMPK inhibits PPARα and PPARγ activity [315]. In that study, the AMPK activators, AICAR, and metformin decreased basal and WY-14,643-stimulated PPARα activity in hepatoma cells. Compound C, which is an AMPK inhibitor, increased agonist-stimulated reporter activity and partially reversed the effect of the AMPK activators. The expression of either a constitutively active or dominant-negative AMPKα subunit inhibits basal and WY-14,643-stimulated PPARα activity. The authors postulated that the AMPK inhibition of PPARα and PPARγ may allow for short-term processes to increase energy generation before the cells devote resources to increasing their capacity for FA oxidation [315]. This contradictory report may indicate further that AMPK–PPAR regulation is cell-type specific.
An agonist of PPARα also can activate AMPK, suggesting that the activity regulation between AMPK and PPARα may be reciprocal. On the one hand, fenofibrate induces the phosphorylation and activation of AMPK via the induction of the small heterodimer partner (SHP; an orphan nuclear receptor) and its target genes [316]. On the other hand, WY-14,643 treatment increases the expression of AMPKα1 and α2 mRNA, leading to an increase in AMPKα subunit phosphorylation and its enzymatic activity [317]. Moreover, pterostilbene, a bioactive component of blueberries and grapes and an agonist of PPARα, activates AMPK, similarly to AICAR and metformin, and modulates several AMPK-dependent metabolic functions in the rat hepatoma cell line H4IIE [318]. The AMPK-mediated activation of PPARα reverses progressive fibrosis in steatohepatitis [316] by endothelial nitric oxide (NO) synthase (eNOS) phosphorylation in endothelial cells, which suppresses microvascular inflammation and apoptosis [319,320].

4.2. AMPK and PPARβ/δ

AMPK and PPARβ/δ, but not PPARα, interact directly and physically in muscle, leading to increased glucose oxidation via the upregulation of lactate dehydrogenase B, which is associated with improved exercise performance [310]. AICAR treatment increases endurance, and the combination of AICAR and GW0742 further potentiates it. The combination significantly increases all running parameters, which is a change that is accompanied by a significant shift to fat as the main energy source with a decline in carbohydrate use during the period near exhaustion [321]. Therefore, agonists of both AMPK and PPARβ/δ are recognized as exercise mimetics [322]. In line with these observations, the deletion of PPARβ/δ specifically in myocytes results in a reduced capacity to sustain running exercise [78].

4.3. AMPK and PPARγ

The activation of AMPK by PPARγ agonists has been documented in many cell lines [261,323,324,325,326], in various tissues ex vivo [327,328], and in nonhuman animals [329,330,331] and people [332]. In general, agonists of PPARγ act via AMPK to improve glucose and fat management. Troglitazone causes rapid increases in phosphorylated AMPK and acetyl-CoA carboxylase (ACC) within minutes after injection in rat skeletal muscle, liver, or adipose tissue. Consistently, the drug results in a two-fold increase in 2-deoxy-d-glucose uptake in skeletal muscle through AMPK activation [328]. In addition, rosiglitazone remarkably enhances AMPK-mediated glucose uptake and glycogen synthesis in muscle and adipose tissues [331]. In cardiac muscle, the impact of troglitazone on glucose uptake is triggered via AMPK and eNOS signaling [333]. Rosiglitazone increases the expression and circulating levels of adiponectin and enhances the expression of hepatic adiponectin receptors in mice, which correlates with the activation of the hepatic Sirt1/AMPK signaling system. This signaling enables rosiglitazone to attenuate alcoholic liver steatosis and nonalcoholic steatohepatitis [329,334]. Another TZD, pioglitazone, increases AMPK phosphorylation two-fold and decreases ACC activity and the concentration of malonyl-CoA by 50% in Wistar rat liver. Moreover, pre-treatment with pioglitazone prevents a 50% decrease in AMPK and ACC phosphorylation in the liver and adipose tissue, which can be triggered by a euglycemic–hyperinsulinemic clamp [330]. In endothelial cells, rosiglitazone reduces glucose-induced oxidative stress mediated by NAD(P)H oxidase hyperactivity induced by high glucose via AMPK activation. It also uses AMPK to stimulate eNOS activity to increase NO synthesis [324,325].
Several TZDs have been shown to reduce insulin resistance via AMPK activation [323,327,335,336]. AMPK-mediated pioglitazone signaling results in an increase in insulin-stimulated glucose disposal, enhanced expression of the genes encoding adiponectin receptors, and coding for factors connected with mitochondrial function and FA oxidation in the muscles of patients with diabetes [332]. Rosiglitazone promotes AMPK-mediated insulin secretion via the phosphorylation of the Kir6.2 subunit of the potassium ATP channel in β-cells [336]. The treatment of pancreatic β-cells with TZDs triggers the phosphorylation of AMPK and ACC and increases glucose-stimulated insulin secretion as well as the response of insulin secretion to the combined stimuli of glucose and palmitate [327]. This treatment also affects β-cell metabolism by reducing glucose oxidation, energy metabolism, and glycerolipid/FA cycling [323]. Thus, the role of TZDs in lowering serum insulin levels and in the protection of β-cells is mainly through AMPK [327].
In addition to mediating PPARγ metabolic functions, AMPK mediates the receptor’s anti-inflammatory activities. In bronchial epithelial cells, PPARγ plays a protective role in CSE-induced inflammation, as noted above (see the section on mTOR and PPARγ). CSE administration inactivates AMPK signaling, which is restored by PPARγ agonists. Consequently, the effects of PPARγ agonists on inflammation and also on autophagy can be abolished by AMPK inhibition [261], showing that AMPK is downstream of PPAR in this pathway. AMPK also mediates the anti-inflammatory effect of PPARγ in endothelial cells, in which the LPS-triggered downregulation of toll-like receptor 4 (TLR4) protein expression is inhibited by pioglitazone. LPS also reduces PPARγ expression, which can be partially restored by the knockdown of TLR4. Therefore, TLR4 and PPARγ affect each other via a negative feedback loop, and this interaction depends on the AMPK signaling pathway [326].
As discussed, agonists of PPAR exert physiological effects by modulating the activity of AMPK, which is an important cellular energy sensor. However, their action seems to be, at least in some instances, independent of the activation of the PPARs. In other words, these agonists can activate AMPK by phosphorylation independently of PPARγ or PPARα [316,317,325,328,337]. This idea is supported by a novel TZD, BLX-1002, with no PPAR affinity, which activates AMPK in β-cells and raises cytoplasmic Ca2+, thereby enhancing glucose-induced insulin secretion at a high glucose level [335]. Similarly, some agonists of PPAR likely exert some effects independently of PPAR, which is in cooperation with other cellular partners.
AMPK also has been reported to feed back to PPARγ. The expression of either a constitutively active or dominant-negative AMPKα inhibits basal and rosiglitazone-stimulated PPARγ activity. AICAR and metformin inhibit PPRE reporter activity, whereas AMPK inhibitor compound C increases basal and rosiglitazone-stimulated PPARγ activity [315].
In brief, there is a very tight interaction between AMPK and PPARs (Figure 4), which involves the factors participating in the metabolic, apoptotic, and anti-inflammatory response to CR.

5. Insulin Signaling

Increased glucose levels in serum after food intake promote insulin secretion from pancreatic β-cells, which in turn activates insulin receptors on the surface of target cells. The tyrosine kinase activity of the insulin receptor triggers a signaling cascade starting with the activation of insulin receptor substrates (IRS 1–4) followed by the phosphorylation of PI3K, which is responsible for metabolic actions including PDK1 and Akt activation. Akt occurs in three isoforms (1–3) with Akt2 being essential for glucose homeostasis, whereas Akt1 is important for growth and Akt3 is important for brain development [338]. The Akt-driven inhibition of AS160 phosphorylation induces GLUT4 to translocate to the cell membrane, which promotes glucose transport into the intracellular compartment. Akt also phosphorylates and deactivates glycogen synthase (GS) kinase 3 (GSK3), which stimulates GS and glycogen production. In parallel, it disrupts the CBP/Torc2/CREB complex and consequently inhibits gluconeogenesis. Moreover, Akt activates mTOR, which facilitates protein synthesis, whereas mTORC2 is a critical regulator of Akt [339]. Another Akt regulator, tumor suppressor PTEN, previously mentioned in the context of mTOR, prevents Akt activation and reduces mTOR activity. In line with the above, the inhibition of IGF-1/PI3K/Akt signaling participates in the anti-cancer and DNA-repair activity of CR [340,341,342]. Further, Akt activation leads to the inhibitory phosphorylation of FOXO1, resulting in its nuclear exclusion [343]. Therefore, Akt functions at the crossroads of several pathways responding to CR.
Among other pathways affected by insulin signaling, the most important include mitogen-activated protein kinase (MAPK), which regulates growth; SREBP-1, which promotes lipid and cholesterol synthesis; and the family of FoxO transcriptional regulators, which regulate metabolism and autophagy. In general, insulin signals an abundance of fuels and thus promotes storage and prevents the further production of energy molecules [344,345,346,347].
The beneficial effects of CR have been associated with changes in metabolism, modification of the activity of the insulin/IGF-1 pathway, reduction in fat mass, and increased stress resistance because of FoxO activation [348,349,350]. Insulin release and insulin action seem to play a major role in the control of aging. The modulation of longevity by insulin signaling is supported by the extended lifespan associated with mutations in the insulin/IRS/growth hormone (GH)/IGF-1/FOXO signaling pathways in humans, mice, C. elegans, and Drosophila [351,352,353,354,355,356,357]. Female, but not male, Igf1r+/− mice live on average 33% longer than their wild-type counterparts [355], and the fat-specific deletion of Igf1r results in an 18% increased longevity in both sexes [351]. Accordingly, GH receptor/binding protein knockout (GHR/BP-KO) mice are characterized by a markedly extended lifespan and show severely reduced plasma IGF-1 and insulin levels, as well as low glucose levels [358,359]. Transgenic Klotho mice, which also have an increased lifespan, are insulin resistant. These findings collectively suggest that aging can be delayed by reducing insulin signaling [360]. It has even been hypothesized that insulin resistance is a physiological protective mechanism against aging and age-related disorders [361].

5.1. Insulin Signaling and PPARα

The immense impact of PPARα on glucose homeostasis and insulin signaling is particularly well illustrated by pancreas malfunction and diabetes models. PPARα directly protects pancreatic islets and their function and improves the adaptive response of the pancreas to pathological conditions. PPARα activation during the fed-to-fasted transition affects the regulation of glucose-stimulated insulin release because of the critical role of FA in insulin secretion [362]. In this condition, the activation of PPARα in β-cells increases pancreatic FA oxidation and potentiates glucose-induced insulin secretion [363,364]. In contrast, PPARα activation can oppose insulin hypersecretion elicited by high-fat feeding [365], suggesting that this activation protects pancreatic islets from lipotoxicity. Similarly, in primary human pancreatic islets, PPARα agonist treatment prevents the FA-induced impairment of glucose-stimulated insulin secretion, apoptosis, and TG accumulation, indicating that PPARα mediates the adaptation of pancreatic β-cells to pathological conditions [366]. PPARα participates in a pathway mediating the effect of metformin on glucagon-like peptide-1 (GLP-1) receptor expression in pancreatic islets and on plasma levels of GLP-1 [367], improving glucose management. In addition, PPARα regulates hepatic glucose metabolism by upregulating glycerol-3-phosphate dehydrogenase, glycerol kinase, glycerol transport proteins [368], and pyruvate dehydrogenase kinase 4 during fasting [369], which leads to the promotion of gluconeogenesis over FA synthesis.
In in vivo models of insulin resistance and diabetes, PPARα activation reverses the pregnancy-related augmentation of glucose-stimulated insulin hypersecretion by increasing insulin sensitivity [370]. Similarly, in nondiabetic patients with hypertriglyceridemia and patients with latent diabetes, the improvement in glucose metabolism observed during short-term clofibrate administration may also result from increased insulin sensitivity. Fasting plasma glucose, oral glucose tolerance test results, and immunoreactive insulin in these patients are significantly decreased, which is accompanied by enhanced glucose use and decreased serum TGs and cholesterol [371]. Furthermore, clofibrate in patients with non-insulin-dependent diabetes decreases fasting plasma glucose and insulin levels, and insulin binding to erythrocytes is enhanced because of increased insulin receptor affinity without a change in receptor number [372]. Another study showed that clofibrate ameliorates glucose tolerance in this patient population without changing the number of insulin receptors and that this increased insulin sensitivity occurs via an unknown post-receptor mechanism [373].
Strikingly, chronic fenofibrate treatment completely prevents the spontaneous sequential hypertrophy and atrophy of pancreatic islets from obese diabetes-prone Otsuka Long Evans Tokushima Fatty (OLETF) rats, decreases body weight and visceral fat, and improves insulin action in skeletal muscle [374]. Along the same line of observations, fenofibrate treatment significantly reduces hyperinsulinemia and hyperglycemia in C57BL/6 mice with insulin resistance triggered by a high-fat diet and in a model of genetic insulin resistance (obese Zucker rats) [375]. Similarly, the treatment of db/db diabetic mice with PPARα agonists significantly reduces plasma insulin and insulin resistance, improves hyperglycemia, albuminuria, and kidney glomerular lesions, and causes a 50% reduction in FA oxidation, with a concomitant increase in glycolysis and glucose oxidation [376,377]. PPARα-deficient ob/ob mice with obesity-related insulin resistance develop pancreatic β-cell dysfunction characterized by reduced mean islet surface area and decreased insulin secretion in response to high glucose [366]. Similarly, PPARα KO mice develop marked age-dependent hyperglycemia [366], and after 24-h fasting, severe hypoglycemia accompanied by elevated plasma insulin concentrations [54,378]. However, PPARα KO mice are protected from high-fat diet-induced insulin resistance, which is most likely because of the development of increased adiposity [379]. Of note, PPARα gene variation in humans can affect the age of onset and progression of T2D in patients with impaired glucose tolerance [51,52].
In the liver, the insulin-stimulated activation of Akt induces the phosphorylation of NCoR1 on serine 1460, which selectively favors its interaction with PPARα. Phosphorylated NCoR1 inhibits the activity of PPARα, attenuating oxidative metabolism, whereas it derepresses liver X receptor α (LXRα), resulting in increased lipogenesis [380]. Glucose levels also affect PPARα activity. The exposure of islets or INS(832/13) β-cells for several days to supraphysiological glucose concentrations, which are detrimental to insulin secretion, leads to a 60–80% reduction in PPARα mRNA expression, DNA-binding activity, and target gene expression, which results in diminished FA oxidation and increased TG accumulation that are potentially associated with pancreatic lipotoxicity [381]. Moreover, insulin-activated MAPK and glucose-activated PKC stimulate PPARα transcriptional activity in HepG2 cells [382]. Strikingly, glucose itself can modulate PPARα activity because PPARα binds glucose and glucose metabolites with high affinity, prompting changes in its secondary structure [383]. Overall, based on the effects of PPARα on glucose homeostasis and its important regulatory role in the transition from feeding to fasting, PPARα might be involved in protecting against hypoglycemia during CR.

5.2. Insulin Signaling and PPARβ/δ

PPARβ/δ cross-reacts with insulin signaling at several points. At first, PPARβ/δ senses elevated glucose levels. Glucose overload leads to cPLA2 activation and the subsequent hydrolysis of arachidonic and linoleic acid and their peroxidation, producing endogenous ligands of PPARβ/δ [384]. In the mouse pancreas, PPARβ/δ represses insulin secretion and the β-cell mass [385]. In adipocytes, it prevents IL-6–dependent STAT3 activation by repressing ERK1/2 and STAT3–Hsp90 association. This effect is thought to prevent cytokine-induced insulin resistance in these cells [386]. Similarly, PPARβ/δ represses IL-6-induced STAT3 activation and suppressor of cytokine signaling-3 (SOCS-3) upregulation in human liver cells and thereby halts the development of insulin resistance [387]. In skeletal muscle cells, PPARβ/δ attenuates ER stress-associated inflammation and prevents insulin resistance in an AMPK-dependent manner [387,388]. Moreover, PPARβ/δ ameliorates hyperglycemia by increasing glucose flux through the pentose phosphate pathway, which enhances FA synthesis. Coupling PPARβ/δ-dependent increased hepatic carbohydrate catabolism and the promotion of β-oxidation in muscle allows PPARβ/δ to regulate metabolic homeostasis and enhance insulin action by complementary effects in distinct tissues [389]. In a primate model of metabolic syndrome, GW501516, an agonist of PPARβ/δ, dose-dependently lowers plasma insulin levels without side effects on glycemic control [390]. GW501516 treatment also markedly improves diabetes by decreasing blood glucose and insulin levels in ob/ob mice [391]. In addition, the treatment of healthy people who are moderately overweight with GW501516 results in a significant reduction in fasting plasma insulin [392], and the dual PPARα/δ agonist GFT505 (elafibranor) improves hepatic and peripheral insulin sensitivity in men with abdominal obesity [393].

5.3. Insulin Signaling and PPARγ

PPARγ is an established regulator of insulin sensitivity, making it an excellent drug target (Figure 5). TZDs form a class of PPARγ agonists that reverse insulin resistance in liver and peripheral tissues, reducing plasma glucose through specific PPARγ activation. Troglitazone was the first TZD approved for this use, but it was withdrawn from the market following reports of serious hepatotoxicity in some patients. TZDs not only improve insulin sensitivity but also preserve pancreatic β-cell function, thus reducing T2D incidence, as demonstrated in clinical trials of T2D prevention in high-risk people [394,395].
PPARγ exerts its insulin-sensitizing properties in several ways. First, it generates functional WAT, which is required for proper glucose homeostasis because lipodystrophy is associated with severe insulin resistance [396]. An early consequence of PPARγ activation that precedes decreased blood TG and glucose is the stimulation of TG production and a reduction in circulating free FA because of FA retention in fat rather than muscle and pancreas. Consequently, increased fat mass triggered by PPARγ activation results in improved glycemic control [397]. Accordingly, the level of insulin sensitization following PPARγ activation is correlated with the reduction in lipid accumulation in skeletal muscle [398]. Furthermore, in mice fed a high-cholesterol/fructose diet, the selective PPARγ agonist pioglitazone improves insulin sensitivity by affecting its signaling pathway, as measured by induction of IRS-2 expression and increased phosphorylation of Akt and GSK-3β [399]. In fact, PPARγ induces the expression of several proteins in the insulin-signaling pathway, including IRS-1 [400], IRS-2 [401], the p85 subunit of PI3K [402], and Cbl-associated protein (CAP) [403,404]. In 3T3-L1 adipocytes and diabetic rodents, PPARγ directly binds the promoter of the Cap gene. Increased CAP expression results in increased insulin-stimulated c-Cbl phosphorylation [403] and consequently in increased glucose uptake [405]. The activation of PPARγ in muscle cells and adipocytes increases the expression and translocation of GLUT1, GLUT2, and GLUT4 to the cell membrane, thus increasing glucose uptake and consequently reducing glucose plasma levels [406,407,408]. In parallel, PPARγ regulates the expression of genes responsible for glucose disposal [400,401,402,403,404].
An important contributor to the insulin-sensitizing effect of PPARγ ligands is the suppression of local and systemic cytokine production. TZD as treatment for patients with obesity and without diabetes reduces circulating levels of inflammatory cytokines and other pro-inflammatory markers, which are accompanied by improved insulin sensitivity [409]. Moreover, hepatic PPARγ reduces the expression of SOCS-3, which has been suggested to play a crucial role in linking inflammation and hepatic insulin resistance [399]. SOCS-3 promotes the ubiquitination and degradation of IRS-2 and thus modulates insulin signaling [410,411]. In vitro studies have confirmed that PPARγ agonists may also exert their antidiabetic activities by counteracting the negative effects of TNFα [412]. In addition, PPARγ elevates blood levels of adipocytokines, such as adiponectin, which are present at low concentrations in the plasma of patients with T2D. The increased adiponectin levels improve insulin sensitivity and free FA oxidation and reduce glucose production in the liver [413,414]. The signaling of PPARγ involves the previously mentioned executor of insulin signaling, FOXO. FOXO1 acts as a transcriptional repressor of Pparγ by binding to its promoter and may reduce PPARγ transcriptional activity through a transrepression mechanism involving direct protein–protein interaction between FOXO1 and PPARγ. This interaction seems to be a crucial part of the pathway responsible for insulin sensitivity in adipocytes [415,416,417]. Moreover, insulin signaling in the liver directly affects PPARγ, as Akt2 stimulates the expression and activity of PPARγ in hepatocytes, resulting in elevated aerobic glycolysis and lipogenesis [260].
As a result of this effect on regulatory pathways, TZDs improve insulin sensitivity, glucose tolerance, and the lipidemic profile in T2D as well as in obesity without diabetes [418]. Dominant-negative mutations in human PPARγ can lead to severe metabolic syndrome, insulin resistance, and diabetes at an unusually young age [419,420], and several point mutations in the PPARγ gene are associated with severe insulin resistance (with or without T2D) and familial partial lipodystrophy phenotypes [421,422,423,424,425]. Both partial and generalized lipodystrophies have consistently been associated with insulin resistance in animals and humans [426]. Therefore, it is likely that the dramatic reduction in limb and gluteal fat found in subjects with PPARγ mutations contributes to their insulin resistance. In addition, the residual adipose tissue in these individuals is dysfunctional, likely resulting in unregulated FA fluxes and impairing insulin action in skeletal muscle and liver [420]. Of interest, lipodystrophic, WAT-specific PPARγ KO mice show an increased expression of PPARγ in the liver, which promotes insulin sensitivity [427,428]. In this context, it is important to note that insulin sensitivity declines with age in humans and is accompanied by a lower expression of PPARγ in preadipocytes [429]. Hence, FA metabolism becomes altered with aging in preadipocytes, which correlates with increased susceptibility to lipotoxicity and impaired FA-induced adipogenesis. In line with these observations, PPARγ, PPARα, and RXR levels are all increased in the liver of GHR-KO long-lived animals [131]. Thus, the enhanced insulin sensitivity in GHR-KO mice may be the result of the increased hepatic activity of PPAR family members.
In addition to TDZs, several other PPARγ agonists influence insulin and glucose management. FMOC-L-Leucine (F-L-Leu) is a partial agonist that selectively activates some PPARγ pathways. F-L-Leu improves insulin sensitivity in normal, diet-induced glucose-intolerant mice and in diabetic db/db mice, yet it has a lower adipogenic activity [430]. Of interest, INT131 besylate, which is a potent non-TZD-selective PPARγ modulator, induces a dose-dependent reduction in fasting plasma glucose without evoking fluid retention or weight gain, which are both unwanted side effects often triggered by TZDs [431]. In addition, food-derived active compounds may contribute to the management of glucose levels. The plant polyphenols quercetin and kaempferol serve as weak partial agonists of PPARγ and increase insulin sensitivity and glucose uptake via PPARγ agonism [432,433]. Another compound, 13-oxo-9(Z),11(E),15(Z)-octadecatrienoic acid (13-oxo-OTA), a linolenic acid derivative in the extracts of tomato (Solanum lycopersicum), Mandarin orange (Citrus reticulata), and bitter gourd (Momordica charantia), modulates gene expression and the production of adiponectin through PPARγ in adipocytes [434]. The reduction of PPARγ activity by antagonists improves the metabolic profile in mice [435,436], and haplodeficient Pparγ+/− mice exhibit increased insulin sensitivity compared with their wild-type littermates [437,438]. These animals are characterized by reduced fat deposits and lower levels of TG accumulation and lipogenesis in WAT, skeletal muscle, and liver [439]. Similarly, genetic variants Pro(12)Ala (heterozygotes) and Ala(12)Ala (homozygotes) of PPARγ, which result in decreased receptor activity, are associated with leanness and improved insulin sensitivity [440,441,442]. A complex U-shaped curve has been proposed to characterize the relationship between PPARγ activity and insulin sensitivity [99].
Altogether, overwhelming evidence points to an important role for all three PPARs in insulin signaling and glucose level management, and to several compounds with similar potential, including some that block the endogenous ligand-induced activation of PPARγ for the treatment of the metabolic syndrome and T2D [436,443,444].

6. Sirtuins

As already mentioned, a CR-related decrease in energy levels leads to the activation of several signaling cascades. Decreased glucose intake reduces the flow of carbon through the glycolytic pathway and the regeneration of ATP from ADP, which eventually alters the NAD+:NADH ratio. This shift activates SIRTs, which serve as both energy sensors and transcriptional effectors by acting as NAD+-dependent HDACs. In addition to CR and fasting, exercise activates SIRTs [445,446], which are remarkably conserved and can even be found in archaebacteria [447]. Originally categorized as class III HDACs, SIRTs are involved in the proper functioning of nucleic acids including DNA repair, homologous recombination, and DNA deacetylation, and they promote transcriptional gene silencing [448,449].
The seven subtypes of SIRTs (SIRT1–7) in mice and humans vary in their cellular distribution and function. SIRT1–SIRT3, SIRT5, and SIRT6 catalyze deacetylation, whereas SIRT4 and SIRT6 have ADP-ribosylation capacity. In addition to histones, SIRT substrates include several transcriptional regulators, such as the nuclear factor kappa-light-chain enhancer of activated B cells (NF-κB), p53, FOXO, and PGC-1α, but also enzymes, including acetyl coenzyme A synthetase 2 (AceCS2), long-chain acyl-coenzyme A dehydrogenase (LCAD), HMGCS2, superoxide dismutase 2, and structural proteins, such as α-tubulin [450,451,452,453,454]. Therefore, SIRTs influence a wide range of cellular processes including circadian clocks, cell cycle, mitochondrial biogenesis, and energy homeostasis, and on the whole-body level regulate aging, apoptosis, inflammation, and stress resistance [455,456].
SIRT1 is the most thoroughly investigated mammalian SIRT and is closely involved in metabolism. Studies in S. cerevisiae have shown that an extra copy of the Sir2 gene, a yeast homolog of mammalian Sirt1, increases lifespan in a dose-dependent manner [457,458], and the deletion of this gene shortens lifespan [457]. In yeast and Drosophila, a lack of Sir2 and dSir2, respectively, prevents CR-associated life extension [459,460,461]. SIR2, a yeast analog of Sirt1, assists in DNA repair and regulates genes that change expression with age [462].
The most important metabolic regulator affected by SIRT1 is PGC-1α, which is activated by SIRT1-mediated deacetylation [463,464]. Deacetylated PGC-1α increases hepatic gluconeogenic activity [463], whereas in muscle and BAT, PGC-1α enhances mitochondrial activity. The activity of PGC-1α translates into increased exercise capacity and thermogenesis, leading to protection against the onset of obesity and associated metabolic dysfunction [465]. The deacetylation of PGC-1α by SIRT1 depends on cellular NAD+ levels, so the status of cellular energy affects PGC-1α activity, which adapts cellular energy production through mitochondrial biogenesis and function. Furthermore, among the SIRT1 substrates are factors that control cell proliferation and apoptosis, including the tumor suppressor protein p53 [466]. The overexpression of SIRT1 hinders p53 transcriptional activity and p53-dependent apoptosis triggered by DNA damage and oxidative stress, whereas the overexpression of dominant-negative SIRT1 can enhance cellular stress responses [466,467].
SIRTs also regulate the activity of the FOXO family of transcription factors [468,469], which affects cell differentiation, transformation, and metabolism as well as plays an important role in CR and longevity regulation [470,471,472]. SIRT1-mediated deacetylation of FOXO1 affects its shuttling between the nucleus and cytoplasm, influencing the expression of FOXO1 target genes and promoting gluconeogenesis and glucose release from hepatocytes [473]. The deacetylation of FOXO3a by SIRT1 increases its translocation from the cytoplasm to the nucleus [474] and its DNA-binding activity. In the nucleus, SIRT1 and FOXO3a form a complex that induces cell-cycle arrest and resistance to oxidative stress, also inhibiting the ability of FOXO3a to induce apoptosis [473]. SIRT1 directly suppresses the expression of uncoupling protein 2 (UCP2), leading to an improved coupling of mitochondrial respiration and ATP synthesis, which induces insulin secretion in β-cells [475]. Confirming the role of SIRT1 in the pancreas, SIRT1−/− mice are characterized by impaired insulin secretion in response to glucose compared with wild-type littermates [475,476]. Conversely, β-cell-specific SIRT1-overexpressing mice exhibit improved glucose tolerance and an enhanced glucose-stimulated insulin secretion [476]. In contrast, SIRT4 has an inhibitory effect on amino acid-stimulated insulin secretion. It represses the activity of glutamate dehydrogenase, reducing α-ketoglutarate production and ATP generation, which are known to activate insulin secretion in pancreatic β-cells [477]. To avoid amino acid-stimulated insulin secretion during CR, when amino acid turnover increases, CR decreases SIRT4 activity, which is opposite to the induction of SIRT1 activity during CR [477]. Considering that NAD+ controls the activities of both SIRT4 and SIRT1, their opposing effects on insulin secretion are surprising, and the full implications remain to be understood.
The role of other SIRT family members has been less investigated; thus, their function is less well known. SIRT2 is localized mainly in the cytoplasm, where it deacetylates tubulin filaments, HOXA10, and FOXO [478,479,480,481]. It takes part in multiple processes including cell cycle regulation [482], lifespan extension [457,483], and glucose and lipid metabolism [451,484]. SIRT3 plays an important role in mitochondria maintenance by acting as a deacetylase for a number of mitochondrial matrix proteins [485,486]. During a prolonged fast, SIRT3 activates FA breakdown by the deacetylation of LCAD [453] and stimulates the production of ketone bodies by activating HMGCS2 [452]. Of note, SIRT3 is genetically linked to lifespan in the elderly [487].
SIRT4 has ADP-ribosylation activity and in addition to blocking amino acid-induced insulin secretion [477], it regulates FA oxidation in hepatocytes and myocytes [488]. Both SIRT4 and SIRT5 show mitochondrial localization [477,489]. SIRT6 resides in the nucleus and is involved in genomic DNA stability and promotes the repair of DNA double-strand breaks [490]. SIRT6-deficient mice present a shortened lifespan and a degenerative aging-like phenotype [491]. In contrast, transgenic male mice overexpressing SIRT6 display lower serum levels of IGF-1, higher levels of IGF-1-binding protein, and modified phosphorylation patterns of different components of the IGF-1 signaling pathway, possibly contributing to about a 15% increase in lifespan when compared to wild-type animals [492].
SIRT1 and SIRT6 are both connected with CR-triggered extension of ovarian lifespan, which is mediated by the inhibition of the transition from primordial to developing follicles and by a delay in the growth phase of follicles to preserve the supply of germ cells [493]. SIRT7 is associated with nucleoli and is implicated in the activation of transcription by RNA polymerase I [494] as well as the repair of double-strand breaks by non-homologous end-joining [495]. SIRT7 knockout mice display features of premature aging [495]. SIRT1, SIRT6, and SIRT7 facilitate DNA repair, and this repair slows the aging process. During CR, except for SIRT4, the expression and activity of SIRTs are increased in many tissues, including adipose and brain [496,497,498], heart [499,500], and liver [501]. SIRT1 mediates a broad array of physiological effects of CR. The overexpression of SIRT in worms and flies increases their lifespan [460,461], and accordingly, mutants of SIRT do not show lifespan extension by CR [459,502]. Moreover, transgenic mice overexpressing SIRT1 show phenotypes similar to those of CR mice [503]. The previously mentioned role of yeast Sir2 in lifespan is particularly critical in the context of CR.
Resveratrol, a polyphenolic compound present in, for example, red grapes and wine, stimulates SIRT1 expression, resulting in extended lifespan and health span in treated animals [504]. SIRT1 activation by resveratrol mimics CR and delays aging in a wide range of organisms, from S. cerevisiae [505] to C. elegans to Drosophila [506] and mice [507]. Resveratrol is considered one of the mimetics not only of CR but also of exercise [504,508]. In mice, resveratrol inhibits gene expression profiles associated with muscle aging and age-related cardiac dysfunction [509]. The compound protects mice against diet-induced obesity and the associated insulin resistance through enhanced mitochondrial function mediated by PGC-1α [465].

6.1. SIRT and PPARα

During fasting, SIRT4 levels decrease in the liver and SIRT4-null mice display an increased expression of hepatic PPARα target genes associated with FA catabolism [510], indicating that PPARα is a negative downstream target of SIRT4. In contrast, the hepatocyte-specific deletion of SIRT1 impairs PPARα signaling and decreases FA β-oxidation, whereas the overexpression of SIRT1 induces the expression of PPARα targets (Figure 6). In fact, SIRT1 interacts with PPARα and is required to activate PGC-1α by deacetylation. Of note, SIRT1-deacetylated PGC-1α can function as a coactivator in PPARα complexes controlling the expression of several metabolic genes. Therefore, SIRT1 activates PPARα to promote FA oxidation in the liver [511]. Similarly, in the heart, PPARα and SIRT1 modulate FA metabolism [512]. Both PPARα and SIRT1 are upregulated by pressure overload in the heart. The haploinsufficiency of either PPARα or SIRT1 reduces pressure overload-induced cardiac hypertrophy and failure, whereas the simultaneous induction of PPARα and SIRT1 aggravates cardiac dysfunction. PPARα and SIRT1 jointly suppress genes involved in mitochondrial functions that are controlled by the estrogen-related receptors (ERRs). PPARα binds and recruits SIRT1 to the ERR response element. In doing so, it represses ERR target genes in an RXR-independent manner. Suppression of the ERR transcriptional pathway by PPARα/SIRT1 also is a physiological response to fasting [513,514,515].

6.2. SIRT and PPARβ/δ

PPARβ/δ markedly increases the transcription [516] and protein levels of SIRT1 [517], whereas PPARα and PPARγ do not stimulate SIRT1 expression [516]. Moreover, PPARα and PPARβ/δ promote osteogenic differentiation in an SIRT1-dependent manner [518,519], and PPARγ prevents it [520]. Based on the PPARβ/δ–SIRT1 interaction, a reasonable inference is that during starvation, increased levels of lipolysis-derived free FAs activate PPARβ/δ. This activation leads to enhanced SIRT1 expression, promoting the deacetylation of factors involved in mitochondrial beta-oxidation and cell survival [516]. The regulation of SIRT1 and PPARβ/δ activity operates bidirectionally. First, in human HaCaT keratinocytes, GW501516 modulates inflammation by acting via AMPK and SIRT1 to reduce TNFα-induced IL-8 mRNA levels and NF-κB DNA-binding activity [517]. Second, the upregulation of SIRT1 by PPARβ/δ attenuates premature senescence in angiotensin (Ang) II-treated human coronary artery endothelial cells. Resveratrol can mimic the action of PPARβ/δ on Ang II-induced premature senescence and reactive oxygen species (ROS) generation [521].

6.3. SIRT1 and PPARγ

The interaction between PPARγ and SIRT1 is twofold (Figure 6). PPARγ inhibits SIRT1 expression by binding to the Sirt1 promoter, and PPARγ also directly interacts with and inhibits SIRT1 activity, forming a negative feedback loop [522]. Pioglitazone prevents NF-κB activation through a reduction in p65 acetylation via the AMPK-SIRT1/p300 pathway [523], whereas SIRT1 represses PPARγ actively via docking with two of its cofactors, NcoR and SMRT [524]. Conversely, the treatment of 3T3-L1 adipocytes with resveratrol represses the expression of PPARγ target genes as well as of PPARγ itself. Furthermore, this treatment increases targeting of the PPARγ protein to the ubiquitin–proteasome system for degradation [525]. Hence, SIRT1 acts as a corepressor of PPARγ-mediated transcription. From a functional point of view, the repression of PPARγ by SIRT1 counters adipogenesis, and the upregulation of SIRT1 triggers lipolysis and the release of fat from differentiated adipocytes [22,524]. Following food withdrawal, SIRT1 promotes fat mobilization by repressing PPARγ, which reduces the expression of genes mediating fat storage [524]. In line with these observations, SIRT1+/− mice show a compromised mobilization of FAs from adipose tissue during fasting [524].

7. Major Outcomes of CR

7.1. Oxidative Stress Reduction

ROS are generated as a by-product of cellular respiration, contributing to the accumulation of oxidative damage and the formation of a range of oxidation products of different macromolecules including lipids, proteins, and nucleic acids [526]. A small amount of ROS is normally beneficial because it plays an important role in cellular processes such as cell cycle progression, the regulation of signaling pathways in response to intra- and extracellular stimuli, and inflammation [527]. However, high uncontrolled levels of ROS are detrimental.
During oxidative stress, the sustained production of ROS and reactive nitrogen species leads to a perturbed equilibrium between pro-oxidants and antioxidants. Consequently, macromolecules, organelles, and cells are altered, and if much damage accumulates, necrotic or apoptotic cell death occurs. The “free radical theory” of aging [528] proposes that the generation of oxidative stress is a major factor contributing to the onset of the aging process and age-related diseases. Therefore, the mammalian lifespan is reduced in relation to the mitochondrial production of oxidizing free radicals [527]. CR likely exerts its diverse benefits through reducing ROS levels and suppressing age-related oxidative stress while supporting the antioxidant defense system [529,530,531]. CR diminishes the impact of ROS through three processes: reduction of oxygen free-radical generation by slowing metabolism, the acceleration of ROS neutralization, and stimulation of the repair of ROS-damaged molecules [532,533,534,535,536].
The oxidative stress-related role of PPARs is first suggested by their name: they were first identified as receptors stimulating peroxisome proliferation. Peroxisomes have oxidative functions that involve use of molecular oxygen and that yield hydrogen peroxide (H2O2). The name of these organelles comes from their hydrogen peroxide-generating and scavenging activities. In addition to the conversion of ROS, peroxisomes play a key role in metabolism, catabolizing very long-chain FAs, branched-chain FAs, bile acid intermediates (in the liver), D-amino acids, and polyamines. The induction of oxidative stress is associated with the downregulation of PPARs, which also occurs during aging [140,537,538]. The reduced expression of PPARα in aging [137,539] has been attributed to increased oxidative stress, and CR has been suggested to prevent this decrease through antioxidative action [140].
PPARα-deficient mice present increased oxidative stress at an earlier age than aged-matched wild-type controls [137]. In fact, the administration PPARα agonists to aged mice restores the cellular redox balance, documented by reduced tissue lipid peroxidation, reduced spontaneous inflammatory cytokine production, and the elimination of constitutively active NF-κB [137]. WY-14643 and fenofibrate protect mice from acetaminophen-induced hepatotoxicity by upregulating UCP-2, which is a PPARα target gene that reduces the generation of mitochondrial ROS [540]. In a gentamicin-induced model of ROS production, different types of PPARα and PPARγ agonists (fenofibrate, pioglitazone, tesaglitazar) provide protection from toxicity. These ligands prevent oxidative stress by increasing the expression of genes controlling ROS production and detoxification (SOD1, glutathione peroxidase 1 (GPx1), CAT, UCP-2), which will restore the ratio of reduced to oxidized glutathione and prevent apoptosis [541].
PPARγ directly modulates the expression of several antioxidant and pro-oxidant enzymes as well as oxidative stress-related proteins. It transcriptionally regulates mouse, rat, and human catalase, which is a major antioxidant enzyme converting H2O2 to O2 and H2O [542,543]. Similarly, it directly regulates the expression of manganese superoxide dismutase (MnSOD), which performs the dismutation of O2 to O2 and H2O. Conversely, heart-specific PPARγ knockout mice show downregulated levels of MnSOD in cardiac muscle with a consequent increase in O2 levels, suggesting that PPARγ protects cardiomyocytes from oxidative damage [544]. In human skeletal muscle cells, the TZD-mediated activation of PPARγ induces GPx3 and protects against oxidative stress [545] because GPx reduces H2O2 to H2O and O2 and scavenges for oxidized lipids. PPARγ also represses the expression of inducible NO synthase (iNOS) and stimulates eNOS [546,547,548,549,550]. These enzymes produce NO from arginine, which forms highly reactive peroxynitrite when it reacts with O2. In mice with an endothelial-specific knockout of PPARγ, aortic segments release less NO than those from controls, and this reduced expression correlates with an increase in oxidative stress parameters [548].
Cyclooxygenase-2 (COX-2) is an inducible form of cyclooxygenase that contributes to the metabolism of arachidonic acid-forming prostaglandin H2 [551,552], which requires the presence of free radicals and may produce O2, contributing to oxidative stress. PPARγ regulates COX-2 expression, but both induction [553,554] and reduction [555,556] in PPARγ expression have been reported, leaving the issue for further investigation. In rats, the activation of PPARγ by oral intake of rosiglitazone upregulates UCP-2 [557], which protects against oxidative stress by preventing O2 accumulation in the mitochondria and facilitating the export of mitochondrial ROS to the cytosol [558]. Moreover, a major target gene of PPARγ, CD36, may act as a scavenger receptor that mediates the recognition and internalization of oxidized lipids [559,560,561]. Finally, PPARγ also has been shown to protect cardiomyocytes and glial cells from oxidative stress-induced apoptosis by increasing Bcl-2 [562,563].
In addition to direct transcriptional regulation, PPARγ can modulate the inflammatory and oxidative status by acting on transcription factors such as NF-κB [547,550,564,565]. NF-κB action is usually pro-inflammatory and pro-oxidant, inducing the expression of genes encoding the inflammatory cytokines sIL-1β, IL-6, and TNFα, as well as the pro-inflammatory enzymes COX-2 and iNOS, but it may also regulate the expression of superoxide dismutases and other anti-inflammatory genes [552,566,567,568,569,570]. PPARγ reduces NF-κB activities in various ways: (1) by transrepressing NF-κB activation through forming a repressor complex in the promoter of NF-κB-target genes; (2) by directly binding with NF-κB [547,550,564]; or (3) by catalase-mediated H2O2 reduction, which activates NF-κB [542,543,571]. Conversely, NF-κB negatively regulates PPARγ transcriptional activity via a mechanism that requires the presence of HDAC3 [572,573].
Of note, PPARγ interacts with a major regulator of the antioxidative response, the nuclear factor erythroid 2-related factor 2 (NRF2). NRF2 is a redox-sensitive transcription regulator that plays a vital role in cryoprotection against oxidative and electrophilic stress as well as in inflammation suppression [574]. NRF2 targets multiple genes, including NADPH-generating enzymes [575], glutathione S-transferases [576], CD36 [560,577], and HO-1 [578,579] and it stimulates the production of defense proteins during oxidative stress. NRF2 also induces PPARγ expression by binding the upstream promoter region of the nuclear receptor [580,581]. Conversely, PPREs have been identified on the NRF2 gene promoter [576,581], confirming a positive feedback loop between PPARγ and NRF2. Therefore, the ability of PPARs to extinguish oxidative stress overlaps with CR effects.

7.2. Mitochondrial Function

One of the several theories tightly connected with the effects of ROS is the “mitochondrial theory of aging”, which proposes that mitochondria are the critical component in the aging process. In fact, mitochondrial DNA damage and dysfunction increase with aging and are associated with a vast number of pathologies. Defective mitochondria determine the turnover not only of the organelles themselves but also whole cells, resulting in the acceleration of aging [527,582,583]. Aging has been linked to a reduced capacity for oxidative phosphorylation in the muscle and heart, most likely because of a decline in mitochondrial content and/or function [584,585,586]. Accordingly, young individuals have higher respiratory function compared to the elderly [587,588,589]. Disturbed mitochondrial electron transfer increases the likelihood of electron leakage and ROS production. Consequently, components of the electron transport chain and mitochondrial DNA become damaged, leading to further increases in intracellular ROS levels and a decline in mitochondrial function. Since mitochondrial DNA is spatially close to the source of ROS production, it is thought to be particularly vulnerable to ROS-mediated lesions [528,590].
An interesting feature of CR, one associated with ROS and changes in metabolism, is mitochondria biogenesis, which is relatively high in various tissues such as in the brain, heart, liver, and particularly the BAT of mice [498,591]. It is associated with activation of the master regulator of mitochondrial biogenesis, PGC-1α [428,592,593]. PGC-1α is expressed at a high level in BAT, heart, skeletal muscle, brain, and kidney, whereas its expression is low in the liver and very low in WAT [594]. Various physiological stimuli highly induce PGC-1α in different organs. It is increased in BAT by cold exposure and in skeletal muscle by exercise and decreased ATP level, whereas in the liver, it is mostly affected by CR [595]. When ectopically expressed in fat or muscle cells, PGC-1α strongly increases mitochondrial biogenesis and oxidative metabolism, which correlates with an increase in mitochondrial DNA and the expression of multiple mitochondrial genes [595,596]. To prevent a mitochondrial biogenesis-associated increase in ROS levels, PGC-1α also induces expression of the antioxidant genes GPx1 and MnSOD [597]. One hypothesis regarding the beneficial outcomes of CR proposes is that CR preserves mitochondrial function by maintaining protein and DNA integrity through decreasing mitochondrial oxidant emission and increasing endogenous antioxidant activity [598,599]. Its impact on mitochondria biogenesis remains a matter of discussion [600,601].
In addition to affecting mitochondria biogenesis, PGC-1α also influences metabolism. It mediates a fasting-induced increase in FA metabolism and the downregulation of pyruvate dehydrogenase, which is part of the mitochondrial pyruvate dehydrogenase complex that catalyzes the reaction representing pyruvate entry into the tricarboxylic acid cycle. In PGC-1α knockout mice, pyruvate dehydrogenase fails to adapt to CR, and the ability of the mice to endure prolonged starvation is decreased [602]. PGC-1α knockout mice also show a reduced content of mitochondrial electron transport chain proteins in skeletal muscle [603,604]. The activity of PGC-1α is directly regulated by the energy sensors SIRT1 and AMPK [276,463]. Functionally, the transcriptional activity of PGC-1α relies on its interactions with transcriptional factors for controlling FA metabolism. Of note, all three PPAR isotypes are subject to transcriptional coactivation by PGC-1α and are major executors of PGC-1α-induced regulation [72,594,605,606].
Evidence has accumulated for an important role of PPARs in maintaining healthy mitochondria. Agonists of PPARα and PPARγ modulate mitochondrial fusion and fission in neurons, leading to a better response to oxidative stress and neuron protection [607]. The abnormal expression of PPARα is linked to an altered mitochondrial structure and metabolic function, with an increase in number of cristae, and myocardial damage and fibrosis in PPARα knockout mice [608]. Through its key role in FA β-oxidation, PPARα is inevitably associated with mitochondrial function [35,609]. The activation of PPARα rescues mitochondrial depletion and failure to oxidize FA in the liver-specific class 3 PI3K-deficient mice. In this model, PPARα stimulates mitochondrial biogenesis and lipid oxidation by the inhibition of HDAC3 [610].
In addition, fenofibrate ameliorates insulin resistance accompanied by an improved mitochondrial oxidative capacity in pediatric burn patients [611]. Fenofibrate and gemfibrozil also reduce mitochondrial membrane potential depolarization, resulting in apoptosis inhibition in lymphoblast cells in Batten disease [612]. Pretreatment of rats with gemfibrozil prior to global cerebral I/R results in neuroprotection by modulating mitochondrial biogenesis and apoptosis [613]. WY-14,643 and fenofibrate protect mice from acetaminophen-induced hepatotoxicity by upregulating UCP-2, which is a PPARα target gene that reduces the generation of mitochondrial ROS [540]. However, fibrates may also trigger mitochondrial dysfunction because they inhibit the activity of mitochondrial respiratory chain complex I in rat skeletal muscles [614]. Moreover, gemfibrozil and WY-14,643 alter mitochondrial energy production by promoting mitochondrial permeability transition, as documented by membrane depolarization and calcium-induced swelling, which inhibits the oxidative phosphorylation and ATP synthesis in the rat liver [615]. Finally, chronic treatment with WY-14,643 impairs myocardial contractile function while decreasing mitochondrial respiratory function and increasing mitochondrial uncoupling in rats [616].
PPARβ/δ has been reported to be essential for the exercise-induced increase in the number of muscle mitochondria [617]. In high-fat–fed C57BL/6 mice, the administration of GW501516, a PPARβ/δ agonist, increases the metabolic rate, reduces fatty liver, decreases lipid accumulation, and increases mitochondrial biogenesis in the muscle [391]. In C2C12 muscle cells, GW501516 induces the mRNA expression of UCP-1, UCP-2, and UCP-3 [618], which are responsible for uncoupling mitochondrial respiration [619]. Accordingly, muscle-specific PPARβ/δ KO mice show a reduction in the expression of genes connected with energy uncoupling, mitochondrial electron transport chain, FA uptake, and catabolism. They develop obesity and diabetes with aging and present a mild defect in glucose metabolism when challenged with a high-fat diet [72,78]. The overexpression of either wild-type PPARβ/δ or constitutively active VP16-PPARβ/δ in muscle results in fiber-type switching and augmented capacity for mitochondrial pyruvate oxidation, which is accompanied by an induction of mitochondria numbers [80,310,620]. This phenotype resembles that of the muscle-specific overexpression of PGC-1α [621] and accords with PGC-1α acting as a co-activator of PPARβ/δ to control mitochondrial biogenesis and muscle fiber-type plasticity [72,78,80]. Furthermore, the expression in BAT of a constitutively active PPARβ/δ results in the increased expression of genes involved in FA oxidation and lipolysis and in energy uncoupling in the mitochondria. However, some studies have shown that the impact of PPARβ/δ on energy metabolism in muscles does not necessarily involve de novo mitochondrial biogenesis [622,623].
PPARγ and TZDs also have roles in mitochondrial structure and function. Transgenic mice overexpressing PPARγ2 have a significantly increased expression of mitochondrial UCP-1, elevated levels of PGC-1α, and reduced mitochondrial ATP concentrations in their subcutaneous fat [624]. The overexpression of cardiac PPARγ results in the production of mitochondria with a distorted architecture of the inner matrix and disrupted cristae [625]. Rosiglitazone treatment of ob/ob mice leads to mitochondrial remodeling, enhanced oxygen consumption, and increased energy expenditure in WAT [626]. Moreover, in adipose tissues of patients with diabetes, agonists of PPARγ increase the relative amounts of mitochondria and mitochondrial DNA copy number and stimulate the expression of factors involved in mitochondrial biogenesis, respiratory complexes I–IV, and FA oxidation [627,628]. A similar response has been described in differentiated 3T3-L1 and C3H/10T1/2 adipocytes treated with rosiglitazone, which showed increased mitochondrial biogenesis, oxygen consumption, and mitochondrial citrate synthase activity [629]. Finally, rosiglitazone also protects T lymphocytes from apoptosis by preventing the loss of mitochondrial membrane potential [630]. The mitochondria-related impact of TZDs has been identified as a basis for their neuroprotective effect [607,631,632,633,634]. Of interest, TZDs can also exert PPARγ-independent effects on mitochondrial respiration, leading to changes in glycolytic metabolism and fuel substrate specificity [635,636]. Taken together, the evidence strongly suggests that these three PPARs contribute to the maintenance of mitochondria in a tissue-specific manner.

7.3. Reduction of Inflammation

The “inflammation hypothesis of aging” posits a molecular mechanism of aging based on inflammation. Inflammation is a complex defense reaction to insult and both physiological and nonphysiological stress, which is induced by agents such as chemicals, drugs, or microbial entities. Inflammation responses are activated by well-coordinated, sequential events controlled by humoral and cellular reactions. Elevated tissue levels of TNFα, IL-1, and IL-6, among other pro-inflammatory mediators, have been observed in experimental animal models of inflammation. With aging, inflammatory responses may be overactive or even cause damage, resulting in pathological conditions [14].
During aging, a shift occurs in the ratio of naive to memory T cells, with associated changes in the cytokine profile in favor of inflammatory cytokines such as TNFα, IL-1, IL-6, INFγ, and transforming growth factor β [637,638,639,640]. There is also a progressively higher dysregulation of immune cells and pro-inflammatory responses. Macrophages from old mice produce more prostaglandin E2 than those from young mice because of higher COX-2 activity [641]. One major causative factor in tissue inflammation is the uncontrolled overproduction ROS/reactive nitrogen species. The transcriptional regulator NF-κB is an inflammatory reaction factor of major importance that is extremely sensitive to oxidants [552,566,567,568,569,570]. Enhanced IL-6 production by activated NF-κB has been implicated in many pathophysiological dysfunctions of aging ranging, from Alzheimer’s disease to atherosclerosis [642]. CR exhibits a broad and effective anti-inflammatory effect. It blunts age-triggered increases in COX-2 levels and activity through the modulation of NF-κB and IκB, in which COX-2-derived ROS generation decreases. In addition, the production of iNOS, IL-β, IL-6, TNFα, and prostanoids such as thromboxane A2 (TXA2), prostacyclin 2, and prostaglandin E2 is suppressed [14,531]. The prevention of the age-related decline triggered by CR correlates with dampening the reduction of PPAR expression and activity seen during aging. Therefore, under CR conditions, higher PPAR expression may play a role in the suppression of the age-induced increase in inflammation [140]. PPARs are implicated in inflammation at the transcriptional level by interfering with pro-inflammatory mediators such as NF-κB, STAT-1, and activating protein-1, leading to the downregulation of the gene targets of these factors [643,644,645,646]. In this way, PPARα and PPARγ inhibit the expression of inflammatory genes, such as COX-2, iNOS, cytokines, metalloproteases, and acute-phase proteins [549,644]. Inflammatory eicosanoids serve as ligands for PPARs, and the levels of these signaling molecules, including prostaglandins and leukotrienes, increase with age [647].
Each of the three PPAR isotypes exhibits a set of individual anti-inflammatory properties [58]. The anti-inflammatory activity of PPARα is in a great part the result of its interaction with NF-κB. The deletion of PPARα results in a premature and enhanced age-dependent increase in oxidative stress and NF-κB activity [137]. Similarly, aged PPARα KO mice display higher oxidative stress at a younger age and an exacerbated inflammatory response to LPS stimulation [137,648]. In contrast, the administration of PPARα agonists to aged wild-type mice restores the cellular redox balance, as attested by the elimination of constitutively active NF-κB and a loss in spontaneous inflammatory cytokine production [137,649]. The interaction of NF-κB and PPARα is intriguing, because high doses of the PPARα ligands activate NF-κB, whereas low or therapeutic doses of the ligands cause decreased NF-κB activation accompanied by reduced IL-6 production and lipid peroxidation [137]. During CR, PPARα is required, at least partially, to mediate the downregulation of acute-phase genes (C4bp, C9, Mbl1, Orm1, Saa4) that are responsive to inflammatory cytokines [141].
PPARβ/δ also shows anti-inflammatory properties and can suppress, in a ligand-independent manner, inflammatory bowel disease by the dampening of inflammatory signaling [650]. In cultured cardiomyocytes, the PPARβ/δ agonist GW0742 inhibits LPS-induced TNFα secretion, whereas the absence of PPARβ/δ exaggerates LPS-induced TNFα production [651]. The intracerebroventricular administration of high-affinity PPARβ/δ agonists significantly decreases the infarct volume at 24 h of reperfusion after cerebral ischemia in rats, again underscoring the anti-inflammatory and neuroprotective properties of PPARβ/δ [652]. Lastly, the activation of PPARβ/δ by GW0742 protects skeletal muscle against metabolic disorders caused by chronic exposure to a high concentration of sugars by affecting the insulin and inflammatory cascades, including reversal of the diet-induced activation of NF-κB and the expression of both iNOS and intercellular adhesion molecule 1 [653].
PPARγ is undeniably one of the most important and best documented anti-inflammatory factors. PPARγ agonists mitigate inflammatory bowel disease symptoms, reduce inflammation, and are effective in multiple models of ulcerative colitis as well as in Crohn’s disease [654,655,656,657,658,659,660,661,662,663,664,665,666,667]. Functionally, the binding of PPARγ to a DNA-bound repressor complex in macrophages blocks the expression of inflammatory genes by preventing the 19S proteasome-mediated degradation of the repressor complex [668]. Accordingly, the ligands of PPARγ inhibit macrophage activation, stimulate macrophage differentiation into non-inflammatory type M2, and suppress the production of inflammatory cytokines in macrophages and dendritic cells, resulting in increased susceptibility to infection in PPARγ-deletion mouse models [549,644,669,670,671,672]. Of interest, the Pro(12)Ala substitution in PPARγ (rs1801282 C>G), which results in a modest decrease in its transcriptional activity and adipogenic potential, mediates anti-inflammatory benefits. The Pro(12)Ala substitution is associated with a 10-year delay in the onset of multiple sclerosis [673] and with a decreased risk for T2D [442]. Males carrying the 12Ala allele and having coronary artery disease show less widespread atherosclerosis and are protected against 10-year vascular morbidity and mortality [674]. Furthermore, another PPARγ polymorphism (rs 1801282 C>G, rs3856806 C>T) is associated with colorectal cancer risk [675,676]. Mice deficient in colonic PPARγ display more acute infectious colitis [663] and are resistant to conjugated linoleic acid therapy for colitis [677]. The molecular mechanism behind the anti-inflammatory activities of PPARγ includes inhibition of the expression of inflammatory genes encoding cytokines, metalloproteases, and acute-phase proteins, and the regulation of multiple signaling pathways, such as those related to p53 [678], Bcl2 [89], c-Myc, [679], Cox-2 [91,680,681,682], iNOS [683], and Apc/β-catenin [684,685]. Most importantly, PPARγ inhibits NF-κB and NF-κB-driven transcription [89,682]. PPARγ may reduce NF-κB activities in various ways (see the section on PPARs and oxidative stress). Therefore, it is likely that PPARs mediate, at least in part, the anti-inflammatory properties of CR.

7.4. Metabolic Adaptation

The shortage of energy during CR leads to a sequence of metabolic changes. Following the depletion of dietary glucose, glycogen is mobilized as an energy supply, and upon prolonged CR, hepatic metabolism shifts to gluconeogenesis to prevent hypoglycemia. Some enzymes connected with hepatic glycolysis, gluconeogenesis, and glycogen metabolism are under the control of PPARα. During fasting, PPARα stimulates glucose import, glycolysis, and glycogenolysis [686,687,688,689]. Accordingly, the expression of several genes involved in gluconeogenesis and glycogen metabolism is reduced in PPARα KO mice [368], and these animals show impaired gluconeogenesis regulation and marked hypoglycemia during fasting [54,55].
Upon prolonged energy restriction, carbohydrate depletion triggers a shift to fat recruitment and ketone body production. This switch between energy sources relies on PPARs. Exercise-elicited glycogen depletion activates PPARβ/δ in rat muscle [690]. We speculate that a similar regulation takes place in fasting-related carbohydrate shortage, which would contribute to PPARβ/δ-driven FA oxidation in muscles. Similarly, the upregulation of the expression of PPARα by CR has been suggested to act as a direct stimulus to enhance FA β-oxidation in the heart [139].
PPARs also regulate the expression of many genes involved in insulin signaling, glucose uptake, lipid metabolism, and ketogenesis, which are affected by CR. Particularly, the metabolism of lipids and ketone bodies in the liver employs PPARα to regulate the expression of most of the rate-limiting enzymes of β-oxidation including ACOX1 (acyl-CoA Oxidase 1), EHHADH (enoyl-CoA hydratase and 3-hydroxyacyl CoA dehydrogenase), carnitine palmitoyltransferases I and II, MCAD (medium-chain acyl-CoA dehydrogenase), LCAD (long chain acyl-CoA dehydrogenase), VLCAD (very long chain acyl-CoA dehydrogenase), and fibroblast growth factor 21, and of ketogenesis, such as HMG-CoA synthase [141,691,692,693,694,695,696,697,698,699,700].
During fasting, PPARα promotes cellular FA uptake and β-oxidation and mediates the adaptation to FA catabolism, lipogenesis, and ketone body synthesis in response to energy depletion [53,54,55]. Consequently, fasting-induced hepatic responses, including elevated FA oxidation and ketogenesis, are all impaired in PPARα-null mice, resulting in hypoketogenesis and liver steatosis [53,54,55]. Similarly, in aged mammals, including humans, the capacity for FA oxidation and hepatic ketogenesis decreases, resulting in reduced energy metabolism as well as increased dyslipidemia [223,224,225]. In healthy men, the L162V substitution of PPARα is associated with higher fasting total cholesterol, low-density lipoprotein cholesterol, and apoB, but not with postprandial parameters [50]. Both PPARα and PPARβ/δ are essential regulators of FA oxidation, and their roles in this process overlap. Of importance, the two PPARs show a distinct primary area of activity, with PPARα activating FA oxidation mainly in the liver and BAT, whereas PPARβ/δ controls lipid metabolism in the pancreas, heart, and skeletal muscle. PPARα does not seem to be involved in the metabolic adaptation of the liver to every-other-day fasting [701].
Reduced energy intake accompanied by increased mobilization of the fat reservoir results in weight loss. The release of energy from WAT involves the inhibition of expression of the lipid-storing PPARγ. Collectively, all three PPARs act as metabolic sensors and play essential roles in lipid and FA metabolism. However, PPARγ is more responsible for fat storage and PPARα and PPARβ/δ are more responsible for energy expenditure. Likely for that reason, a high-fat diet increases the expression of PPARγ in the liver, whereas intermittent fasting decreases it [702]. Genetic variation in the Pparγ gene and its target gene Acsl5 determine the capacity for weight loss under CR [703], and six Pparγ single nucleotide polymorphisms are significantly associated with weight reduction in response to CR [704]. Most of the data concerning Pparγ polymorphisms focus on the Pro(12)Ala substitution. Based on a report of a population of children in Mexico, Pro(12)Pro homozygosity is the more represented, followed by Pro(12)Ala heterozygosity, and more rarely Ala(12)Ala homozygosity (73.9%:24.5%:1.6%) [705]. The (12)Ala PPARγ protein shows a decreased binding affinity for PPRE and consequently is a weaker stimulator of target gene expression [441,706]. The presence of (12)Ala PPARγ and resistance to CR-induced weight loss were associated in a comparison of women with obesity losing the most weight to those losing the least after 6 weeks of a 900 kcal/day CR [703]. Moreover, PPARγ polymorphism is associated with changes in body mass index (BMI) in response to the total fat intake [707,708], FA composition in the diet [709], and plasma TG response to ω3 FA [710]. This polymorphism also influences weight regain following CR, with women homozygous for Ala(12)Ala gaining more weight compared to women with Pro(12)Pro homozygosity [711], likely indicating lesser metabolic flexibility for Ala(12)Ala individuals.
Long-term CR leads to energy-saving adaptations that can result in a lower resting metabolic rate and decreased body temperature [712,713,714], which is possibly because of reduced thyroid hormone levels. Bezafibrate, a panagonist for all three PPARs, has been reported to induce WAT beiging and thus shows potential for regulating body temperature [715]. Similarly, the activation of PPARα in WAT and BAT results in increased UCP-1 expression and consequently elevated energy dissipation and higher body temperature [716]. Crosstalk between thyroid hormone receptors and PPARs appears to be important for regulating thermogenesis and metabolism [717,718]. In summary, the involvement of PPARs in the metabolic feeding-to-fasting adaptation places these receptors at the center of the proper body response to CR.

7.5. Physical Exercise

Exercise, similar to CR, yields multiple beneficial effects. Research outcomes point toward the effectiveness of regular moderate exercise in preventing and delaying several metabolic disorders, chronic diseases, and premature death. Increased physical activity reduces mortality risk from many age-related diseases, including cardiovascular disease, stroke, T2D, certain cancers, hypertension, obesity, depression, and osteoporosis [719,720,721,722,723]. However, in rodents, exercise improves the mean lifespan without increasing maximum longevity [724,725]. Similarly, high physical activity fails to extend maximum lifespan in humans [726]. Compared to exercise, long-term CR in humans improves several biomarkers related to aging [727,728]. Accordingly, exercise has been deemed as unable to fully mimic the beneficial hormonal and/or metabolic changes associated with CR [729]. Therefore, despite a mutual influence with CR on similar molecular pathways and providing multiple advantages, physical activity has been recognized as yielding inferior benefits compared to CR.
Physical activity results in the release of stored energy and elevated levels of FAs, increasing the availability of ligands for PPARs [730]. In rat liver, exercise increases PPARα expression and transcription, and PPARα may mediate the impact of exercise on plasma glucose, TG, and cholesterol [731]. Similarly, PPARα may be involved in the protective effects of exercise against myocardial infarction and for cardiac function by changing the expression of metabolic and inflammatory response regulators and by reducing myocardial apoptosis [732]. However, among PPARs, PPARβ/δ is particularly well-known for its impact on physical performance. First, PPARβ/δ controls muscle development and the adaptive response to exercise, and its overexpression results in a switch to type I muscle fiber [620]. Second, exercise can increase PPARβ/δ expression in skeletal muscle, and this activation is essential for increasing the number of exercise-induced muscle mitochondria [617]. As previously mentioned (see the AMPK and PPARβ/δ section), AMPK and PPARβ/δ are exercise mimetics [322], and their stimulation significantly increases running parameters and promotes muscle remodeling [321,322]. During exercise, the depletion of carbohydrates in skeletal muscle limits endurance. PPARβ/δ represses glycolytic genes in muscle to slow glucose catabolism, reducing the use of carbohydrates at the period near exhaustion. In parallel, PPARβ/δ induces a shift to FAs as the main energy source and thus extends the possible exercise time [623,733]. Consequently, transgenic mice overexpressing muscle-specific PPARβ/δ show enhanced exercise performance, but PPARα-overexpressing animals do not [80,310,620].
Exercise is associated with increased PPARγ DNA-binding activity and expression of its target genes in leukocytes [734]. Similarly, in skeletal muscle and subcutaneous WAT, PPARγ and PGC-1α mRNA expression increases in response to physical training, and these expression changes are proposed to mediate the effect of exercise on insulin sensitivity [735]. Furthermore, the beneficial outcome of low-intensity exercise on plasma lipid levels is exerted via PPARγ [734], and PPARγ1 promotes exercise-induced lipoprotein lipase expression [736]. In addition, the Pro(12)Ala substitution in PPARγ polymorphism is associated with reduced glucose and insulin levels as well as body weight loss in response to exercise [737,738,739,740]. Therefore, as in the case of CR, PPARs play an active role in upstream molecular signaling and beneficial outcomes of exercise.

7.6. Hunger

Food withdrawal or limitation inevitably results in a hunger sensation. The physiology of hunger involves a complex network of sensors, hormones, and neuronal signaling. Hunger signaling relies on PPARs, particularly PPARα. For example, allele “A” in PPARA rs4253747 (a single nucleotide polymorphism in an intron region) in young men of Han Chinese ancestry is significantly associated with an increased risk for appetite loss at high altitude. In contrast, the “AC” haplotype of PPARA rs7292407-rs6520015 in the same cohort had a protective role for high altitude appetite loss [741]. If PPARα is implicated in appetite control, some of its natural ligands should affect hunger sensation. Oleoylethanolamide (OEA), an endogenous ligand of PPARα, is one such compound. This endocannabinoid is produced by enterocytes in response to fat consumption [742], and bile acids modulate its biosynthesis, which requires sympathetic innervation [743,744]. The administration of OEA has an anorectic effect by acting peripherally, reducing meal size or prolonging eating latency, leading to body weight loss [742,745,746,747]. PPARα activation in the proximal small intestine mediates this effect [742,748,749]. The intraperitoneal administration of OEA acutely decreases energy expenditure, as well as ambulatory and spontaneous locomotor activity [750]. Via PPARα, OEA stimulates lipolysis and decreases the neutral lipid content in hepatocytes, as well as serum cholesterol and TG levels, and thereby regulates lipid metabolism [751]. OEA engages afferent sensory fibers of the vagal nerve in the intestine, leading to an increased expression of proto-oncogene c-fos in the nucleus solitary tract and the paraventricular nucleus of the brainstem and hypothalamus, respectively [751], which promotes oxytocin secretion and satiety [752]. Since enterocytes in the small intestine are the first cells to respond to dietary fat intake by increasing OEA production, OEA has been suggested to serve as a gut-derived satiety factor [742].
The other PPARs may also indirectly affect appetite. As a master regulator of adipogenesis, PPARγ plays a crucial role in regulating food intake because WAT secretes a number of endocrine and paracrine satiety mediators, including leptin, adiponectin, and resistin [753]. Inflammation, which is well established as being under PPAR influence, reduces appetite [754]. Particularly, NF-kB, which interacts with all PPARs, has been implicated in appetite suppression [755,756].

7.7. Longevity and Aging

According to the “rate-of-living” theory, lifespan differs between species and it correlates with energy metabolic rate and, in general, with body size [757,758]. Thereupon, the level of CR needed to prolong life by a certain percentage varies from species to species. In fact, CR increases lifespan to different extents with a stronger impact on short-living animals, such as C. elegans (up to 150%) [759], Drosophila [760], or rodents (up to 50%) [761] compared to long-lived species including lemurs [762] and rhesus monkeys [6,763]. Therefore, it is not possible to extrapolate the results obtained for one species to others. Importantly, the level of restriction (10–50%) applied in various studies strongly impacts the outcome [11,12]. Although it is proven that CR causes beneficial metabolic modifications in women and men, the exact amount of calorie intake that is necessary to reach maximum longevity and sustain good health is not known, but it is likely different from person to person. It is also obvious that excessive CR leads to malnutrition with adverse health effects.
Both genetic and environmental factors control the progression of aging. Aging is associated with immunosenescence, increased oxidative stress, decreased hormonal secretion, changes in metabolic rate, mitochondrial function, insulin resistance, and dysregulated lipid metabolism [764,765,766]. The preservation of insulin sensitivity by reducing levels of blood glucose and insulin without compromising glucose fuel may prevent age-related metabolic phenotypes [353]. Glucose metabolism maintenance is a key feature of the anti-aging actions of CR [3]. In fact, genes connected with the insulin/IGF-1 signaling pathway have been proposed as longevity candidate markers [356,357,767]. Paradoxically, impaired insulin signaling through the insulin receptor or its substrates increases rather than decreases lifespan in a number of mouse models [351,356,357]. With regard to insulin sensitivity, the undeniable role of PPARα and PPARβ/δ has already been described above. Worth highlighting here in the context of PPARγ is the role of WAT in proper glucose homeostasis, as attested by the association of lipodystrophy with severe insulin resistance [396].
Aging is accompanied by increased body fat, particularly in the visceral areas as well as in liver and skeletal muscle, which initiates the development of age-associated hypertension, atherosclerosis, inflammatory diseases, dyslipidemia, and T2D [764,765,766]. In fact, WAT, which has also been directly associated with lifespan, is a central physiological component of energy metabolism [351,396], and its development and function rely on PPARγ [106,768,769]. In addition to the direct regulation of adipogenesis from fat-produced PPARγ, intestinal PPARγ regulates body adiposity by sympathetic nervous system signaling in mice on CR [101]. Therefore, through the coordination of glucose homeostasis and adipogenesis, PPARγ might affect longevity [770].
A low expression of PPARγ reduces the lifespan in both lipodystrophic PPARγ1/2-hypomorphic and PPARγ2-deficient mice [771]. We suggest that the reduced fat mass observed during CR is not what results in longevity, but rather that the key factor is proper adipose tissue functionality, including insulin-sensitizing effects. Correspondingly, the human genetic variant genotypes Pro(12)Ala and Ala(12)Ala of PPARγ are associated with leanness, improved insulin sensitivity, and increased lifespan in both humans and mice [440,441,442]. Moreover, gene network analysis has identified PPARγ as one of the “longevity genes” in mouse WAT [771]. However, reports are contradictory concerning the expression of PPARγ in WAT in response to CR. One group found that both 8 weeks of 50% CR and intermittent fasting downregulated PPARγ mRNA and protein expression in the adipose tissue of obese rats [132]. A similar effect was observed in the subcutaneous adipose tissue of obese humans following 10-week CR [772]. In contrast, in intermittently fasting rats, PPARγ2 mRNA levels were approximately two-fold higher than in control or CR animals [773], resulting in a not well-understood effect of restrictive diets on PPARγ2 expression in WAT.
Most of the prominent features of aging are related to PPAR activity, mTOR activity, oxidative stress, inflammation, and metabolism. Moreover, changes in PPAR expression and activity often occur in aging and are reversed by CR [140,224,539,774]. PPAR activity also may be affected indirectly through age-dependent decreases in RXR, the heterodimerization partner of PPARs [775,776,777]. The impact of PPARs can be particularly well observed in mutant models of longevity, such as the dwarf mice. Snell dwarf, Ames dwarf, and “Little” mice display low levels of GH or a defect in GH signaling because of a mutation (generating the GHR-KO strain). All of these dwarf mice are characterized by having a markedly longer lifespan than their wild-type counterparts and share a number of beneficial phenotypic characteristics with rodents on CR diets. Similar to CR animals, dwarf mice are protected from spontaneous and chemically induced cancer, age-dependent declines in immune function, collagen cross-linking, decreased levels of insulin and IGF-1, and increased insulin sensitivity [428]. The increased expression of PPARα and constitutive activation of some of its target genes have been detected in the liver of the dwarf mice [131,778]. The increased expression of genes involved in β- and ω-oxidation of FAs (Acox1, Cyp4a10, Cyp4a14) in the liver of these mice suggests increased FA oxidation, which could be beneficial for insulin sensitivity. PPARα levels are decreased in the muscle of GHR-KO animals, and PPARβ/δ protein levels are downregulated in the liver and skeletal muscle, which mimics the expression profile in wild-type CR mice [136]. The protein levels of PPARγ are elevated in the liver but downregulated in the skeletal muscle of the GHR-KO animals [136]. Furthermore, the overexpression of fibroblast growth factor 21, previously mentioned as a PPARα target gene, extends the lifespan in mice without affecting AMPK or mTOR but blunting GH/IGF-1 signaling in the liver [779].
In contrast to GHR-KO mice, animals overexpressing the bovine GH gene have a markedly shorter lifespan in comparison to their wild-type counterparts. The hepatic expression of PPARα is decreased in these mice, as is the expression of genes involved in FA activation, peroxisomal and mitochondrial β-oxidation, and the production of ketone bodies. Consequently, bovine GH mice exhibit a reduced ability to produce ketone bodies in the fed and fasted states [780]. The antagonistic relationship between PPARs and GH is demonstrated by the fact that the surgical removal of the pituitary gland (hypophysectomization) of rats enhances the expression of PPAR-inducible proteins, which can be reversed by GH infusion [781]. Moreover, STAT5b, a GH-inducible transcription factor, inhibits the ability of PPARα to activate PPARα-dependent reporter gene transcription [782,783], and PPARα downregulates STAT5b [784]. Consequently, PPARs may control lifespan at the level of glucose and lipid metabolism and hormonal regulation.

7.8. Microbiota Composition

Microbiota composition changes upon CR have been repeatedly observed [138,785,786,787,788]. CR increases the abundance of bacteria that positively correlate with lifespan, mainly Firmicutes including Lactobacillus, Allobaculum, Papillibacter, or Lachnospiraceae. In parallel, CR reduces the occurrence of bacteria that negatively correlate with lifespan, such as Clostridiales, Riminococcaceae, Alistipes, or Rikenella [787,788,789,790,791]. The exact effect of microbiota on the outcome of CR is not fully known, but the microbiota mediates some of the beneficial outcomes of CR, including reduced body weight and decreased blood leptin and insulin levels [791]. We could speculate that there is an effect on metabolism, body fat storage, and the endocrine system of microbiota-driven changes in the production of signaling molecules and ligands for nuclear receptors, including PPARs [101]. Indeed, the interaction of PPARs with the microbiota has been well documented. The expression of PPARα and its target genes coding for rate-limiting enzymes of ketogenesis depends on stimulation by commensal gut microbiota [691,698,792]. Using germ-free mice, we have shown that the microbiota not only promotes harvesting energy from the food but is also generating signals, which regulate the hepatic clock genes and their effector genes such as the PPARs, and several PPARα target genes [793]. Of note, PPARα also mediates signals received from the microbiota via TLRs and contributes to the circadian expression of genes in the intestine and intestinal corticosterone production [794]. Thus, PPARα forwards information from the gastrointestinal flora, which affects host physiology. Furthermore, PPARα has been identified as an important factor in the inflammatory response of the intestine to commensal microbiota [795]. It regulates the expression of IL-22, the antimicrobial peptides Reg3β and Reg3γ, and calprotectin [795]. In the context of restrictive diets, the microbiota mediates the stimulatory effect of intermittent fasting on beige fat development [796]. Similarly, the deletion of PPARα triggers the upregulation of UCP1 expression in WAT [796]. PPARβ/δ, which is constitutively expressed in the intestine at a high level [45] and takes part in the differentiation of intestinal cells, is indirectly involved in the secretion of antimicrobial peptides [74,650,797]. Therefore, it influences gut microbiota composition.
The expression and activity of PPARγ are induced in the gut by multiple nutrients [114], bacterial metabolites, and bacterial by-products [115,116,117,118], and the presence of specific bacterial strains [117,119,120] stimulates PPARγ expression and activity. However, CR has been shown to limit the production of butyrate [787], which is one of the short chain fatty acids (SCFAs) that is known to activate PPARγ [116,117]. Moreover, the microbiota affects the liver circadian rhythm by modulating the activity of PPARγ expressed in the liver [798]. Of note, PPARγ is responsible for the selective killing of bacteria associated with inflammatory bowel disease by stimulating the expression of β-defensins and the maintenance of innate antimicrobial immunity in the colon [799].
Thus, there are reciprocal interactions between PPARs and gut microbiota in which PPARs can be activated by bacteria and regulate the intestinal microbiota composition [800]. The additional impact of CR on the expression PPARs points to a potential role for PPARs in the response of microbiota to CR.

8. Conclusions

After remarkable achievements in medical research that have translated into a notable increase in life expectancy, the current focus is more on increasing disease-free years. With the potential to alleviate numerous health conditions while extending the lifespan, CR remains a relevant candidate in health-related research. Therefore, the current recommendation on energy intake should be revised, particularly for individuals with a high risk of developing metabolic, inflammatory, or neurodegenerative diseases.
The beneficial impact of several restrictive approaches including multiple models of intermittent fasting [801,802,803,804] and fasting-mimicking diet [805,806,807] has been proven. These diets imply temporal restriction without long-term energy deprivation or prolonged hunger making them less restrictive than CR, easier to apply in everyday life, more flexible for various lifestyles, and therefore more plausible for a wide population. The majority of available studies concerning intermittent fasting are observational studies focused on weight loss, cardiovascular risk, and inflammation. However, knowledge of the molecular mechanism behind the observed effects is still limited [808,809]. More studies comparing CR and other specific restrictive diets, in terms of molecular pathways and health outcomes, are needed to identify which restrictive approach is more efficient. Eventually, the aim is to encourage the use of such diets as a means to prevent diseases. Currently, the large variety of intermittent fasting and CR protocols increases the complexity of this task. Interestingly, the involvement of PPARs in the impact of the intermittent fasting and fasting-mimicking diet has not yet been verified.
Despite the lack of a fully revealed network of pathways underlying CR, the current review gives a detailed overview of how CR exerts its effects on the whole organism and which of the many outcomes are mediated by PPARs. During CR, the energy and nutrient sensor pathways involving mTOR, AMPK, insulin signaling, and SIRT are tightly interconnected, resulting in reduced oxidative stress and inflammation, increased autophagy, improved mitochondrial function, and the regulation of metabolism, hunger, and microbiota composition. As discussed, multiple connections have been found between PPARs and regulatory responses to CR and fasting. PPARs take part in managing the initial shortage of energy, modulating the main signaling pathways, adjusting metabolism to prolonged low energy intake, and mediating direct and long-term outcomes. Each of these roles could contribute to improvements related to common diseases such as diabetes and cancer, and even neurodegenerative diseases. Therefore, PPAR agonist treatments together with CR has great potential for synergistic effects to be explored in future experimental and clinical studies.

Author Contributions

K.D. wrote the manuscript and prepared the figures, A.G. assisted with research and figure preparation, H.G. edited the manuscript and supported the writing with expert advice, J.K. supported the writing with expert advice, and W.W. counseled on content, corrected the manuscript, and supported the writing with expert advice. All authors have read and agreed to the published version of the manuscript.

Funding

Open access funding was provided by University of Vienna.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Osborne, T.B.; Mendel, L.B.; Ferry, E.L. The Effect of Retardation of Growth Upon the Breeding Period and Duration of Life of Rats. Science 1917, 45, 294–295. [Google Scholar] [CrossRef] [Green Version]
  2. McCay, C.M.; Crowell, M.F.; Maynard, L.A. The effect of retarded growth upon the length of life span and upon the ultimate body size. 1935. Nutrition 1989, 5, 155–171, discussion 172. [Google Scholar]
  3. Masoro, E.J. Overview of caloric restriction and ageing. Mech. Ageing Dev. 2005, 126, 913–922. [Google Scholar] [CrossRef]
  4. Speakman, J.R.; Mitchell, S.E. Caloric restriction. Mol. Asp. Med. 2011, 32, 159–221. [Google Scholar] [CrossRef]
  5. Weindruch, R.; Walford, R.L.; Fligiel, S.; Guthrie, D. The retardation of aging in mice by dietary restriction: Longevity, cancer, immunity and lifetime energy intake. J. Nutr. 1986, 116, 641–654. [Google Scholar] [CrossRef] [Green Version]
  6. Colman, R.J.; Anderson, R.M.; Johnson, S.C.; Kastman, E.K.; Kosmatka, K.J.; Beasley, T.M.; Allison, D.B.; Cruzen, C.; Simmons, H.A.; Kemnitz, J.W.; et al. Caloric restriction delays disease onset and mortality in rhesus monkeys. Science 2009, 325, 201–204. [Google Scholar] [CrossRef] [Green Version]
  7. Anderson, R.M.; Shanmuganayagam, D.; Weindruch, R. Caloric restriction and aging: Studies in mice and monkeys. Toxicol. Pathol. 2009, 37, 47–51. [Google Scholar] [CrossRef] [Green Version]
  8. Fontana, L.; Meyer, T.E.; Klein, S.; Holloszy, J.O. Long-term calorie restriction is highly effective in reducing the risk for atherosclerosis in humans. Proc. Natl. Acad. Sci. USA 2004, 101, 6659–6663. [Google Scholar] [CrossRef] [Green Version]
  9. Fontana, L.; Klein, S. Aging, adiposity, and calorie restriction. JAMA 2007, 297, 986–994. [Google Scholar] [CrossRef]
  10. Cerqueira, F.M.; Kowaltowski, A.J. Commonly adopted caloric restriction protocols often involve malnutrition. Ageing Res. Rev. 2010, 9, 424–430. [Google Scholar] [CrossRef]
  11. Dogan, S.; Ray, A.; Cleary, M.P. The influence of different calorie restriction protocols on serum pro-inflammatory cytokines, adipokines and IGF-I levels in female C57BL6 mice: Short term and long term diet effects. Meta Gene 2017, 12, 22–32. [Google Scholar] [CrossRef]
  12. Dogan, S.; Rogozina, O.P.; Lokshin, A.E.; Grande, J.P.; Cleary, M.P. Effects of chronic vs. intermittent calorie restriction on mammary tumor incidence and serum adiponectin and leptin levels in MMTV-TGF-alpha mice at different ages. Oncol. Lett. 2010, 1, 167–176. [Google Scholar] [CrossRef] [Green Version]
  13. Phelan, J.P.; Rose, M.R. Why dietary restriction substantially increases longevity in animal models but won’t in humans. Ageing Res. Rev. 2005, 4, 339–350. [Google Scholar] [CrossRef]
  14. Chung, H.Y.; Kim, H.J.; Kim, J.W.; Yu, B.P. The inflammation hypothesis of aging: Molecular modulation by calorie restriction. Ann. N. Y. Acad. Sci. 2001, 928, 327–335. [Google Scholar]
  15. Nuclear Receptors Nomenclature, C. A unified nomenclature system for the nuclear receptor superfamily. Cell 1999, 97, 161–163. [Google Scholar]
  16. Feige, J.N.; Gelman, L.; Michalik, L.; Desvergne, B.; Wahli, W. From molecular action to physiological outputs: Peroxisome proliferator-activated receptors are nuclear receptors at the crossroads of key cellular functions. Prog. Lipid Res. 2006, 45, 120–159. [Google Scholar] [CrossRef]
  17. Michalik, L.; Auwerx, J.; Berger, J.P.; Chatterjee, V.K.; Glass, C.K.; Gonzalez, F.J.; Grimaldi, P.A.; Kadowaki, T.; Lazar, M.A.; O’Rahilly, S.; et al. International Union of Pharmacology. LXI. Peroxisome proliferator-activated receptors. Pharmacol. Rev. 2006, 58, 726–741. [Google Scholar] [CrossRef]
  18. Viswakarma, N.; Jia, Y.; Bai, L.; Vluggens, A.; Borensztajn, J.; Xu, J.; Reddy, J.K. Coactivators in PPAR-Regulated Gene Expression. PPAR Res. 2010, 2010. [Google Scholar] [CrossRef] [Green Version]
  19. Horlein, A.J.; Naar, A.M.; Heinzel, T.; Torchia, J.; Gloss, B.; Kurokawa, R.; Ryan, A.; Kamei, Y.; Soderstrom, M.; Glass, C.K.; et al. Ligand-independent repression by the thyroid hormone receptor mediated by a nuclear receptor co-repressor. Nature 1995, 377, 397–404. [Google Scholar] [CrossRef]
  20. Chen, J.D.; Evans, R.M. A transcriptional co-repressor that interacts with nuclear hormone receptors. Nature 1995, 377, 454–457. [Google Scholar] [CrossRef]
  21. Kang, Z.; Fan, R. PPARalpha and NCOR/SMRT corepressor network in liver metabolic regulation. FASEB J. 2020. [Google Scholar] [CrossRef]
  22. Yu, C.; Markan, K.; Temple, K.A.; Deplewski, D.; Brady, M.J.; Cohen, R.N. The nuclear receptor corepressors NCoR and SMRT decrease peroxisome proliferator-activated receptor gamma transcriptional activity and repress 3T3-L1 adipogenesis. J. Biol. Chem. 2005, 280, 13600–13605. [Google Scholar] [CrossRef] [Green Version]
  23. Krogsdam, A.M.; Nielsen, C.A.; Neve, S.; Holst, D.; Helledie, T.; Thomsen, B.; Bendixen, C.; Mandrup, S.; Kristiansen, K. Nuclear receptor corepressor-dependent repression of peroxisome-proliferator-activated receptor delta-mediated transactivation. Biochem. J. 2002, 363, 157–165. [Google Scholar] [CrossRef]
  24. Neschen, S.; Morino, K.; Dong, J.; Wang-Fischer, Y.; Cline, G.W.; Romanelli, A.J.; Rossbacher, J.C.; Moore, I.K.; Regittnig, W.; Munoz, D.S.; et al. n-3 Fatty acids preserve insulin sensitivity in vivo in a peroxisome proliferator-activated receptor-alpha-dependent manner. Diabetes 2007, 56, 1034–1041. [Google Scholar] [CrossRef] [Green Version]
  25. Krey, G.; Braissant, O.; L’Horset, F.; Kalkhoven, E.; Perroud, M.; Parker, M.G.; Wahli, W. Fatty acids, eicosanoids, and hypolipidemic agents identified as ligands of peroxisome proliferator-activated receptors by coactivator-dependent receptor ligand assay. Mol. Endocrinol. 1997, 11, 779–791. [Google Scholar] [CrossRef]
  26. Plutzky, J. Peroxisome proliferator-activated receptors in vascular biology and atherosclerosis: Emerging insights for evolving paradigms. Curr. Atheroscler. Rep. 2000, 2, 327–335. [Google Scholar]
  27. Kliewer, S.A.; Sundseth, S.S.; Jones, S.A.; Brown, P.J.; Wisely, G.B.; Koble, C.S.; Devchand, P.; Wahli, W.; Willson, T.M.; Lenhard, J.M.; et al. Fatty acids and eicosanoids regulate gene expression through direct interactions with peroxisome proliferator-activated receptors alpha and gamma. Proc. Natl. Acad. Sci. USA 1997, 94, 4318–4323. [Google Scholar] [CrossRef] [Green Version]
  28. Moller, D.E.; Berger, J.P. Role of PPARs in the regulation of obesity-related insulin sensitivity and inflammation. Int. J. Obes. Relat. Metab. Disord. 2003, 27 (Suppl. S3), S17–S21. [Google Scholar] [CrossRef] [Green Version]
  29. Tan, C.K.; Zhuang, Y.; Wahli, W. Synthetic and natural Peroxisome Proliferator-Activated Receptor (PPAR) agonists as candidates for the therapy of the metabolic syndrome. Expert Opin. Ther. Targets 2017, 21, 333–348. [Google Scholar] [CrossRef]
  30. Cheng, H.S.; Tan, W.R.; Low, Z.S.; Marvalim, C.; Lee, J.Y.H.; Tan, N.S. Exploration and Development of PPAR Modulators in Health and Disease: An Update of Clinical Evidence. Int. J. Mol. Sci. 2019, 20, 5055. [Google Scholar] [CrossRef] [Green Version]
  31. Jain, M.R.; Giri, S.R.; Bhoi, B.; Trivedi, C.; Rath, A.; Rathod, R.; Ranvir, R.; Kadam, S.; Patel, H.; Swain, P.; et al. Dual PPARalpha/gamma agonist saroglitazar improves liver histopathology and biochemistry in experimental NASH models. Liver Int. 2018, 38, 1084–1094. [Google Scholar] [CrossRef] [Green Version]
  32. Staels, B.; Rubenstrunk, A.; Noel, B.; Rigou, G.; Delataille, P.; Millatt, L.J.; Baron, M.; Lucas, A.; Tailleux, A.; Hum, D.W.; et al. Hepatoprotective effects of the dual peroxisome proliferator-activated receptor alpha/delta agonist, GFT505, in rodent models of nonalcoholic fatty liver disease/nonalcoholic steatohepatitis. Hepatology 2013, 58, 1941–1952. [Google Scholar] [CrossRef]
  33. Tenenbaum, A.; Fisman, E.Z. Balanced pan-PPAR activator bezafibrate in combination with statin: Comprehensive lipids control and diabetes prevention? Cardiovasc. Diabetol. 2012, 11, 140. [Google Scholar] [CrossRef] [Green Version]
  34. Boubia, B.; Poupardin, O.; Barth, M.; Binet, J.; Peralba, P.; Mounier, L.; Jacquier, E.; Gauthier, E.; Lepais, V.; Chatar, M.; et al. Design, Synthesis, and Evaluation of a Novel Series of Indole Sulfonamide Peroxisome Proliferator Activated Receptor (PPAR) alpha/gamma/delta Triple Activators: Discovery of Lanifibranor, a New Antifibrotic Clinical Candidate. J. Med. Chem. 2018, 61, 2246–2265. [Google Scholar] [CrossRef]
  35. Desvergne, B.; Wahli, W. Peroxisome proliferator-activated receptors: Nuclear control of metabolism. Endocr. Rev. 1999, 20, 649–688. [Google Scholar] [CrossRef] [Green Version]
  36. Green, S.; Wahli, W. Peroxisome proliferator-activated receptors: Finding the orphan a home. Mol. Cell Endocrinol. 1994, 100, 149–153. [Google Scholar]
  37. Pawlak, M.; Lefebvre, P.; Staels, B. Molecular mechanism of PPARalpha action and its impact on lipid metabolism, inflammation and fibrosis in non-alcoholic fatty liver disease. J. Hepatol. 2015, 62, 720–733. [Google Scholar] [CrossRef] [Green Version]
  38. Lazennec, G.; Canaple, L.; Saugy, D.; Wahli, W. Activation of peroxisome proliferator-activated receptors (PPARs) by their ligands and protein kinase A activators. Mol. Endocrinol. 2000, 14, 1962–1975. [Google Scholar] [CrossRef]
  39. Wadosky, K.M.; Willis, M.S. The story so far: Post-translational regulation of peroxisome proliferator-activated receptors by ubiquitination and SUMOylation. Am. J. Physiol. Heart Circ. Physiol. 2012, 302, H515–H526. [Google Scholar] [CrossRef] [Green Version]
  40. Floyd, Z.E.; Stephens, J.M. Controlling a master switch of adipocyte development and insulin sensitivity: Covalent modifications of PPARgamma. Biochim. Biophys. Acta 2012, 1822, 1090–1095. [Google Scholar] [CrossRef] [Green Version]
  41. Hu, E.; Kim, J.B.; Sarraf, P.; Spiegelman, B.M. Inhibition of adipogenesis through MAP kinase-mediated phosphorylation of PPARgamma. Science 1996, 274, 2100–2103. [Google Scholar] [CrossRef] [Green Version]
  42. Diradourian, C.; Girard, J.; Pegorier, J.P. Phosphorylation of PPARs: From molecular characterization to physiological relevance. Biochimie 2005, 87, 33–38. [Google Scholar] [CrossRef]
  43. Leuenberger, N.; Pradervand, S.; Wahli, W. Sumoylated PPARalpha mediates sex-specific gene repression and protects the liver from estrogen-induced toxicity in mice. J. Clin. Investig. 2009, 119, 3138–3148. [Google Scholar] [CrossRef] [Green Version]
  44. 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]
  45. Braissant, O.; Foufelle, F.; Scotto, C.; Dauca, M.; Wahli, W. Differential expression of peroxisome proliferator-activated receptors (PPARs): Tissue distribution of PPAR-alpha, -beta, and -gamma in the adult rat. Endocrinology 1996, 137, 354–366. [Google Scholar] [CrossRef] [Green Version]
  46. Robitaille, J.; Brouillette, C.; Houde, A.; Lemieux, S.; Perusse, L.; Tchernof, A.; Gaudet, D.; Vohl, M.C. Association between the PPARalpha-L162V polymorphism and components of the metabolic syndrome. J. Hum. Genet. 2004, 49, 482–489. [Google Scholar] [CrossRef] [Green Version]
  47. Lacquemant, C.; Lepretre, F.; Pineda Torra, I.; Manraj, M.; Charpentier, G.; Ruiz, J.; Staels, B.; Froguel, P. Mutation screening of the PPARalpha gene in type 2 diabetes associated with coronary heart disease. Diabetes Metab. 2000, 26, 393–401. [Google Scholar]
  48. Vohl, M.C.; Lepage, P.; Gaudet, D.; Brewer, C.G.; Betard, C.; Perron, P.; Houde, G.; Cellier, C.; Faith, J.M.; Despres, J.P.; et al. Molecular scanning of the human PPARa gene: Association of the L162v mutation with hyperapobetalipoproteinemia. J. Lipid Res. 2000, 41, 945–952. [Google Scholar]
  49. Flavell, D.M.; Pineda Torra, I.; Jamshidi, Y.; Evans, D.; Diamond, J.R.; Elkeles, R.S.; Bujac, S.R.; Miller, G.; Talmud, P.J.; Staels, B.; et al. Variation in the PPARalpha gene is associated with altered function in vitro and plasma lipid concentrations in Type II diabetic subjects. Diabetologia 2000, 43, 673–680. [Google Scholar] [CrossRef]
  50. Tanaka, T.; Ordovas, J.M.; Delgado-Lista, J.; Perez-Jimenez, F.; Marin, C.; Perez-Martinez, P.; Gomez, P.; Lopez-Miranda, J. Peroxisome proliferator-activated receptor alpha polymorphisms and postprandial lipemia in healthy men. J. Lipid Res. 2007, 48, 1402–1408. [Google Scholar] [CrossRef] [Green Version]
  51. Andrulionyte, L.; Kuulasmaa, T.; Chiasson, J.L.; Laakso, M.; Group, S.-N.S. Single nucleotide polymorphisms of the peroxisome proliferator-activated receptor-alpha gene (PPARA) influence the conversion from impaired glucose tolerance to type 2 diabetes: The STOP-NIDDM trial. Diabetes 2007, 56, 1181–1186. [Google Scholar] [CrossRef] [Green Version]
  52. Flavell, D.M.; Ireland, H.; Stephens, J.W.; Hawe, E.; Acharya, J.; Mather, H.; Hurel, S.J.; Humphries, S.E. Peroxisome proliferator-activated receptor alpha gene variation influences age of onset and progression of type 2 diabetes. Diabetes 2005, 54, 582–586. [Google Scholar] [CrossRef] [Green Version]
  53. Hashimoto, T.; Cook, W.S.; Qi, C.; Yeldandi, A.V.; Reddy, J.K.; Rao, M.S. Defect in peroxisome proliferator-activated receptor alpha-inducible fatty acid oxidation determines the severity of hepatic steatosis in response to fasting. J. Biol. Chem. 2000, 275, 28918–28928. [Google Scholar] [CrossRef] [Green Version]
  54. Kersten, S.; Seydoux, J.; Peters, J.M.; Gonzalez, F.J.; Desvergne, B.; Wahli, W. Peroxisome proliferator-activated receptor alpha mediates the adaptive response to fasting. J. Clin. Investig. 1999, 103, 1489–1498. [Google Scholar] [CrossRef] [Green Version]
  55. Leone, T.C.; Weinheimer, C.J.; Kelly, D.P. A critical role for the peroxisome proliferator-activated receptor alpha (PPARalpha) in the cellular fasting response: The PPARalpha-null mouse as a model of fatty acid oxidation disorders. Proc. Natl. Acad. Sci. USA 1999, 96, 7473–7478. [Google Scholar] [CrossRef] [Green Version]
  56. Smati, S.; Regnier, M.; Fougeray, T.; Polizzi, A.; Fougerat, A.; Lasserre, F.; Lukowicz, C.; Tramunt, B.; Guillaume, M.; Burnol, A.F.; et al. Regulation of hepatokine gene expression in response to fasting and feeding: Influence of PPAR-alpha and insulin-dependent signalling in hepatocytes. Diabetes Metab. 2019. [Google Scholar] [CrossRef]
  57. Paumelle, R.; Haas, J.T.; Hennuyer, N.; Bauge, E.; Deleye, Y.; Mesotten, D.; Langouche, L.; Vanhoutte, J.; Cudejko, C.; Wouters, K.; et al. Hepatic PPARalpha is critical in the metabolic adaptation to sepsis. J. Hepatol. 2019, 70, 963–973. [Google Scholar] [CrossRef]
  58. Wahli, W.; Michalik, L. PPARs at the crossroads of lipid signaling and inflammation. Trends Endocrinol. Metab. 2012, 23, 351–363. [Google Scholar] [CrossRef]
  59. Grygiel-Gorniak, B. Peroxisome proliferator-activated receptors and their ligands: Nutritional and clinical implications—A review. Nutr. J. 2014, 13, 17. [Google Scholar] [CrossRef] [Green Version]
  60. Liu, Z.M.; Hu, M.; Chan, P.; Tomlinson, B. Early investigational drugs targeting PPAR-alpha for the treatment of metabolic disease. Expert Opin. Investig. Drugs 2015, 24, 611–621. [Google Scholar] [CrossRef]
  61. Taniguchi, A.; Fukushima, M.; Sakai, M.; Tokuyama, K.; Nagata, I.; Fukunaga, A.; Kishimoto, H.; Doi, K.; Yamashita, Y.; Matsuura, T.; et al. Effects of bezafibrate on insulin sensitivity and insulin secretion in non-obese Japanese type 2 diabetic patients. Metabolism 2001, 50, 477–480. [Google Scholar] [CrossRef]
  62. Fruchart, J.C.; Staels, B.; Duriez, P. The role of fibric acids in atherosclerosis. Curr. Atheroscler. Rep. 2001, 3, 83–92. [Google Scholar]
  63. Fruchart, J.C.; Santos, R.D.; Aguilar-Salinas, C.; Aikawa, M.; Al Rasadi, K.; Amarenco, P.; Barter, P.J.; Ceska, R.; Corsini, A.; Despres, J.P.; et al. The selective peroxisome proliferator-activated receptor alpha modulator (SPPARMalpha) paradigm: Conceptual framework and therapeutic potential: A consensus statement from the International Atherosclerosis Society (IAS) and the Residual Risk Reduction Initiative (R3i) Foundation. Cardiovasc. Diabetol. 2019, 18, 71. [Google Scholar] [CrossRef]
  64. Girroir, E.E.; Hollingshead, H.E.; He, P.; Zhu, B.; Perdew, G.H.; Peters, J.M. Quantitative expression patterns of peroxisome proliferator-activated receptor-beta/delta (PPARbeta/delta) protein in mice. Biochem. Biophys. Res. Commun. 2008, 371, 456–461. [Google Scholar] [CrossRef] [Green Version]
  65. Braissant, O.; Wahli, W. Differential expression of peroxisome proliferator-activated receptor-alpha, -beta, and -gamma during rat embryonic development. Endocrinology 1998, 139, 2748–2754. [Google Scholar] [CrossRef]
  66. Tan, N.S.; Michalik, L.; Desvergne, B.; Wahli, W. Peroxisome proliferator-activated receptor (PPAR)-beta as a target for wound healing drugs: What is possible? Am. J. Clin. Dermatol. 2003, 4, 523–530. [Google Scholar] [CrossRef]
  67. Di-Poi, N.; Tan, N.S.; Michalik, L.; Wahli, W.; Desvergne, B. Antiapoptotic role of PPARbeta in keratinocytes via transcriptional control of the Akt1 signaling pathway. Mol. Cell 2002, 10, 721–733. [Google Scholar] [CrossRef]
  68. Letavernier, E.; Perez, J.; Joye, E.; Bellocq, A.; Fouqueray, B.; Haymann, J.P.; Heudes, D.; Wahli, W.; Desvergne, B.; Baud, L. Peroxisome proliferator-activated receptor beta/delta exerts a strong protection from ischemic acute renal failure. J. Am. Soc. Nephrol. 2005, 16, 2395–2402. [Google Scholar] [CrossRef] [Green Version]
  69. Michalik, L.; Wahli, W. Involvement of PPAR nuclear receptors in tissue injury and wound repair. J. Clin. Investig. 2006, 116, 598–606. [Google Scholar] [CrossRef]
  70. Peters, J.M.; Lee, S.S.; Li, W.; Ward, J.M.; Gavrilova, O.; Everett, C.; Reitman, M.L.; Hudson, L.D.; Gonzalez, F.J. Growth, adipose, brain, and skin alterations resulting from targeted disruption of the mouse peroxisome proliferator-activated receptor beta(delta). Mol. Cell Biol. 2000, 20, 5119–5128. [Google Scholar] [CrossRef] [Green Version]
  71. Barak, Y.; Liao, D.; He, W.; Ong, E.S.; Nelson, M.C.; Olefsky, J.M.; Boland, R.; Evans, R.M. Effects of peroxisome proliferator-activated receptor delta on placentation, adiposity, and colorectal cancer. Proc. Natl. Acad. Sci. USA 2002, 99, 303–308. [Google Scholar] [CrossRef] [Green Version]
  72. Wang, Y.X.; Lee, C.H.; Tiep, S.; Yu, R.T.; Ham, J.; Kang, H.; Evans, R.M. Peroxisome-proliferator-activated receptor delta activates fat metabolism to prevent obesity. Cell 2003, 113, 159–170. [Google Scholar] [CrossRef] [Green Version]
  73. Nadra, K.; Anghel, S.I.; Joye, E.; Tan, N.S.; Basu-Modak, S.; Trono, D.; Wahli, W.; Desvergne, B. Differentiation of trophoblast giant cells and their metabolic functions are dependent on peroxisome proliferator-activated receptor beta/delta. Mol. Cell Biol. 2006, 26, 3266–3281. [Google Scholar] [CrossRef] [Green Version]
  74. Varnat, F.; Heggeler, B.B.; Grisel, P.; Boucard, N.; Corthesy-Theulaz, I.; Wahli, W.; Desvergne, B. PPARbeta/delta regulates paneth cell differentiation via controlling the hedgehog signaling pathway. Gastroenterology 2006, 131, 538–553. [Google Scholar] [CrossRef]
  75. Doktorova, M.; Zwarts, I.; Zutphen, T.V.; Dijk, T.H.; Bloks, V.W.; Harkema, L.; Bruin, A.; Downes, M.; Evans, R.M.; Verkade, H.J.; et al. Intestinal PPARdelta protects against diet-induced obesity, insulin resistance and dyslipidemia. Sci. Rep. 2017, 7, 846. [Google Scholar] [CrossRef] [Green Version]
  76. Dreyer, C.; Krey, G.; Keller, H.; Givel, F.; Helftenbein, G.; Wahli, W. Control of the peroxisomal beta-oxidation pathway by a novel family of nuclear hormone receptors. Cell 1992, 68, 879–887. [Google Scholar] [CrossRef]
  77. Daoudi, M.; Hennuyer, N.; Borland, M.G.; Touche, V.; Duhem, C.; Gross, B.; Caiazzo, R.; Kerr-Conte, J.; Pattou, F.; Peters, J.M.; et al. PPARbeta/delta activation induces enteroendocrine L cell GLP-1 production. Gastroenterology 2011, 140, 1564–1574. [Google Scholar] [CrossRef] [Green Version]
  78. Schuler, M.; Ali, F.; Chambon, C.; Duteil, D.; Bornert, J.M.; Tardivel, A.; Desvergne, B.; Wahli, W.; Chambon, P.; Metzger, D. PGC1alpha expression is controlled in skeletal muscles by PPARbeta, whose ablation results in fiber-type switching, obesity, and type 2 diabetes. Cell Metab. 2006, 4, 407–414. [Google Scholar] [CrossRef]
  79. Manickam, R.; Wahli, W. Roles of Peroxisome Proliferator-Activated Receptor beta/delta in skeletal muscle physiology. Biochimie 2017, 136, 42–48. [Google Scholar] [CrossRef]
  80. Wang, Y.X.; Zhang, C.L.; Yu, R.T.; Cho, H.K.; Nelson, M.C.; Bayuga-Ocampo, C.R.; Ham, J.; Kang, H.; Evans, R.M. Regulation of muscle fiber type and running endurance by PPARdelta. PLoS Biol. 2004, 2, e294. [Google Scholar] [CrossRef] [Green Version]
  81. Giordano Attianese, G.M.; Desvergne, B. Integrative and systemic approaches for evaluating PPARbeta/delta (PPARD) function. Nucl. Recept. Signal. 2015, 13, e001. [Google Scholar] [CrossRef] [Green Version]
  82. Anghel, S.I.; Wahli, W. Fat poetry: A kingdom for PPAR gamma. Cell Res. 2007, 17, 486–511. [Google Scholar] [CrossRef] [Green Version]
  83. Lehmann, J.M.; Moore, L.B.; Smith-Oliver, T.A.; Wilkison, W.O.; Willson, T.M.; Kliewer, S.A. An antidiabetic thiazolidinedione is a high affinity ligand for peroxisome proliferator-activated receptor gamma (PPAR gamma). J. Biol. Chem. 1995, 270, 12953–12956. [Google Scholar] [CrossRef] [Green Version]
  84. Imai, T.; Takakuwa, R.; Marchand, S.; Dentz, E.; Bornert, J.M.; Messaddeq, N.; Wendling, O.; Mark, M.; Desvergne, B.; Wahli, W.; et al. Peroxisome proliferator-activated receptor gamma is required in mature white and brown adipocytes for their survival in the mouse. Proc. Natl. Acad. Sci. USA 2004, 101, 4543–4547. [Google Scholar] [CrossRef] [Green Version]
  85. Cerbone, A.; Toaldo, C.; Laurora, S.; Briatore, F.; Pizzimenti, S.; Dianzani, M.U.; Ferretti, C.; Barrera, G. 4-Hydroxynonenal and PPARgamma ligands affect proliferation, differentiation, and apoptosis in colon cancer cells. Free Radic. Biol. Med. 2007, 42, 1661–1670. [Google Scholar] [CrossRef]
  86. Martinasso, G.; Oraldi, M.; Trombetta, A.; Maggiora, M.; Bertetto, O.; Canuto, R.A.; Muzio, G. Involvement of PPARs in Cell Proliferation and Apoptosis in Human Colon Cancer Specimens and in Normal and Cancer Cell Lines. PPAR Res. 2007, 2007, 93416. [Google Scholar] [CrossRef] [Green Version]
  87. Theocharis, S.; Margeli, A.; Vielh, P.; Kouraklis, G. Peroxisome proliferator-activated receptor-gamma ligands as cell-cycle modulators. Cancer Treat. Rev. 2004, 30, 545–554. [Google Scholar] [CrossRef]
  88. Xu, W.P.; Zhang, X.; Xie, W.F. Differentiation therapy for solid tumors. J. Dig. Dis. 2014, 15, 159–165. [Google Scholar] [CrossRef]
  89. Chen, G.G.; Lee, J.F.; Wang, S.H.; Chan, U.P.; Ip, P.C.; Lau, W.Y. Apoptosis induced by activation of peroxisome-proliferator activated receptor-gamma is associated with Bcl-2 and NF-kappaB in human colon cancer. Life Sci. 2002, 70, 2631–2646. [Google Scholar]
  90. Chen, G.G.; Xu, H.; Lee, J.F.; Subramaniam, M.; Leung, K.L.; Wang, S.H.; Chan, U.P.; Spelsberg, T.C. 15-hydroxy-eicosatetraenoic acid arrests growth of colorectal cancer cells via a peroxisome proliferator-activated receptor gamma-dependent pathway. Int. J. Cancer 2003, 107, 837–843. [Google Scholar] [CrossRef]
  91. Lee, C.J.; Han, J.S.; Seo, C.Y.; Park, T.H.; Kwon, H.C.; Jeong, J.S.; Kim, I.H.; Yun, J.; Bae, Y.S.; Kwak, J.Y.; et al. Pioglitazone, a synthetic ligand for PPARgamma, induces apoptosis in RB-deficient human colorectal cancer cells. Apoptosis Int. J. Program. Cell Death 2006, 11, 401–411. [Google Scholar] [CrossRef]
  92. Clay, C.E.; Monjazeb, A.; Thorburn, J.; Chilton, F.H.; High, K.P. 15-Deoxy-delta12,14-prostaglandin J2-induced apoptosis does not require PPARgamma in breast cancer cells. J. Lipid Res. 2002, 43, 1818–1828. [Google Scholar] [CrossRef] [Green Version]
  93. Guan, Y.F.; Zhang, Y.H.; Breyer, R.M.; Davis, L.; Breyer, M.D. Expression of peroxisome proliferator-activated receptor gamma (PPARgamma) in human transitional bladder cancer and its role in inducing cell death. Neoplasia 1999, 1, 330–339. [Google Scholar] [CrossRef] [Green Version]
  94. Sharma, C.; Pradeep, A.; Wong, L.; Rana, A.; Rana, B. Peroxisome proliferator-activated receptor gamma activation can regulate beta-catenin levels via a proteasome-mediated and adenomatous polyposis coli-independent pathway. J. Biol. Chem. 2004, 279, 35583–35594. [Google Scholar] [CrossRef] [Green Version]
  95. Auwerx, J. Nuclear receptors. I. PPAR gamma in the gastrointestinal tract: Gain or pain? Am. J. Physiol. Gastrointest. Liver Physiol. 2002, 282, G581–G585. [Google Scholar] [CrossRef] [Green Version]
  96. Dubuquoy, L.; Rousseaux, C.; Thuru, X.; Peyrin-Biroulet, L.; Romano, O.; Chavatte, P.; Chamaillard, M.; Desreumaux, P. PPARgamma as a new therapeutic target in inflammatory bowel diseases. Gut 2006, 55, 1341–1349. [Google Scholar] [CrossRef] [Green Version]
  97. Leonardini, A.; Laviola, L.; Perrini, S.; Natalicchio, A.; Giorgino, F. Cross-Talk between PPARgamma and Insulin Signaling and Modulation of Insulin Sensitivity. PPAR Res. 2009, 2009, 818945. [Google Scholar] [CrossRef] [Green Version]
  98. Tontonoz, P.; Spiegelman, B.M. Fat and beyond: The diverse biology of PPARgamma. Annu. Rev. Biochem. 2008, 77, 289–312. [Google Scholar] [CrossRef]
  99. Picard, F.; Auwerx, J. PPAR(gamma) and glucose homeostasis. Annu. Rev. Nutr. 2002, 22, 167–197. [Google Scholar] [CrossRef]
  100. Duszka, K.; Oresic, M.; Le May, C.; Konig, J.; Wahli, W. PPARgamma Modulates Long Chain Fatty Acid Processing in the Intestinal Epithelium. Int. J. Mol. Sci. 2017, 18, 2559. [Google Scholar] [CrossRef] [Green Version]
  101. Duszka, K.; Picard, A.; Ellero-Simatos, S.; Chen, J.; Defernez, M.; Paramalingam, E.; Pigram, A.; Vanoaica, L.; Canlet, C.; Parini, P.; et al. Intestinal PPARgamma signalling is required for sympathetic nervous system activation in response to caloric restriction. Sci. Rep. 2016, 6, 36937. [Google Scholar] [CrossRef]
  102. Tomas, J.; Mulet, C.; Saffarian, A.; Cavin, J.B.; Ducroc, R.; Regnault, B.; Kun Tan, C.; Duszka, K.; Burcelin, R.; Wahli, W.; et al. High-fat diet modifies the PPAR-gamma pathway leading to disruption of microbial and physiological ecosystem in murine small intestine. Proc. Natl. Acad. Sci. USA 2016, 113, E5934–E5943. [Google Scholar] [CrossRef] [Green Version]
  103. Zou, Y.; Watters, A.; Cheng, N.; Perry, C.E.; Xu, K.; Alicea, G.M.; Parris, J.L.D.; Baraban, E.; Ray, P.; Nayak, A.; et al. Polyunsaturated Fatty Acids from Astrocytes Activate PPARgamma Signaling in Cancer Cells to Promote Brain Metastasis. Cancer Discov. 2019, 9, 1720–1735. [Google Scholar] [CrossRef] [Green Version]
  104. Patitucci, C.; Couchy, G.; Bagattin, A.; Caneque, T.; de Reynies, A.; Scoazec, J.Y.; Rodriguez, R.; Pontoglio, M.; Zucman-Rossi, J.; Pende, M.; et al. Hepatocyte nuclear factor 1alpha suppresses steatosis-associated liver cancer by inhibiting PPARgamma transcription. J. Clin. Investig. 2017, 127, 1873–1888. [Google Scholar] [CrossRef] [Green Version]
  105. Zhu, Y.; Qi, C.; Korenberg, J.R.; Chen, X.N.; Noya, D.; Rao, M.S.; Reddy, J.K. Structural organization of mouse peroxisome proliferator-activated receptor gamma (mPPAR gamma) gene: Alternative promoter use and different splicing yield two mPPAR gamma isoforms. Proc. Natl. Acad. Sci. USA 1995, 92, 7921–7925. [Google Scholar] [CrossRef] [Green Version]
  106. Chawla, A.; Schwarz, E.J.; Dimaculangan, D.D.; Lazar, M.A. Peroxisome proliferator-activated receptor (PPAR) gamma: Adipose-predominant expression and induction early in adipocyte differentiation. Endocrinology 1994, 135, 798–800. [Google Scholar] [CrossRef]
  107. Tontonoz, P.; Hu, E.; Graves, R.A.; Budavari, A.I.; Spiegelman, B.M. mPPAR gamma 2: Tissue-specific regulator of an adipocyte enhancer. Genes Dev. 1994, 8, 1224–1234. [Google Scholar] [CrossRef] [Green Version]
  108. Elbrecht, A.; Chen, Y.; Cullinan, C.A.; Hayes, N.; Leibowitz, M.; Moller, D.E.; Berger, J. Molecular cloning, expression and characterization of human peroxisome proliferator activated receptors gamma 1 and gamma 2. Biochem. Biophys. Res. Commun. 1996, 224, 431–437. [Google Scholar] [CrossRef]
  109. Greene, M.E.; Blumberg, B.; McBride, O.W.; Yi, H.F.; Kronquist, K.; Kwan, K.; Hsieh, L.; Greene, G.; Nimer, S.D. Isolation of the human peroxisome proliferator activated receptor gamma cDNA: Expression in hematopoietic cells and chromosomal mapping. Gene Expr. 1995, 4, 281–299. [Google Scholar]
  110. Mukherjee, R.; Jow, L.; Croston, G.E.; Paterniti, J.R., Jr. Identification, characterization, and tissue distribution of human peroxisome proliferator-activated receptor (PPAR) isoforms PPARgamma2 versus PPARgamma1 and activation with retinoid X receptor agonists and antagonists. J. Biol. Chem. 1997, 272, 8071–8076. [Google Scholar] [CrossRef] [Green Version]
  111. Trombetta, A.; Maggiora, M.; Martinasso, G.; Cotogni, P.; Canuto, R.A.; Muzio, G. Arachidonic and docosahexaenoic acids reduce the growth of A549 human lung-tumor cells increasing lipid peroxidation and PPARs. Chem. Biol. Interact. 2007, 165, 239–250. [Google Scholar] [CrossRef]
  112. Forman, B.M.; Tontonoz, P.; Chen, J.; Brun, R.P.; Spiegelman, B.M.; Evans, R.M. 15-Deoxy-delta 12, 14-prostaglandin J2 is a ligand for the adipocyte determination factor PPAR gamma. Cell 1995, 83, 803–812. [Google Scholar] [CrossRef] [Green Version]
  113. Heim, M.; Johnson, J.; Boess, F.; Bendik, I.; Weber, P.; Hunziker, W.; Fluhmann, B. Phytanic acid, a natural peroxisome proliferator-activated receptor (PPAR) agonist, regulates glucose metabolism in rat primary hepatocytes. FASEB J. 2002, 16, 718–720. [Google Scholar] [CrossRef] [Green Version]
  114. Marion-Letellier, R.; Dechelotte, P.; Iacucci, M.; Ghosh, S. Dietary modulation of peroxisome proliferator-activated receptor gamma. Gut 2009, 58, 586–593. [Google Scholar] [CrossRef]
  115. Schwab, M.; Reynders, V.; Loitsch, S.; Steinhilber, D.; Stein, J.; Schroder, O. Involvement of different nuclear hormone receptors in butyrate-mediated inhibition of inducible NF kappa B signalling. Mol. Immunol. 2007, 44, 3625–3632. [Google Scholar] [CrossRef]
  116. Wachtershauser, A.; Loitsch, S.M.; Stein, J. PPAR-gamma is selectively upregulated in Caco-2 cells by butyrate. Biochem. Biophys. Res. Commun. 2000, 272, 380–385. [Google Scholar]
  117. Nepelska, M.; de Wouters, T.; Jacouton, E.; Beguet-Crespel, F.; Lapaque, N.; Dore, J.; Arulampalam, V.; Blottiere, H.M. Commensal gut bacteria modulate phosphorylation-dependent PPARgamma transcriptional activity in human intestinal epithelial cells. Sci. Rep. 2017, 7, 43199. [Google Scholar] [CrossRef] [Green Version]
  118. Voltan, S.; Martines, D.; Elli, M.; Brun, P.; Longo, S.; Porzionato, A.; Macchi, V.; D’Inca, R.; Scarpa, M.; Palu, G.; et al. Lactobacillus crispatus M247-derived H2O2 acts as a signal transducing molecule activating peroxisome proliferator activated receptor-gamma in the intestinal mucosa. Gastroenterology 2008, 135, 1216–1227. [Google Scholar] [CrossRef]
  119. Are, A.; Aronsson, L.; Wang, S.; Greicius, G.; Lee, Y.K.; Gustafsson, J.A.; Pettersson, S.; Arulampalam, V. Enterococcus faecalis from newborn babies regulate endogenous PPARgamma activity and IL-10 levels in colonic epithelial cells. Proc. Natl. Acad. Sci. USA 2008, 105, 1943–1948. [Google Scholar] [CrossRef] [Green Version]
  120. Couvigny, B.; de Wouters, T.; Kaci, G.; Jacouton, E.; Delorme, C.; Dore, J.; Renault, P.; Blottiere, H.M.; Guedon, E.; Lapaque, N. Commensal Streptococcus salivarius Modulates PPARgamma Transcriptional Activity in Human Intestinal Epithelial Cells. PLoS ONE 2015, 10, e0125371. [Google Scholar] [CrossRef] [Green Version]
  121. Soccio, R.E.; Chen, E.R.; Lazar, M.A. Thiazolidinediones and the promise of insulin sensitization in type 2 diabetes. Cell Metab. 2014, 20, 573–591. [Google Scholar] [CrossRef] [Green Version]
  122. Burgermeister, E.; Schnoebelen, A.; Flament, A.; Benz, J.; Stihle, M.; Gsell, B.; Rufer, A.; Ruf, A.; Kuhn, B.; Marki, H.P.; et al. A novel partial agonist of peroxisome proliferator-activated receptor-gamma (PPARgamma) recruits PPARgamma-coactivator-1alpha, prevents triglyceride accumulation, and potentiates insulin signaling in vitro. Mol. Endocrinol. 2006, 20, 809–830. [Google Scholar] [CrossRef] [Green Version]
  123. Hou, Y.; Cao, X.; Hu, X.; Li, X.; Shi, X.; Wang, H.; Peng, C.; Li, J.; Li, J.; Li, Q.; et al. CMHX008, a PPARgamma partial agonist, enhances insulin sensitivity with minor influences on bone loss. Genes Dis. 2018, 5, 290–299. [Google Scholar] [CrossRef]
  124. Henke, B.R. Peroxisome proliferator-activated receptor alpha/gamma dual agonists for the treatment of type 2 diabetes. J. Med. Chem. 2004, 47, 4118–4127. [Google Scholar] [CrossRef]
  125. Kim, S.H.; Hong, S.H.; Park, Y.J.; Sung, J.H.; Suh, W.; Lee, K.W.; Jung, K.; Lim, C.; Kim, J.H.; Kim, H.; et al. MD001, a Novel Peroxisome Proliferator-activated Receptor alpha/gamma Agonist, Improves Glucose and Lipid Metabolism. Sci. Rep. 2019, 9, 1656. [Google Scholar] [CrossRef]
  126. Jeong, H.W.; Lee, J.W.; Kim, W.S.; Choe, S.S.; Kim, K.H.; Park, H.S.; Shin, H.J.; Lee, G.Y.; Shin, D.; Lee, H.; et al. A newly identified CG301269 improves lipid and glucose metabolism without body weight gain through activation of peroxisome proliferator-activated receptor alpha and gamma. Diabetes 2011, 60, 496–506. [Google Scholar] [CrossRef] [Green Version]
  127. Xu, C.; Wang, L.L.; Liu, H.Y.; Zhou, X.B.; Cao, Y.L.; Li, S. C333H, a novel PPARalpha/gamma dual agonist, has beneficial effects on insulin resistance and lipid metabolism. Acta Pharmacol. Sin. 2006, 27, 223–228. [Google Scholar] [CrossRef] [Green Version]
  128. Park, M.H.; Park, J.Y.; Lee, H.J.; Kim, D.H.; Park, D.; Jeong, H.O.; Park, C.H.; Chun, P.; Moon, H.R.; Chung, H.Y. Potent anti-diabetic effects of MHY908, a newly synthesized PPAR alpha/gamma dual agonist in db/db mice. PLoS ONE 2013, 8, e78815. [Google Scholar] [CrossRef] [Green Version]
  129. Waites, C.R.; Dominick, M.A.; Sanderson, T.P.; Schilling, B.E. Nonclinical safety evaluation of muraglitazar, a novel PPARalpha/gamma agonist. Toxicol. Sci. 2007, 100, 248–258. [Google Scholar] [CrossRef] [Green Version]
  130. Chakrabarti, R.; Vikramadithyan, R.K.; Misra, P.; Hiriyan, J.; Raichur, S.; Damarla, R.K.; Gershome, C.; Suresh, J.; Rajagopalan, R. Ragaglitazar: A novel PPAR alpha PPAR gamma agonist with potent lipid-lowering and insulin-sensitizing efficacy in animal models. Br. J. Pharmacol. 2003, 140, 527–537. [Google Scholar] [CrossRef] [Green Version]
  131. Masternak, M.M.; Bartke, A. PPARs in Calorie Restricted and Genetically Long-Lived Mice. PPAR Res. 2007, 2007, 28436. [Google Scholar] [CrossRef] [Green Version]
  132. Yang, Y.B.; Wu, X.L.; Ke, B.; Huang, Y.J.; Chen, S.Q.; Su, Y.Q.; Qin, J. Effects of caloric restriction on peroxisome proliferator-activated receptors and positive transcription elongation factor b expression in obese rats. Eur. Rev. Med. Pharmacol. Sci. 2017, 21, 4369–4378. [Google Scholar]
  133. Masternak, M.M.; Al-Regaiey, K.A.; Del Rosario Lim, M.M.; Jimenez-Ortega, V.; Panici, J.A.; Bonkowski, M.S.; Kopchick, J.J.; Wang, Z.; Bartke, A. Caloric restriction and growth hormone receptor knockout: Effects on expression of genes involved in insulin action in the heart. Exp. Gerontol. 2006, 41, 417–429. [Google Scholar] [CrossRef] [Green Version]
  134. Masternak, M.M.; Al-Regaiey, K.A.; Del Rosario Lim, M.M.; Jimenez-Ortega, V.; Panici, J.A.; Bonkowski, M.S.; Kopchick, J.J.; Bartke, A. Effects of caloric restriction and growth hormone resistance on the expression level of peroxisome proliferator-activated receptors superfamily in liver of normal and long-lived growth hormone receptor/binding protein knockout mice. J. Gerontol. Biol. Sci. Med. Sci. 2005, 60, 1394–1398. [Google Scholar] [CrossRef]
  135. Masternak, M.M.; Al-Regaiey, K.; Bonkowski, M.S.; Panici, J.; Sun, L.; Wang, J.; Przybylski, G.K.; Bartke, A. Divergent effects of caloric restriction on gene expression in normal and long-lived mice. J. Gerontol. A Biol. Sci. Med. Sci. 2004, 59, 784–788. [Google Scholar] [CrossRef] [Green Version]
  136. Masternak, M.M.; Al-Regaiey, K.A.; Del Rosario Lim, M.M.; Bonkowski, M.S.; Panici, J.A.; Przybylski, G.K.; Bartke, A. Caloric restriction results in decreased expression of peroxisome proliferator-activated receptor superfamily in muscle of normal and long-lived growth hormone receptor/binding protein knockout mice. J. Gerontol. Biol. Sci. Med. Sci. 2005, 60, 1238–1245. [Google Scholar] [CrossRef]
  137. Poynter, M.E.; Daynes, R.A. Peroxisome proliferator-activated receptor alpha activation modulates cellular redox status, represses nuclear factor-kappaB signaling, and reduces inflammatory cytokine production in aging. J. Biol. Chem. 1998, 273, 32833–32841. [Google Scholar]
  138. Duszka, K.; Ellero-Simatos, S.; Ow, G.S.; Defernez, M.; Paramalingam, E.; Tett, A.; Ying, S.; Konig, J.; Narbad, A.; Kuznetsov, V.A.; et al. Complementary intestinal mucosa and microbiota responses to caloric restriction. Sci. Rep. 2018, 8, 11338. [Google Scholar] [CrossRef]
  139. Dhahbi, J.M.; Tsuchiya, T.; Kim, H.J.; Mote, P.L.; Spindler, S.R. Gene expression and physiologic responses of the heart to the initiation and withdrawal of caloric restriction. J. Gerontol. Biol. Sci. Med. Sci. 2006, 61, 218–231. [Google Scholar] [CrossRef] [Green Version]
  140. Sung, B.; Park, S.; Yu, B.P.; Chung, H.Y. Modulation of PPAR in aging, inflammation, and calorie restriction. J. Gerontol. Biol. Sci. Med. Sci. 2004, 59, 997–1006. [Google Scholar] [CrossRef] [Green Version]
  141. Corton, J.C.; Apte, U.; Anderson, S.P.; Limaye, P.; Yoon, L.; Latendresse, J.; Dunn, C.; Everitt, J.I.; Voss, K.A.; Swanson, C.; et al. Mimetics of caloric restriction include agonists of lipid-activated nuclear receptors. J. Biol. Chem. 2004, 279, 46204–46212. [Google Scholar] [CrossRef] [Green Version]
  142. Bhaskar, P.T.; Hay, N. The two TORCs and Akt. Dev. Cell 2007, 12, 487–502. [Google Scholar] [CrossRef] [Green Version]
  143. Wullschleger, S.; Loewith, R.; Hall, M.N. TOR signaling in growth and metabolism. Cell 2006, 124, 471–484. [Google Scholar] [CrossRef] [Green Version]
  144. Jacinto, E.; Facchinetti, V.; Liu, D.; Soto, N.; Wei, S.; Jung, S.Y.; Huang, Q.; Qin, J.; Su, B. SIN1/MIP1 maintains rictor-mTOR complex integrity and regulates Akt phosphorylation and substrate specificity. Cell 2006, 127, 125–137. [Google Scholar] [CrossRef] [Green Version]
  145. Pearce, L.R.; Huang, X.; Boudeau, J.; Pawlowski, R.; Wullschleger, S.; Deak, M.; Ibrahim, A.F.; Gourlay, R.; Magnuson, M.A.; Alessi, D.R. Identification of Protor as a novel Rictor-binding component of mTOR complex-2. Biochem. J. 2007, 405, 513–522. [Google Scholar] [CrossRef] [Green Version]
  146. Sarbassov, D.D.; Ali, S.M.; Kim, D.H.; Guertin, D.A.; Latek, R.R.; Erdjument-Bromage, H.; Tempst, P.; Sabatini, D.M. Rictor, a novel binding partner of mTOR, defines a rapamycin-insensitive and raptor-independent pathway that regulates the cytoskeleton. Curr. Biol. 2004, 14, 1296–1302. [Google Scholar] [CrossRef] [Green Version]
  147. Zheng, X.; Liang, Y.; He, Q.; Yao, R.; Bao, W.; Bao, L.; Wang, Y.; Wang, Z. Current models of mammalian target of rapamycin complex 1 (mTORC1) activation by growth factors and amino acids. Int. J. Mol. Sci. 2014, 15, 20753–20769. [Google Scholar] [CrossRef] [Green Version]
  148. Tatebe, H.; Shiozaki, K. Evolutionary Conservation of the Components in the TOR Signaling Pathways. Biomolecules 2017, 7, 77. [Google Scholar] [CrossRef] [Green Version]
  149. Loewith, R.; Jacinto, E.; Wullschleger, S.; Lorberg, A.; Crespo, J.L.; Bonenfant, D.; Oppliger, W.; Jenoe, P.; Hall, M.N. Two TOR complexes, only one of which is rapamycin sensitive, have distinct roles in cell growth control. Mol. Cell 2002, 10, 457–468. [Google Scholar]
  150. Jacinto, E.; Loewith, R.; Schmidt, A.; Lin, S.; Ruegg, M.A.; Hall, A.; Hall, M.N. Mammalian TOR complex 2 controls the actin cytoskeleton and is rapamycin insensitive. Nat. Cell Biol. 2004, 6, 1122–1128. [Google Scholar] [CrossRef]
  151. Yamagata, K.; Sanders, L.K.; Kaufmann, W.E.; Yee, W.; Barnes, C.A.; Nathans, D.; Worley, P.F. rheb, a growth factor- and synaptic activity-regulated gene, encodes a novel Ras-related protein. J. Biol. Chem. 1994, 269, 16333–16339. [Google Scholar]
  152. Long, X.; Lin, Y.; Ortiz-Vega, S.; Yonezawa, K.; Avruch, J. Rheb binds and regulates the mTOR kinase. Curr. Biol. 2005, 15, 702–713. [Google Scholar] [CrossRef] [Green Version]
  153. Sato, T.; Nakashima, A.; Guo, L.; Tamanoi, F. Specific activation of mTORC1 by Rheb G-protein in vitro involves enhanced recruitment of its substrate protein. J. Biol. Chem. 2009, 284, 12783–12791. [Google Scholar] [CrossRef] [Green Version]
  154. Inoki, K.; Li, Y.; Xu, T.; Guan, K.L. Rheb GTPase is a direct target of TSC2 GAP activity and regulates mTOR signaling. Genes Dev. 2003, 17, 1829–1834. [Google Scholar] [CrossRef] [Green Version]
  155. Garami, A.; Zwartkruis, F.J.; Nobukuni, T.; Joaquin, M.; Roccio, M.; Stocker, H.; Kozma, S.C.; Hafen, E.; Bos, J.L.; Thomas, G. Insulin activation of Rheb, a mediator of mTOR/S6K/4E-BP signaling, is inhibited by TSC1 and 2. Mol. Cell 2003, 11, 1457–1466. [Google Scholar]
  156. Wang, X.; Proud, C.G. The mTOR pathway in the control of protein synthesis. Physiology 2006, 21, 362–369. [Google Scholar] [CrossRef] [Green Version]
  157. Tee, A.R.; Manning, B.D.; Roux, P.P.; Cantley, L.C.; Blenis, J. Tuberous sclerosis complex gene products, Tuberin and Hamartin, control mTOR signaling by acting as a GTPase-activating protein complex toward Rheb. Curr. Biol. 2003, 13, 1259–1268. [Google Scholar]
  158. Zhang, Y.; Gao, X.; Saucedo, L.J.; Ru, B.; Edgar, B.A.; Pan, D. Rheb is a direct target of the tuberous sclerosis tumour suppressor proteins. Nat. Cell Biol. 2003, 5, 578–581. [Google Scholar] [CrossRef]
  159. Kwiatkowski, D.J. Tuberous sclerosis: From tubers to mTOR. Ann. Hum. Genet. 2003, 67, 87–96. [Google Scholar]
  160. Miyazaki, M.; McCarthy, J.J.; Esser, K.A. Insulin like growth factor-1-induced phosphorylation and altered distribution of tuberous sclerosis complex (TSC)1/TSC2 in C2C12 myotubes. FEBS J. 2010, 277, 2180–2191. [Google Scholar] [CrossRef] [Green Version]
  161. Dan, H.C.; Sun, M.; Yang, L.; Feldman, R.I.; Sui, X.M.; Ou, C.C.; Nellist, M.; Yeung, R.S.; Halley, D.J.; Nicosia, S.V.; et al. Phosphatidylinositol 3-kinase/Akt pathway regulates tuberous sclerosis tumor suppressor complex by phosphorylation of tuberin. J. Biol. Chem. 2002, 277, 35364–35370. [Google Scholar] [CrossRef] [Green Version]
  162. Manning, B.D.; Tee, A.R.; Logsdon, M.N.; Blenis, J.; Cantley, L.C. Identification of the tuberous sclerosis complex-2 tumor suppressor gene product tuberin as a target of the phosphoinositide 3-kinase/akt pathway. Mol. Cell 2002, 10, 151–162. [Google Scholar]
  163. Inoki, K.; Li, Y.; Zhu, T.; Wu, J.; Guan, K.L. TSC2 is phosphorylated and inhibited by Akt and suppresses mTOR signalling. Nat. Cell Biol. 2002, 4, 648–657. [Google Scholar] [CrossRef]
  164. Ma, L.; Chen, Z.; Erdjument-Bromage, H.; Tempst, P.; Pandolfi, P.P. Phosphorylation and functional inactivation of TSC2 by Erk implications for tuberous sclerosis and cancer pathogenesis. Cell 2005, 121, 179–193. [Google Scholar] [CrossRef] [Green Version]
  165. Roux, P.P.; Ballif, B.A.; Anjum, R.; Gygi, S.P.; Blenis, J. Tumor-promoting phorbol esters and activated Ras inactivate the tuberous sclerosis tumor suppressor complex via p90 ribosomal S6 kinase. Proc. Natl. Acad. Sci. USA 2004, 101, 13489–13494. [Google Scholar] [CrossRef] [Green Version]
  166. Potter, C.J.; Pedraza, L.G.; Xu, T. Akt regulates growth by directly phosphorylating Tsc2. Nat. Cell Biol. 2002, 4, 658–665. [Google Scholar] [CrossRef]
  167. Vander Haar, E.; Lee, S.I.; Bandhakavi, S.; Griffin, T.J.; Kim, D.H. Insulin signalling to mTOR mediated by the Akt/PKB substrate PRAS40. Nat. Cell Biol. 2007, 9, 316–323. [Google Scholar] [CrossRef]
  168. Zheng, M.; Wang, Y.H.; Wu, X.N.; Wu, S.Q.; Lu, B.J.; Dong, M.Q.; Zhang, H.; Sun, P.; Lin, S.C.; Guan, K.L.; et al. Inactivation of Rheb by PRAK-mediated phosphorylation is essential for energy-depletion-induced suppression of mTORC1. Nat. Cell Biol. 2011, 13, 263–272. [Google Scholar] [CrossRef]
  169. Yin, Y.; Hua, H.; Li, M.; Liu, S.; Kong, Q.; Shao, T.; Wang, J.; Luo, Y.; Wang, Q.; Luo, T.; et al. mTORC2 promotes type I insulin-like growth factor receptor and insulin receptor activation through the tyrosine kinase activity of mTOR. Cell Res. 2016, 26, 46–65. [Google Scholar] [CrossRef] [Green Version]
  170. Arsham, A.M.; Neufeld, T.P. Thinking globally and acting locally with TOR. Curr. Opin. Cell Biol. 2006, 18, 589–597. [Google Scholar] [CrossRef]
  171. Lee, D.F.; Kuo, H.P.; Chen, C.T.; Hsu, J.M.; Chou, C.K.; Wei, Y.; Sun, H.L.; Li, L.Y.; Ping, B.; Huang, W.C.; et al. IKK beta suppression of TSC1 links inflammation and tumor angiogenesis via the mTOR pathway. Cell 2007, 130, 440–455. [Google Scholar] [CrossRef] [Green Version]
  172. Inoki, K.; Ouyang, H.; Zhu, T.; Lindvall, C.; Wang, Y.; Zhang, X.; Yang, Q.; Bennett, C.; Harada, Y.; Stankunas, K.; et al. TSC2 integrates Wnt and energy signals via a coordinated phosphorylation by AMPK and GSK3 to regulate cell growth. Cell 2006, 126, 955–968. [Google Scholar] [CrossRef] [Green Version]
  173. Inoki, K.; Zhu, T.; Guan, K.L. TSC2 mediates cellular energy response to control cell growth and survival. Cell 2003, 115, 577–590. [Google Scholar]
  174. Brugarolas, J.; Lei, K.; Hurley, R.L.; Manning, B.D.; Reiling, J.H.; Hafen, E.; Witters, L.A.; Ellisen, L.W.; Kaelin, W.G., Jr. Regulation of mTOR function in response to hypoxia by REDD1 and the TSC1/TSC2 tumor suppressor complex. Genes Dev. 2004, 18, 2893–2904. [Google Scholar] [CrossRef] [Green Version]
  175. DeYoung, M.P.; Horak, P.; Sofer, A.; Sgroi, D.; Ellisen, L.W. Hypoxia regulates TSC1/2-mTOR signaling and tumor suppression through REDD1-mediated 14-3-3 shuttling. Genes Dev. 2008, 22, 239–251. [Google Scholar] [CrossRef] [Green Version]
  176. Reiling, J.H.; Hafen, E. The hypoxia-induced paralogs Scylla and CharybDis. inhibit growth by down-regulating S6K activity upstream of TSC in Drosophila. Genes Dev. 2004, 18, 2879–2892. [Google Scholar] [CrossRef] [Green Version]
  177. Gwinn, D.M.; Shackelford, D.B.; Egan, D.F.; Mihaylova, M.M.; Mery, A.; Vasquez, D.S.; Turk, B.E.; Shaw, R.J. AMPK phosphorylation of raptor mediates a metabolic checkpoint. Mol. Cell 2008, 30, 214–226. [Google Scholar] [CrossRef] [Green Version]
  178. Fang, Y.; Vilella-Bach, M.; Bachmann, R.; Flanigan, A.; Chen, J. Phosphatidic acid-mediated mitogenic activation of mTOR signaling. Science 2001, 294, 1942–1945. [Google Scholar] [CrossRef]
  179. Sarbassov, D.D.; Sabatini, D.M. Redox regulation of the nutrient-sensitive raptor-mTOR pathway and complex. J. Biol. Chem. 2005, 280, 39505–39509. [Google Scholar] [CrossRef] [Green Version]
  180. Proud, C.G. Signalling to translation: How signal transduction pathways control the protein synthetic machinery. Biochem. J. 2007, 403, 217–234. [Google Scholar] [CrossRef] [Green Version]
  181. Dann, S.G.; Thomas, G. The amino acid sensitive TOR pathway from yeast to mammals. FEBS Lett. 2006, 580, 2821–2829. [Google Scholar] [CrossRef] [Green Version]
  182. Feng, Z.; Zhang, H.; Levine, A.J.; Jin, S. The coordinate regulation of the p53 and mTOR pathways in cells. Proc. Natl. Acad. Sci. USA 2005, 102, 8204–8209. [Google Scholar] [CrossRef] [Green Version]
  183. Stambolic, V.; MacPherson, D.; Sas, D.; Lin, Y.; Snow, B.; Jang, Y.; Benchimol, S.; Mak, T.W. Regulation of PTEN transcription by p53. Mol. Cell 2001, 8, 317–325. [Google Scholar]
  184. Budanov, A.V.; Karin, M. p53 target genes sestrin1 and sestrin2 connect genotoxic stress and mTOR signaling. Cell 2008, 134, 451–460. [Google Scholar] [CrossRef] [Green Version]
  185. Duvel, K.; Yecies, J.L.; Menon, S.; Raman, P.; Lipovsky, A.I.; Souza, A.L.; Triantafellow, E.; Ma, Q.; Gorski, R.; Cleaver, S.; et al. Activation of a metabolic gene regulatory network downstream of mTOR complex 1. Mol. Cell 2010, 39, 171–183. [Google Scholar] [CrossRef] [Green Version]
  186. Hudson, C.C.; Liu, M.; Chiang, G.G.; Otterness, D.M.; Loomis, D.C.; Kaper, F.; Giaccia, A.J.; Abraham, R.T. Regulation of hypoxia-inducible factor 1alpha expression and function by the mammalian target of rapamycin. Mol. Cell Biol. 2002, 22, 7004–7014. [Google Scholar]
  187. Kim, J.E.; Chen, J. regulation of peroxisome proliferator-activated receptor-gamma activity by mammalian target of rapamycin and amino acids in adipogenesis. Diabetes 2004, 53, 2748–2756. [Google Scholar]
  188. Zhang, H.H.; Huang, J.; Duvel, K.; Boback, B.; Wu, S.; Squillace, R.M.; Wu, C.L.; Manning, B.D. Insulin stimulates adipogenesis through the Akt-TSC2-mTORC1 pathway. PLoS ONE 2009, 4, e6189. [Google Scholar] [CrossRef]
  189. Porstmann, T.; Santos, C.R.; Griffiths, B.; Cully, M.; Wu, M.; Leevers, S.; Griffiths, J.R.; Chung, Y.L.; Schulze, A. SREBP activity is regulated by mTORC1 and contributes to Akt-dependent cell growth. Cell Metab. 2008, 8, 224–236. [Google Scholar] [CrossRef] [Green Version]
  190. Wang, B.T.; Ducker, G.S.; Barczak, A.J.; Barbeau, R.; Erle, D.J.; Shokat, K.M. The mammalian target of rapamycin regulates cholesterol biosynthetic gene expression and exhibits a rapamycin-resistant transcriptional profile. Proc. Natl. Acad. Sci. USA 2011, 108, 15201–15206. [Google Scholar] [CrossRef] [Green Version]
  191. Yecies, J.L.; Zhang, H.H.; Menon, S.; Liu, S.; Yecies, D.; Lipovsky, A.I.; Gorgun, C.; Kwiatkowski, D.J.; Hotamisligil, G.S.; Lee, C.H.; et al. Akt stimulates hepatic SREBP1c and lipogenesis through parallel mTORC1-dependent and independent pathways. Cell Metab. 2011, 14, 21–32. [Google Scholar] [CrossRef] [Green Version]
  192. Li, S.; Brown, M.S.; Goldstein, J.L. Bifurcation of insulin signaling pathway in rat liver: mTORC1 required for stimulation of lipogenesis, but not inhibition of gluconeogenesis. Proc. Natl. Acad. Sci. USA 2010, 107, 3441–3446. [Google Scholar] [CrossRef] [Green Version]
  193. Cunningham, J.T.; Rodgers, J.T.; Arlow, D.H.; Vazquez, F.; Mootha, V.K.; Puigserver, P. mTOR controls mitochondrial oxidative function through a YY1-PGC-1alpha transcriptional complex. Nature 2007, 450, 736–740. [Google Scholar] [CrossRef]
  194. Corradetti, M.N.; Guan, K.L. Upstream of the mammalian target of rapamycin: Do all roads pass through mTOR? Oncogene 2006, 25, 6347–6360. [Google Scholar] [CrossRef] [Green Version]
  195. Hay, N.; Sonenberg, N. Upstream and downstream of mTOR. Genes Dev. 2004, 18, 1926–1945. [Google Scholar] [CrossRef] [Green Version]
  196. Shahbazian, D.; Roux, P.P.; Mieulet, V.; Cohen, M.S.; Raught, B.; Taunton, J.; Hershey, J.W.; Blenis, J.; Pende, M.; Sonenberg, N. The mTOR/PI3K and MAPK pathways converge on eIF4B to control its phosphorylation and activity. EMBO J. 2006, 25, 2781–2791. [Google Scholar] [CrossRef]
  197. Dorrello, N.V.; Peschiaroli, A.; Guardavaccaro, D.; Colburn, N.H.; Sherman, N.E.; Pagano, M. S6K1- and betaTRCP-mediated degradation of PDCD4 promotes protein translation and cell growth. Science 2006, 314, 467–471. [Google Scholar] [CrossRef]
  198. Richardson, C.J.; Broenstrup, M.; Fingar, D.C.; Julich, K.; Ballif, B.A.; Gygi, S.; Blenis, J. SKAR is a specific target of S6 kinase 1 in cell growth control. Curr. Biol. 2004, 14, 1540–1549. [Google Scholar] [CrossRef] [Green Version]
  199. Blenis, J.; Kuo, C.J.; Erikson, R.L. Identification of a ribosomal protein S6 kinase regulated by transformation and growth-promoting stimuli. J. Biol. Chem. 1987, 262, 14373–14376. [Google Scholar]
  200. Kim, J.; Guan, K.L. mTOR as a central hub of nutrient signalling and cell growth. Nat. Cell Biol. 2019, 21, 63–71. [Google Scholar] [CrossRef]
  201. Zhao, J.; Zhai, B.; Gygi, S.P.; Goldberg, A.L. mTOR inhibition activates overall protein degradation by the ubiquitin proteasome system as well as by autophagy. Proc. Natl. Acad. Sci. USA 2015, 112, 15790–15797. [Google Scholar] [CrossRef] [Green Version]
  202. Betz, C.; Stracka, D.; Prescianotto-Baschong, C.; Frieden, M.; Demaurex, N.; Hall, M.N. Feature Article: mTOR complex 2-Akt signaling at mitochondria-associated endoplasmic reticulum membranes (MAM) regulates mitochondrial physiology. Proc. Natl. Acad. Sci. USA 2013, 110, 12526–12534. [Google Scholar] [CrossRef] [Green Version]
  203. Sarbassov, D.D.; Guertin, D.A.; Ali, S.M.; Sabatini, D.M. Phosphorylation and regulation of Akt/PKB by the rictor-mTOR complex. Science 2005, 307, 1098–1101. [Google Scholar] [CrossRef] [Green Version]
  204. Garcia-Martinez, J.M.; Alessi, D.R. mTOR complex 2 (mTORC2) controls hydrophobic motif phosphorylation and activation of serum- and glucocorticoid-induced protein kinase 1 (SGK1). Biochem. J. 2008, 416, 375–385. [Google Scholar] [CrossRef] [Green Version]
  205. Brunet, A.; Park, J.; Tran, H.; Hu, L.S.; Hemmings, B.A.; Greenberg, M.E. Protein kinase SGK mediates survival signals by phosphorylating the forkhead transcription factor FKHRL1 (FOXO3a). Mol. Cell Biol. 2001, 21, 952–965. [Google Scholar] [CrossRef] [Green Version]
  206. Kaiser, G.; Gerst, F.; Michael, D.; Berchtold, S.; Friedrich, B.; Strutz-Seebohm, N.; Lang, F.; Haring, H.U.; Ullrich, S. Regulation of forkhead box O1 (FOXO1) by protein kinase B and glucocorticoids: Different mechanisms of induction of beta cell death in vitro. Diabetologia 2013, 56, 1587–1595. [Google Scholar] [CrossRef]
  207. Wang, X.; Hu, S.; Liu, L. Phosphorylation and acetylation modifications of FOXO3a: Independently or synergistically? Oncol. Lett. 2017, 13, 2867–2872. [Google Scholar] [CrossRef] [Green Version]
  208. Kaeberlein, M.; Powers, R.W., 3rd; Steffen, K.K.; Westman, E.A.; Hu, D.; Dang, N.; Kerr, E.O.; Kirkland, K.T.; Fields, S.; Kennedy, B.K. Regulation of yeast replicative life span by TOR and Sch9 in response to nutrients. Science 2005, 310, 1193–1196. [Google Scholar] [CrossRef] [Green Version]
  209. Vellai, T.; Takacs-Vellai, K.; Zhang, Y.; Kovacs, A.L.; Orosz, L.; Muller, F. Genetics: Influence of TOR kinase on lifespan in C. elegans. Nature 2003, 426, 620. [Google Scholar] [CrossRef]
  210. Kapahi, P.; Zid, B.M.; Harper, T.; Koslover, D.; Sapin, V.; Benzer, S. Regulation of lifespan in Drosophila by modulation of genes in the TOR signaling pathway. Curr. Biol. 2004, 14, 885–890. [Google Scholar] [CrossRef] [Green Version]
  211. Wu, J.J.; Liu, J.; Chen, E.B.; Wang, J.J.; Cao, L.; Narayan, N.; Fergusson, M.M.; Rovira, I.I.; Allen, M.; Springer, D.A.; et al. Increased mammalian lifespan and a segmental and tissue-specific slowing of aging after genetic reduction of mTOR expression. Cell Rep. 2013, 4, 913–920. [Google Scholar] [CrossRef] [Green Version]
  212. Medvedik, O.; Lamming, D.W.; Kim, K.D.; Sinclair, D.A. MSN2 and MSN4 link calorie restriction and TOR to sirtuin-mediated lifespan extension in Saccharomyces cerevisiae. PLoS Biol. 2007, 5, e261. [Google Scholar] [CrossRef] [Green Version]
  213. Bjedov, I.; Toivonen, J.M.; Kerr, F.; Slack, C.; Jacobson, J.; Foley, A.; Partridge, L. Mechanisms of life span extension by rapamycin in the fruit fly Drosophila melanogaster. Cell Metab. 2010, 11, 35–46. [Google Scholar] [CrossRef] [Green Version]
  214. Hansen, M.; Chandra, A.; Mitic, L.L.; Onken, B.; Driscoll, M.; Kenyon, C. A role for autophagy in the extension of lifespan by dietary restriction in C. elegans. PLoS Genet. 2008, 4, e24. [Google Scholar] [CrossRef] [Green Version]
  215. Toth, M.L.; Sigmond, T.; Borsos, E.; Barna, J.; Erdelyi, P.; Takacs-Vellai, K.; Orosz, L.; Kovacs, A.L.; Csikos, G.; Sass, M.; et al. Longevity pathways converge on autophagy genes to regulate life span in Caenorhabditis elegans. Autophagy 2008, 4, 330–338. [Google Scholar] [CrossRef] [Green Version]
  216. Hansen, M.; Taubert, S.; Crawford, D.; Libina, N.; Lee, S.J.; Kenyon, C. Lifespan extension by conditions that inhibit translation in Caenorhabditis elegans. Aging Cell 2007, 6, 95–110. [Google Scholar] [CrossRef] [Green Version]
  217. Blouet, C.; Ono, H.; Schwartz, G.J. Mediobasal hypothalamic p70 S6 kinase 1 modulates the control of energy homeostasis. Cell Metab. 2008, 8, 459–467. [Google Scholar] [CrossRef] [Green Version]
  218. Cota, D.; Matter, E.K.; Woods, S.C.; Seeley, R.J. The role of hypothalamic mammalian target of rapamycin complex 1 signaling in diet-induced obesity. J. Neurosci. 2008, 28, 7202–7208. [Google Scholar] [CrossRef] [Green Version]
  219. Sengupta, S.; Peterson, T.R.; Laplante, M.; Oh, S.; Sabatini, D.M. mTORC1 controls fasting-induced ketogenesis and its modulation by ageing. Nature 2010, 468, 1100–1104. [Google Scholar] [CrossRef]
  220. Kim, K.; Pyo, S.; Um, S.H. S6 kinase 2 deficiency enhances ketone body production and increases peroxisome proliferator-activated receptor alpha activity in the liver. Hepatology 2012, 55, 1727–1737. [Google Scholar] [CrossRef]
  221. Gebert, N.; Cheng, C.W.; Kirkpatrick, J.M.; Di Fraia, D.; Yun, J.; Schadel, P.; Pace, S.; Garside, G.B.; Werz, O.; Rudolph, K.L.; et al. Region-Specific Proteome Changes of the Intestinal Epithelium during Aging and Dietary Restriction. Cell Rep. 2020, 31, 107565. [Google Scholar] [CrossRef]
  222. Pentinmikko, N.; Iqbal, S.; Mana, M.; Andersson, S.; Cognetta, A.B., 3rd; Suciu, R.M.; Roper, J.; Luopajarvi, K.; Markelin, E.; Gopalakrishnan, S.; et al. Notum produced by Paneth cells attenuates regeneration of aged intestinal epithelium. Nature 2019, 571, 398–402. [Google Scholar] [CrossRef]
  223. Okuda, Y.; Kawai, K.; Yamashita, K. Age-related change in ketone body metabolism: Diminished glucagon effect on ketogenesis in adult rats. Endocrinology 1987, 120, 2152–2157. [Google Scholar] [CrossRef]
  224. Sanguino, E.; Ramon, M.; Michalik, L.; Wahli, W.; Alegret, M.; Sanchez, R.M.; Vazquez-Carrera, M.; Laguna, J.C. Lack of hypotriglyceridemic effect of gemfibrozil as a consequence of age-related changes in rat liver PPARalpha. Biochem. Pharmacol. 2004, 67, 157–166. [Google Scholar] [CrossRef]
  225. Sastre, J.; Pallardo, F.V.; Pla, R.; Pellin, A.; Juan, G.; O’Connor, J.E.; Estrela, J.M.; Miquel, J.; Vina, J. Aging of the liver: Age-associated mitochondrial damage in intact hepatocytes. Hepatology 1996, 24, 1199–1205. [Google Scholar] [CrossRef]
  226. Jiao, M.; Ren, F.; Zhou, L.; Zhang, X.; Zhang, L.; Wen, T.; Wei, L.; Wang, X.; Shi, H.; Bai, L.; et al. Peroxisome proliferator-activated receptor alpha activation attenuates the inflammatory response to protect the liver from acute failure by promoting the autophagy pathway. Cell Death Dis. 2014, 5, e1397. [Google Scholar] [CrossRef]
  227. Kim, Y.S.; Lee, H.M.; Kim, J.K.; Yang, C.S.; Kim, T.S.; Jung, M.; Jin, H.S.; Kim, S.; Jang, J.; Oh, G.T.; et al. PPAR-alpha Activation Mediates Innate Host Defense through Induction of TFEB and Lipid Catabolism. J. Immunol. 2017, 198, 3283–3295. [Google Scholar] [CrossRef] [Green Version]
  228. Pineda Torra, I.; Claudel, T.; Duval, C.; Kosykh, V.; Fruchart, J.C.; Staels, B. Bile acids induce the expression of the human peroxisome proliferator-activated receptor alpha gene via activation of the farnesoid X receptor. Mol. Endocrinol. 2003, 17, 259–272. [Google Scholar] [CrossRef] [Green Version]
  229. 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]
  230. Kim, K.H.; Moore, D.D. Regulation of Liver Energy Balance by the Nuclear Receptors Farnesoid X Receptor and Peroxisome Proliferator Activated Receptor alpha. Dig. Dis. 2017, 35, 203–209. [Google Scholar] [CrossRef]
  231. Sun, X.; Ritzenthaler, J.D.; Zhong, X.; Zheng, Y.; Roman, J.; Han, S. Nicotine stimulates PPARbeta/delta expression in human lung carcinoma cells through activation of PI3K/mTOR and suppression of AP-2alpha. Cancer Res. 2009, 69, 6445–6453. [Google Scholar] [CrossRef] [Green Version]
  232. Han, S.; Ritzenthaler, J.D.; Zheng, Y.; Roman, J. PPARbeta/delta agonist stimulates human lung carcinoma cell growth through inhibition of PTEN expression: The involvement of PI3K and NF-kappaB signals. Am. J. Physiol. Lung Cell Mol. Physiol. 2008, 294, L1238–L1249. [Google Scholar] [CrossRef]
  233. Foster, D.A. Phosphatidic acid and lipid-sensing by mTOR. Trends Endocrinol. Metab. 2013, 24, 272–278. [Google Scholar] [CrossRef] [Green Version]
  234. Cho, H.J.; Park, J.; Lee, H.W.; Lee, Y.S.; Kim, J.B. Regulation of adipocyte differentiation and insulin action with rapamycin. Biochem. Biophys. Res. Commun. 2004, 321, 942–948. [Google Scholar] [CrossRef]
  235. Gagnon, A.; Lau, S.; Sorisky, A. Rapamycin-sensitive phase of 3T3-L1 preadipocyte differentiation after clonal expansion. J. Cell Physiol. 2001, 189, 14–22. [Google Scholar] [CrossRef]
  236. Polak, P.; Cybulski, N.; Feige, J.N.; Auwerx, J.; Ruegg, M.A.; Hall, M.N. Adipose-specific knockout of raptor results in lean mice with enhanced mitochondrial respiration. Cell Metab. 2008, 8, 399–410. [Google Scholar] [CrossRef]
  237. Bell, A.; Grunder, L.; Sorisky, A. Rapamycin inhibits human adipocyte differentiation in primary culture. Obes. Res. 2000, 8, 249–254. [Google Scholar] [CrossRef]
  238. Carnevalli, L.S.; Masuda, K.; Frigerio, F.; Le Bacquer, O.; Um, S.H.; Gandin, V.; Topisirovic, I.; Sonenberg, N.; Thomas, G.; Kozma, S.C. S6K1 plays a critical role in early adipocyte differentiation. Dev. Cell 2010, 18, 763–774. [Google Scholar] [CrossRef] [Green Version]
  239. Blanchard, P.G.; Festuccia, W.T.; Houde, V.P.; St-Pierre, P.; Brule, S.; Turcotte, V.; Cote, M.; Bellmann, K.; Marette, A.; Deshaies, Y. Major involvement of mTOR in the PPARgamma-induced stimulation of adipose tissue lipid uptake and fat accretion. J. Lipid Res. 2012, 53, 1117–1125. [Google Scholar] [CrossRef] [Green Version]
  240. Kim, J.B.; Wright, H.M.; Wright, M.; Spiegelman, B.M. ADD1/SREBP1 activates PPARgamma through the production of endogenous ligand. Proc. Natl. Acad. Sci. USA 1998, 95, 4333–4337. [Google Scholar] [CrossRef] [Green Version]
  241. Bakan, I.; Laplante, M. Connecting mTORC1 signaling to SREBP-1 activation. Curr. Opin. Lipidol. 2012, 23, 226–234. [Google Scholar] [CrossRef]
  242. Huffman, T.A.; Mothe-Satney, I.; Lawrence, J.C., Jr. Insulin-stimulated phosphorylation of lipin mediated by the mammalian target of rapamycin. Proc. Natl. Acad. Sci. USA 2002, 99, 1047–1052. [Google Scholar] [CrossRef] [Green Version]
  243. Koh, Y.K.; Lee, M.Y.; Kim, J.W.; Kim, M.; Moon, J.S.; Lee, Y.J.; Ahn, Y.H.; Kim, K.S. Lipin1 is a key factor for the maturation and maintenance of adipocytes in the regulatory network with CCAAT/enhancer-binding protein alpha and peroxisome proliferator-activated receptor gamma 2. J. Biol. Chem. 2008, 283, 34896–34906. [Google Scholar] [CrossRef] [Green Version]
  244. Kim, H.E.; Bae, E.; Jeong, D.Y.; Kim, M.J.; Jin, W.J.; Park, S.W.; Han, G.S.; Carman, G.M.; Koh, E.; Kim, K.S. Lipin1 regulates PPARgamma transcriptional activity. Biochem. J. 2013, 453, 49–60. [Google Scholar] [CrossRef]
  245. Finck, B.N.; Gropler, M.C.; Chen, Z.; Leone, T.C.; Croce, M.A.; Harris, T.E.; Lawrence, J.C., Jr.; Kelly, D.P. Lipin 1 is an inducible amplifier of the hepatic PGC-1alpha/PPARalpha regulatory pathway. Cell Metab. 2006, 4, 199–210. [Google Scholar] [CrossRef] [Green Version]
  246. San, Y.Z.; Liu, Y.; Zhang, Y.; Shi, P.P.; Zhu, Y.L. Peroxisome proliferator-activated receptor-gamma agonist inhibits the mammalian target of rapamycin signaling pathway and has a protective effect in a rat model of status epilepticus. Mol. Med. Rep. 2015, 12, 1877–1883. [Google Scholar] [CrossRef] [Green Version]
  247. Purnell, P.; Teepper, C.; Liu, S.; Cardiff, R.; Gregg, J. Induction of PPAR gamma signaling and autophagy as a mechanism mediating acquired rapamycin resistance in breast cancer models. Cancer Res. 2008, 68. [Google Scholar]
  248. Assumpcao, J.A.F.; Magalhaes, K.G.; Correa, J.R. The role of ppargamma and autophagy in ros production, lipid droplets biogenesis and its involvement with colorectal cancer cells modulation. Cancer Cell Int. 2017, 17, 82. [Google Scholar] [CrossRef] [Green Version]
  249. Kang, J.H.; Chang, Y.C.; Maurizi, M.R. 4-O-carboxymethyl ascochlorin causes ER stress and induced autophagy in human hepatocellular carcinoma cells. J. Biol. Chem. 2012, 287, 15661–15671. [Google Scholar] [CrossRef] [Green Version]
  250. Pellerito, O.; Notaro, A.; Sabella, S.; De Blasio, A.; Vento, R.; Calvaruso, G.; Giuliano, M. WIN induces apoptotic cell death in human colon cancer cells through a block of autophagic flux dependent on PPARgamma down-regulation. Apoptosis 2014, 19, 1029–1042. [Google Scholar] [CrossRef]
  251. Weng, J.R.; Bai, L.Y.; Chiu, C.F.; Hu, J.L.; Chiu, S.J.; Wu, C.Y. Cucurbitane Triterpenoid from Momordica charantia Induces Apoptosis and Autophagy in Breast Cancer Cells, in Part, through Peroxisome Proliferator-Activated Receptor gamma Activation. Evid. Based Complement. Altern. Med. 2013, 2013, 935675. [Google Scholar] [CrossRef] [Green Version]
  252. Zhou, J.; Zhang, W.; Liang, B.; Casimiro, M.C.; Whitaker-Menezes, D.; Wang, M.; Lisanti, M.P.; Lanza-Jacoby, S.; Pestell, R.G.; Wang, C. PPARgamma activation induces autophagy in breast cancer cells. Int. J. Biochem. Cell Biol. 2009, 41, 2334–2342. [Google Scholar] [CrossRef] [Green Version]
  253. Yan, S.; Yang, X.; Chen, T.; Xi, Z.; Jiang, X. The PPARgamma agonist Troglitazone induces autophagy, apoptosis and necroptosis in bladder cancer cells. Cancer Gene Ther. 2014, 21, 188–193. [Google Scholar] [CrossRef] [Green Version]
  254. Cerquetti, L.; Sampaoli, C.; Amendola, D.; Bucci, B.; Masuelli, L.; Marchese, R.; Misiti, S.; De Venanzi, A.; Poggi, M.; Toscano, V.; et al. Rosiglitazone induces autophagy in H295R and cell cycle deregulation in SW13 adrenocortical cancer cells. Exp. Cell Res. 2011, 317, 1397–1410. [Google Scholar] [CrossRef]
  255. Teresi, R.E.; Waite, K.A. PPARgamma, PTEN, and the Fight against Cancer. PPAR Res. 2008, 2008, 932632. [Google Scholar] [CrossRef] [Green Version]
  256. Patel, L.; Pass, I.; Coxon, P.; Downes, C.P.; Smith, S.A.; Macphee, C.H. Tumor suppressor and anti-inflammatory actions of PPARgamma agonists are mediated via upregulation of PTEN. Curr. Biol. 2001, 11, 764–768. [Google Scholar]
  257. Zhang, W.; Wu, N.; Li, Z.; Wang, L.; Jin, J.; Zha, X.L. PPARgamma activator rosiglitazone inhibits cell migration via upregulation of PTEN in human hepatocarcinoma cell line BEL-7404. Cancer Biol. Ther. 2006, 5, 1008–1014. [Google Scholar]
  258. Lin, C.F.; Young, K.C.; Bai, C.H.; Yu, B.C.; Ma, C.T.; Chien, Y.C.; Chiang, C.L.; Liao, C.S.; Lai, H.W.; Tsao, C.W. Rosiglitazone regulates anti-inflammation and growth inhibition via PTEN. Biomed. Res. Int. 2014, 2014, 787924. [Google Scholar] [CrossRef]
  259. Song, M.S.; Salmena, L.; Pandolfi, P.P. The functions and regulation of the PTEN tumour suppressor. Nat. Rev. Mol. Cell Biol. 2012, 13, 283–296. [Google Scholar] [CrossRef]
  260. Panasyuk, G.; Espeillac, C.; Chauvin, C.; Pradelli, L.A.; Horie, Y.; Suzuki, A.; Annicotte, J.S.; Fajas, L.; Foretz, M.; Verdeguer, F.; et al. PPARgamma contributes to PKM2 and HK2 expression in fatty liver. Nat. Commun. 2012, 3, 672. [Google Scholar] [CrossRef] [Green Version]
  261. Wang, L.; Yin, Y.; Hou, G.; Kang, J.; Wang, Q. Peroxisome Proliferator-Activated Receptor (PPARgamma) Plays a Protective Role in Cigarette Smoking-Induced Inflammation via AMP-Activated Protein Kinase (AMPK) Signaling. Med. Sci. Monit. 2018, 24, 5168–5177. [Google Scholar] [CrossRef]
  262. He, G.; Sung, Y.M.; Digiovanni, J.; Fischer, S.M. Thiazolidinediones inhibit insulin-like growth factor-i-induced activation of p70S6 kinase and suppress insulin-like growth factor-I tumor-promoting activity. Cancer Res. 2006, 66, 1873–1878. [Google Scholar] [CrossRef] [Green Version]
  263. Lee, Y.H.; Lee, H.Y.; Kim, T.G.; Lee, N.H.; Yu, M.K.; Yi, H.K. PPARgamma Maintains Homeostasis through Autophagy Regulation in Dental Pulp. J. Dent. Res. 2015, 94, 729–737. [Google Scholar] [CrossRef]
  264. Xu, F.; Li, J.; Ni, W.; Shen, Y.W.; Zhang, X.P. Peroxisome proliferator-activated receptor-gamma agonist 15d-prostaglandin J2 mediates neuronal autophagy after cerebral ischemia-reperfusion injury. PLoS ONE 2013, 8, e55080. [Google Scholar] [CrossRef] [Green Version]
  265. Li, H.; Zhang, Q.; Yang, X.; Wang, L. PPAR-gamma agonist rosiglitazone reduces autophagy and promotes functional recovery in experimental traumaticspinal cord injury. Neurosci. Lett. 2017, 650, 89–96. [Google Scholar] [CrossRef]
  266. Mahmood, D.F.; Jguirim-Souissi, I.; Khadija, E.H.; Blondeau, N.; Diderot, V.; Amrani, S.; Slimane, M.N.; Syrovets, T.; Simmet, T.; Rouis, M. Peroxisome proliferator-activated receptor gamma induces apoptosis and inhibits autophagy of human monocyte-derived macrophages via induction of cathepsin L: Potential role in atherosclerosis. J. Biol. Chem. 2011, 286, 28858–28866. [Google Scholar] [CrossRef] [Green Version]
  267. Xiao, B.; Sanders, M.J.; Underwood, E.; Heath, R.; Mayer, F.V.; Carmena, D.; Jing, C.; Walker, P.A.; Eccleston, J.F.; Haire, L.F.; et al. Structure of mammalian AMPK and its regulation by ADP. Nature 2011, 472, 230–233. [Google Scholar] [CrossRef] [Green Version]
  268. Ross, F.A.; MacKintosh, C.; Hardie, D.G. AMP-activated protein kinase: A cellular energy sensor that comes in 12 flavours. FEBS J. 2016, 283, 2987–3001. [Google Scholar] [CrossRef]
  269. Suter, M.; Riek, U.; Tuerk, R.; Schlattner, U.; Wallimann, T.; Neumann, D. Dissecting the role of 5′-AMP for allosteric stimulation, activation, and deactivation of AMP-activated protein kinase. J. Biol. Chem. 2006, 281, 32207–32216. [Google Scholar] [CrossRef] [Green Version]
  270. Davies, S.P.; Helps, N.R.; Cohen, P.T.; Hardie, D.G. 5′-AMP inhibits dephosphorylation, as well as promoting phosphorylation, of the AMP-activated protein kinase. Studies using bacterially expressed human protein phosphatase-2C alpha and native bovine protein phosphatase-2AC. FEBS Lett. 1995, 377, 421–425. [Google Scholar] [CrossRef] [Green Version]
  271. Fogarty, S.; Hawley, S.A.; Green, K.A.; Saner, N.; Mustard, K.J.; Hardie, D.G. Calmodulin-dependent protein kinase kinase-beta activates AMPK without forming a stable complex: Synergistic effects of Ca2+ and AMP. Biochem. J. 2010, 426, 109–118. [Google Scholar] [CrossRef]
  272. Winder, W.W.; Hardie, D.G. AMP-activated protein kinase, a metabolic master switch: Possible roles in type 2 diabetes. Am. J. Physiol. 1999, 277, E1–E10. [Google Scholar] [CrossRef]
  273. Koo, S.H.; Flechner, L.; Qi, L.; Zhang, X.; Screaton, R.A.; Jeffries, S.; Hedrick, S.; Xu, W.; Boussouar, F.; Brindle, P.; et al. The CREB coactivator TORC2 is a key regulator of fasting glucose metabolism. Nature 2005, 437, 1109–1111. [Google Scholar] [CrossRef]
  274. Chen, S.; Murphy, J.; Toth, R.; Campbell, D.G.; Morrice, N.A.; Mackintosh, C. Complementary regulation of TBC1D1 and AS160 by growth factors, insulin and AMPK activators. Biochem. J. 2008, 409, 449–459. [Google Scholar] [CrossRef] [Green Version]
  275. Geraghty, K.M.; Chen, S.; Harthill, J.E.; Ibrahim, A.F.; Toth, R.; Morrice, N.A.; Vandermoere, F.; Moorhead, G.B.; Hardie, D.G.; MacKintosh, C. Regulation of multisite phosphorylation and 14-3-3 binding of AS160 in response to IGF-1, EGF, PMA and AICAR. Biochem. J. 2007, 407, 231–241. [Google Scholar] [CrossRef] [Green Version]
  276. Jager, S.; Handschin, C.; St-Pierre, J.; Spiegelman, B.M. AMP-activated protein kinase (AMPK) action in skeletal muscle via direct phosphorylation of PGC-1alpha. Proc. Natl. Acad. Sci. USA 2007, 104, 12017–12022. [Google Scholar] [CrossRef] [Green Version]
  277. McGee, S.L.; van Denderen, B.J.; Howlett, K.F.; Mollica, J.; Schertzer, J.D.; Kemp, B.E.; Hargreaves, M. AMP-activated protein kinase regulates GLUT4 transcription by phosphorylating histone deacetylase 5. Diabetes 2008, 57, 860–867. [Google Scholar] [CrossRef]
  278. Davies, S.P.; Sim, A.T.; Hardie, D.G. Location and function of three sites phosphorylated on rat acetyl-CoA carboxylase by the AMP-activated protein kinase. Eur. J. Biochem. 1990, 187, 183–190. [Google Scholar]
  279. Davies, S.P.; Carling, D.; Munday, M.R.; Hardie, D.G. Diurnal rhythm of phosphorylation of rat liver acetyl-CoA carboxylase by the AMP-activated protein kinase, demonstrated using freeze-clamping. Effects of high fat diets. Eur. J. Biochem. 1992, 203, 615–623. [Google Scholar]
  280. Li, Y.; Xu, S.; Mihaylova, M.M.; Zheng, B.; Hou, X.; Jiang, B.; Park, O.; Luo, Z.; Lefai, E.; Shyy, J.Y.; et al. AMPK phosphorylates and inhibits SREBP activity to attenuate hepatic steatosis and atherosclerosis in diet-induced insulin-resistant mice. Cell Metab. 2011, 13, 376–388. [Google Scholar] [CrossRef] [Green Version]
  281. Muoio, D.M.; Seefeld, K.; Witters, L.A.; Coleman, R.A. AMP-activated kinase reciprocally regulates triacylglycerol synthesis and fatty acid oxidation in liver and muscle: Evidence that sn-glycerol-3-phosphate acyltransferase is a novel target. Biochem. J. 1999, 338 Pt 3, 783–791. [Google Scholar]
  282. Clarke, P.R.; Hardie, D.G. Regulation of HMG-CoA reductase: Identification of the site phosphorylated by the AMP-activated protein kinase in vitro and in intact rat liver. EMBO J. 1990, 9, 2439–2446. [Google Scholar]
  283. Hoppe, S.; Bierhoff, H.; Cado, I.; Weber, A.; Tiebe, M.; Grummt, I.; Voit, R. AMP-activated protein kinase adapts rRNA synthesis to cellular energy supply. Proc. Natl. Acad. Sci. USA 2009, 106, 17781–17786. [Google Scholar] [CrossRef] [Green Version]
  284. Dufer, M.; Noack, K.; Krippeit-Drews, P.; Drews, G. Activation of the AMP-activated protein kinase enhances glucose-stimulated insulin secretion in mouse beta-cells. Islets 2010, 2, 156–163. [Google Scholar] [CrossRef] [Green Version]
  285. Chavez, J.A.; Roach, W.G.; Keller, S.R.; Lane, W.S.; Lienhard, G.E. Inhibition of GLUT4 translocation by Tbc1d1, a Rab GTPase-activating protein abundant in skeletal muscle, is partially relieved by AMP-activated protein kinase activation. J. Biol. Chem. 2008, 283, 9187–9195. [Google Scholar] [CrossRef] [Green Version]
  286. Marsin, A.S.; Bouzin, C.; Bertrand, L.; Hue, L. The stimulation of glycolysis by hypoxia in activated monocytes is mediated by AMP-activated protein kinase and inducible 6-phosphofructo-2-kinase. J. Biol. Chem. 2002, 277, 30778–30783. [Google Scholar] [CrossRef] [Green Version]
  287. Jorgensen, S.B.; Nielsen, J.N.; Birk, J.B.; Olsen, G.S.; Viollet, B.; Andreelli, F.; Schjerling, P.; Vaulont, S.; Hardie, D.G.; Hansen, B.F.; et al. The alpha2-5′ AMP-activated protein kinase is a site 2 glycogen synthase kinase in skeletal muscle and is responsive to glucose loading. Diabetes 2004, 53, 3074–3081. [Google Scholar]
  288. Hardie, D.G.; Ross, F.A.; Hawley, S.A. AMPK: A nutrient and energy sensor that maintains energy homeostasis. Nat. Rev. Mol. Cell Biol. 2012, 13, 251–262. [Google Scholar] [CrossRef] [Green Version]
  289. Leclerc, I.; Lenzner, C.; Gourdon, L.; Vaulont, S.; Kahn, A.; Viollet, B. Hepatocyte nuclear factor-4alpha involved in type 1 maturity-onset diabetes of the young is a novel target of AMP-activated protein kinase. Diabetes 2001, 50, 1515–1521. [Google Scholar] [CrossRef] [Green Version]
  290. Lee, J.M.; Seo, W.Y.; Song, K.H.; Chanda, D.; Kim, Y.D.; Kim, D.K.; Lee, M.W.; Ryu, D.; Kim, Y.H.; Noh, J.R.; et al. AMPK-dependent repression of hepatic gluconeogenesis via disruption of CREB.CRTC2 complex by orphan nuclear receptor small heterodimer partner. J. Biol. Chem. 2010, 285, 32182–32191. [Google Scholar] [CrossRef] [Green Version]
  291. Minokoshi, Y.; Alquier, T.; Furukawa, N.; Kim, Y.B.; Lee, A.; Xue, B.; Mu, J.; Foufelle, F.; Ferre, P.; Birnbaum, M.J.; et al. AMP-kinase regulates food intake by responding to hormonal and nutrient signals in the hypothalamus. Nature 2004, 428, 569–574. [Google Scholar] [CrossRef]
  292. Andersson, U.; Filipsson, K.; Abbott, C.R.; Woods, A.; Smith, K.; Bloom, S.R.; Carling, D.; Small, C.J. AMP-activated protein kinase plays a role in the control of food intake. J. Biol. Chem. 2004, 279, 12005–12008. [Google Scholar] [CrossRef] [Green Version]
  293. Steinberg, G.R.; Kemp, B.E. AMPK in Health and Disease. Physiol. Rev. 2009, 89, 1025–1078. [Google Scholar] [CrossRef]
  294. Wang, Z.; Wang, N.; Liu, P.; Xie, X. AMPK and Cancer. Exp. Suppl. 2016, 107, 203–226. [Google Scholar] [CrossRef]
  295. Umezawa, S.; Higurashi, T.; Nakajima, A. AMPK: Therapeutic Target for Diabetes and Cancer Prevention. Curr. Pharm. Des. 2017, 23, 3629–3644. [Google Scholar] [CrossRef]
  296. Greer, E.L.; Dowlatshahi, D.; Banko, M.R.; Villen, J.; Hoang, K.; Blanchard, D.; Gygi, S.P.; Brunet, A. An AMPK-FOXO pathway mediates longevity induced by a novel method of dietary restriction in C. elegans. Curr. Biol. 2007, 17, 1646–1656. [Google Scholar] [CrossRef] [Green Version]
  297. Schulz, T.J.; Zarse, K.; Voigt, A.; Urban, N.; Birringer, M.; Ristow, M. Glucose restriction extends Caenorhabditis elegans life span by inducing mitochondrial respiration and increasing oxidative stress. Cell Metab. 2007, 6, 280–293. [Google Scholar] [CrossRef] [Green Version]
  298. Funakoshi, M.; Tsuda, M.; Muramatsu, K.; Hatsuda, H.; Morishita, S.; Aigaki, T. A gain-of-function screen identifies wdb and lkb1 as lifespan-extending genes in Drosophila. Biochem. Biophys. Res. Commun. 2011, 405, 667–672. [Google Scholar] [CrossRef]
  299. Stenesen, D.; Suh, J.M.; Seo, J.; Yu, K.; Lee, K.S.; Kim, J.S.; Min, K.J.; Graff, J.M. Adenosine nucleotide biosynthesis and AMPK regulate adult life span and mediate the longevity benefit of caloric restriction in flies. Cell Metab. 2013, 17, 101–112. [Google Scholar] [CrossRef] [Green Version]
  300. Hardie, D.G.; Carling, D.; Carlson, M. The AMP-activated/SNF1 protein kinase subfamily: Metabolic sensors of the eukaryotic cell? Annu. Rev. Biochem. 1998, 67, 821–855. [Google Scholar] [CrossRef]
  301. Viollet, B.; Andreelli, F. AMP-activated protein kinase and metabolic control. Handb. Exp. Pharmacol. 2011, 303–330. [Google Scholar] [CrossRef] [Green Version]
  302. Burns, K.A.; Vanden Heuvel, J.P. Modulation of PPAR activity via phosphorylation. Biochim. Biophys. Acta 2007, 1771, 952–960. [Google Scholar] [CrossRef] [Green Version]
  303. Li, J.; Miller, E.J.; Ninomiya-Tsuji, J.; Russell, R.R., 3rd; Young, L.H. AMP-activated protein kinase activates p38 mitogen-activated protein kinase by increasing recruitment of p38 MAPK to TAB1 in the ischemic heart. Circ. Res. 2005, 97, 872–879. [Google Scholar] [CrossRef] [Green Version]
  304. Lemieux, K.; Konrad, D.; Klip, A.; Marette, A. The AMP-activated protein kinase activator AICAR does not induce GLUT4 translocation to transverse tubules but stimulates glucose uptake and p38 mitogen-activated protein kinases alpha and beta in skeletal muscle. FASEB J. 2003, 17, 1658–1665. [Google Scholar] [CrossRef]
  305. Hunter, R.W.; Treebak, J.T.; Wojtaszewski, J.F.; Sakamoto, K. Molecular mechanism by which AMP-activated protein kinase activation promotes glycogen accumulation in muscle. Diabetes 2011, 60, 766–774. [Google Scholar] [CrossRef] [Green Version]
  306. Hinds, T.D., Jr.; Burns, K.A.; Hosick, P.A.; McBeth, L.; Nestor-Kalinoski, A.; Drummond, H.A.; AlAmodi, A.A.; Hankins, M.W.; Vanden Heuvel, J.P.; Stec, D.E. Biliverdin Reductase A Attenuates Hepatic Steatosis by Inhibition of Glycogen Synthase Kinase (GSK) 3beta Phosphorylation of Serine 73 of Peroxisome Proliferator-activated Receptor (PPAR) alpha. J. Biol. Chem. 2016, 291, 25179–25191. [Google Scholar] [CrossRef] [Green Version]
  307. Lee, W.J.; Kim, M.; Park, H.S.; Kim, H.S.; Jeon, M.J.; Oh, K.S.; Koh, E.H.; Won, J.C.; Kim, M.S.; Oh, G.T.; et al. AMPK activation increases fatty acid oxidation in skeletal muscle by activating PPARalpha and PGC-1. Biochem. Biophys. Res. Commun. 2006, 340, 291–295. [Google Scholar] [CrossRef]
  308. Yoon, M.J.; Lee, G.Y.; Chung, J.J.; Ahn, Y.H.; Hong, S.H.; Kim, J.B. Adiponectin increases fatty acid oxidation in skeletal muscle cells by sequential activation of AMP-activated protein kinase, p38 mitogen-activated protein kinase, and peroxisome proliferator-activated receptor alpha. Diabetes 2006, 55, 2562–2570. [Google Scholar] [CrossRef] [Green Version]
  309. Hotta, K.; Funahashi, T.; Arita, Y.; Takahashi, M.; Matsuda, M.; Okamoto, Y.; Iwahashi, H.; Kuriyama, H.; Ouchi, N.; Maeda, K.; et al. Plasma concentrations of a novel, adipose-specific protein, adiponectin, in type 2 diabetic patients. Arterioscler. Thromb. Vasc. Biol. 2000, 20, 1595–1599. [Google Scholar] [CrossRef] [Green Version]
  310. Gan, Z.; Burkart-Hartman, E.M.; Han, D.H.; Finck, B.; Leone, T.C.; Smith, E.Y.; Ayala, J.E.; Holloszy, J.; Kelly, D.P. The nuclear receptor PPARbeta/delta programs muscle glucose metabolism in cooperation with AMPK and MEF2. Genes Dev. 2011, 25, 2619–2630. [Google Scholar] [CrossRef] [Green Version]
  311. Bai, F.; Liu, Y.; Tu, T.; Li, B.; Xiao, Y.; Ma, Y.; Qin, F.; Xie, J.; Zhou, S.; Liu, Q. Metformin regulates lipid metabolism in a canine model of atrial fibrillation through AMPK/PPAR-alpha/VLCAD pathway. Lipids Health Dis. 2019, 18, 109. [Google Scholar] [CrossRef] [Green Version]
  312. Joly, E.; Roduit, R.; Peyot, M.L.; Habinowski, S.A.; Ruderman, N.B.; Witters, L.A.; Prentki, M. Glucose represses PPARalpha gene expression via AMP-activated protein kinase but not via p38 mitogen-activated protein kinase in the pancreatic beta-cell. J. Diabetes 2009, 1, 263–272. [Google Scholar] [CrossRef]
  313. Ravnskjaer, K.; Boergesen, M.; Dalgaard, L.T.; Mandrup, S. Glucose-induced repression of PPARalpha gene expression in pancreatic beta-cells involves PP2A activation and AMPK inactivation. J. Mol. Endocrinol. 2006, 36, 289–299. [Google Scholar] [CrossRef] [Green Version]
  314. Bronner, M.; Hertz, R.; Bar-Tana, J. Kinase-independent transcriptional co-activation of peroxisome proliferator-activated receptor alpha by AMP-activated protein kinase. Biochem. J. 2004, 384, 295–305. [Google Scholar] [CrossRef] [Green Version]
  315. Sozio, M.S.; Lu, C.; Zeng, Y.; Liangpunsakul, S.; Crabb, D.W. Activated AMPK inhibits PPAR-{alpha} and PPAR-{gamma} transcriptional activity in hepatoma cells. Am. J. Physiol. Gastrointest. Liver Physiol. 2011, 301, G739–G747. [Google Scholar] [CrossRef] [Green Version]
  316. Chanda, D.; Lee, C.H.; Kim, Y.H.; Noh, J.R.; Kim, D.K.; Park, J.H.; Hwang, J.H.; Lee, M.R.; Jeong, K.H.; Lee, I.K.; et al. Fenofibrate differentially regulates plasminogen activator inhibitor-1 gene expression via adenosine monophosphate-activated protein kinase-dependent induction of orphan nuclear receptor small heterodimer partner. Hepatology 2009, 50, 880–892. [Google Scholar] [CrossRef] [Green Version]
  317. Liangpunsakul, S.; Wou, S.E.; Wineinger, K.D.; Zeng, Y.; Cyganek, I.; Jayaram, H.N.; Crabb, D.W. Effects of WY-14,643 on the phosphorylation and activation of AMP-dependent protein kinase. Arch. Biochem. Biophys. 2009, 485, 10–15. [Google Scholar] [CrossRef] [Green Version]
  318. Ren, G.; Rimando, A.M.; Mathews, S.T. AMPK activation by pterostilbene contributes to suppression of hepatic gluconeogenic gene expression and glucose production in H4IIE cells. Biochem. Biophys. Res. Commun. 2018, 498, 640–645. [Google Scholar] [CrossRef]
  319. Murakami, H.; Murakami, R.; Kambe, F.; Cao, X.; Takahashi, R.; Asai, T.; Hirai, T.; Numaguchi, Y.; Okumura, K.; Seo, H.; et al. Fenofibrate activates AMPK and increases eNOS phosphorylation in HUVEC. Biochem. Biophys. Res. Commun. 2006, 341, 973–978. [Google Scholar] [CrossRef]
  320. Okayasu, T.; Tomizawa, A.; Suzuki, K.; Manaka, K.; Hattori, Y. PPARalpha activators upregulate eNOS activity and inhibit cytokine-induced NF-kappaB activation through AMP-activated protein kinase activation. Life Sci. 2008, 82, 884–891. [Google Scholar] [CrossRef]
  321. Manio, M.C.; Inoue, K.; Fujitani, M.; Matsumura, S.; Fushiki, T. Combined pharmacological activation of AMPK and PPARdelta potentiates the effects of exercise in trained mice. Physiol. Rep. 2016, 4. [Google Scholar] [CrossRef] [Green Version]
  322. Narkar, V.A.; Downes, M.; Yu, R.T.; Embler, E.; Wang, Y.X.; Banayo, E.; Mihaylova, M.M.; Nelson, M.C.; Zou, Y.; Juguilon, H.; et al. AMPK and PPARdelta agonists are exercise mimetics. Cell 2008, 134, 405–415. [Google Scholar] [CrossRef] [Green Version]
  323. Lamontagne, J.; Pepin, E.; Peyot, M.L.; Joly, E.; Ruderman, N.B.; Poitout, V.; Madiraju, S.R.; Nolan, C.J.; Prentki, M. Pioglitazone acutely reduces insulin secretion and causes metabolic deceleration of the pancreatic beta-cell at submaximal glucose concentrations. Endocrinology 2009, 150, 3465–3474. [Google Scholar] [CrossRef] [Green Version]
  324. Boyle, J.G.; Logan, P.J.; Ewart, M.A.; Reihill, J.A.; Ritchie, S.A.; Connell, J.M.; Cleland, S.J.; Salt, I.P. Rosiglitazone stimulates nitric oxide synthesis in human aortic endothelial cells via AMP-activated protein kinase. J. Biol. Chem. 2008, 283, 11210–11217. [Google Scholar] [CrossRef] [Green Version]
  325. Ceolotto, G.; Gallo, A.; Papparella, I.; Franco, L.; Murphy, E.; Iori, E.; Pagnin, E.; Fadini, G.P.; Albiero, M.; Semplicini, A.; et al. Rosiglitazone reduces glucose-induced oxidative stress mediated by NAD(P)H oxidase via AMPK-dependent mechanism. Arterioscler. Thromb. Vasc. Biol. 2007, 27, 2627–2633. [Google Scholar] [CrossRef] [Green Version]
  326. Sun, H.; Zhu, X.; Lin, W.; Zhou, Y.; Cai, W.; Qiu, L. Interactions of TLR4 and PPARgamma, Dependent on AMPK Signalling Pathway Contribute to Anti-Inflammatory Effects of Vaccariae Hypaphorine in Endothelial Cells. Cell Physiol. Biochem. 2017, 42, 1227–1239. [Google Scholar] [CrossRef] [Green Version]
  327. Wang, X.; Zhou, L.; Shao, L.; Qian, L.; Fu, X.; Li, G.; Luo, T.; Gu, Y.; Li, F.; Li, J.; et al. Troglitazone acutely activates AMP-activated protein kinase and inhibits insulin secretion from beta cells. Life Sci. 2007, 81, 160–165. [Google Scholar] [CrossRef]
  328. LeBrasseur, N.K.; Kelly, M.; Tsao, T.S.; Farmer, S.R.; Saha, A.K.; Ruderman, N.B.; Tomas, E. Thiazolidinediones can rapidly activate AMP-activated protein kinase in mammalian tissues. Am. J. Physiol. Endocrinol. Metab. 2006, 291, E175–E181. [Google Scholar] [CrossRef] [Green Version]
  329. Shen, Z.; Liang, X.; Rogers, C.Q.; Rideout, D.; You, M. Involvement of adiponectin-SIRT1-AMPK signaling in the protective action of rosiglitazone against alcoholic fatty liver in mice. Am. J. Physiol. Gastrointest. Liver Physiol. 2010, 298, G364–G374. [Google Scholar] [CrossRef] [Green Version]
  330. Saha, A.K.; Avilucea, P.R.; Ye, J.M.; Assifi, M.M.; Kraegen, E.W.; Ruderman, N.B. Pioglitazone treatment activates AMP-activated protein kinase in rat liver and adipose tissue in vivo. Biochem. Biophys. Res. Commun. 2004, 314, 580–585. [Google Scholar] [CrossRef]
  331. Ye, J.M.; Dzamko, N.; Hoy, A.J.; Iglesias, M.A.; Kemp, B.; Kraegen, E. Rosiglitazone treatment enhances acute AMP-activated protein kinase-mediated muscle and adipose tissue glucose uptake in high-fat-fed rats. Diabetes 2006, 55, 2797–2804. [Google Scholar] [CrossRef] [Green Version]
  332. Coletta, D.K.; Sriwijitkamol, A.; Wajcberg, E.; Tantiwong, P.; Li, M.; Prentki, M.; Madiraju, M.; Jenkinson, C.P.; Cersosimo, E.; Musi, N.; et al. Pioglitazone stimulates AMP-activated protein kinase signalling and increases the expression of genes involved in adiponectin signalling, mitochondrial function and fat oxidation in human skeletal muscle in vivo: A randomised trial. Diabetologia 2009, 52, 723–732. [Google Scholar] [CrossRef] [Green Version]
  333. Xiao, X.; Su, G.; Brown, S.N.; Chen, L.; Ren, J.; Zhao, P. Peroxisome proliferator-activated receptors gamma and alpha agonists stimulate cardiac glucose uptake via activation of AMP-activated protein kinase. J. Nutr. Biochem. 2010, 21, 621–626. [Google Scholar] [CrossRef]
  334. Ratziu, V.; Charlotte, F.; Bernhardt, C.; Giral, P.; Halbron, M.; Lenaour, G.; Hartmann-Heurtier, A.; Bruckert, E.; Poynard, T.; Group, L.S. Long-term efficacy of rosiglitazone in nonalcoholic steatohepatitis: Results of the fatty liver improvement by rosiglitazone therapy (FLIRT 2) extension trial. Hepatology 2010, 51, 445–453. [Google Scholar] [CrossRef]
  335. Zhang, F.; Dey, D.; Branstrom, R.; Forsberg, L.; Lu, M.; Zhang, Q.; Sjoholm, A. BLX-1002, a novel thiazolidinedione with no PPAR affinity, stimulates AMP-activated protein kinase activity, raises cytosolic Ca2+, and enhances glucose-stimulated insulin secretion in a PI3K-dependent manner. Am. J. Physiol. Cell Physiol. 2009, 296, C346–C354. [Google Scholar] [CrossRef]
  336. Chang, T.J.; Chen, W.P.; Yang, C.; Lu, P.H.; Liang, Y.C.; Su, M.J.; Lee, S.C.; Chuang, L.M. Serine-385 phosphorylation of inwardly rectifying K+ channel subunit (Kir6.2) by AMP-dependent protein kinase plays a key role in rosiglitazone-induced closure of the K(ATP) channel and insulin secretion in rats. Diabetologia 2009, 52, 1112–1121. [Google Scholar] [CrossRef] [Green Version]
  337. Fryer, L.G.; Parbu-Patel, A.; Carling, D. The Anti-diabetic drugs rosiglitazone and metformin stimulate AMP-activated protein kinase through distinct signaling pathways. J. Biol. Chem. 2002, 277, 25226–25232. [Google Scholar] [CrossRef] [Green Version]
  338. Gonzalez, E.; McGraw, T.E. The Akt kinases: Isoform specificity in metabolism and cancer. Cell Cycle 2009, 8, 2502–2508. [Google Scholar] [CrossRef]
  339. Oh, W.J.; Jacinto, E. mTOR complex 2 signaling and functions. Cell Cycle 2011, 10, 2305–2316. [Google Scholar] [CrossRef]
  340. Piscitello, D.; Varshney, D.; Lilla, S.; Vizioli, M.G.; Reid, C.; Gorbunova, V.; Seluanov, A.; Gillespie, D.A.; Adams, P.D. AKT overactivation can suppress DNA repair via p70S6 kinase-dependent downregulation of MRE11. Oncogene 2018, 37, 427–438. [Google Scholar] [CrossRef] [Green Version]
  341. Jia, Y.; Song, W.; Zhang, F.; Yan, J.; Yang, Q. Akt1 inhibits homologous recombination in Brca1-deficient cells by blocking the Chk1-Rad51 pathway. Oncogene 2013, 32, 1943–1949. [Google Scholar] [CrossRef] [Green Version]
  342. Liu, P.; Gan, W.; Guo, C.; Xie, A.; Gao, D.; Guo, J.; Zhang, J.; Willis, N.; Su, A.; Asara, J.M.; et al. Akt-mediated phosphorylation of XLF impairs non-homologous end-joining DNA repair. Mol. Cell 2015, 57, 648–661. [Google Scholar] [CrossRef] [Green Version]
  343. Haeusler, R.A.; McGraw, T.E.; Accili, D. Biochemical and cellular properties of insulin receptor signalling. Nat. Rev. Mol. Cell Biol. 2018, 19, 31–44. [Google Scholar] [CrossRef]
  344. Rowland, A.F.; Fazakerley, D.J.; James, D.E. Mapping insulin/GLUT4 circuitry. Traffic 2011, 12, 672–681. [Google Scholar] [CrossRef]
  345. Bevan, P. Insulin signalling. J. Cell Sci. 2001, 114, 1429–1430. [Google Scholar]
  346. Siddle, K. Signalling by insulin and IGF receptors: Supporting acts and new players. J. Mol. Endocrinol. 2011, 47, R1–R10. [Google Scholar] [CrossRef] [Green Version]
  347. Wong, R.H.; Sul, H.S. Insulin signaling in fatty acid and fat synthesis: A transcriptional perspective. Curr. Opin. Pharmacol. 2010, 10, 684–691. [Google Scholar] [CrossRef] [Green Version]
  348. Weindruch, R. The retardation of aging by caloric restriction: Studies in rodents and primates. Toxicol. Pathol. 1996, 24, 742–745. [Google Scholar] [CrossRef]
  349. Shimokawa, I.; Komatsu, T.; Hayashi, N.; Kim, S.E.; Kawata, T.; Park, S.; Hayashi, H.; Yamaza, H.; Chiba, T.; Mori, R. The life-extending effect of dietary restriction requires Foxo3 in mice. Aging Cell 2015, 14, 707–709. [Google Scholar] [CrossRef]
  350. Kim, D.H.; Park, M.H.; Lee, E.K.; Choi, Y.J.; Chung, K.W.; Moon, K.M.; Kim, M.J.; An, H.J.; Park, J.W.; Kim, N.D.; et al. The roles of FoxOs in modulation of aging by calorie restriction. Biogerontology 2015, 16, 1–14. [Google Scholar] [CrossRef]
  351. Bluher, M.; Kahn, B.B.; Kahn, C.R. Extended longevity in mice lacking the insulin receptor in adipose tissue. Science 2003, 299, 572–574. [Google Scholar] [CrossRef] [Green Version]
  352. Dosch, J.; Meissner, U.; Rascher, W. Prolonged lifespan by defective insulin signalling? Eur. J. Endocrinol. 2003, 148, 489–490. [Google Scholar] [CrossRef] [Green Version]
  353. Tatar, M.; Bartke, A.; Antebi, A. The endocrine regulation of aging by insulin-like signals. Science 2003, 299, 1346–1351. [Google Scholar] [CrossRef]
  354. Kenyon, C.; Chang, J.; Gensch, E.; Rudner, A.; Tabtiang, R. A C. elegans mutant that lives twice as long as wild type. Nature 1993, 366, 461–464. [Google Scholar] [CrossRef]
  355. Holzenberger, M.; Dupont, J.; Ducos, B.; Leneuve, P.; Geloen, A.; Even, P.C.; Cervera, P.; Le Bouc, Y. IGF-1 receptor regulates lifespan and resistance to oxidative stress in mice. Nature 2003, 421, 182–187. [Google Scholar] [CrossRef]
  356. Taguchi, A.; Wartschow, L.M.; White, M.F. Brain IRS2 signaling coordinates life span and nutrient homeostasis. Science 2007, 317, 369–372. [Google Scholar] [CrossRef] [Green Version]
  357. Selman, C.; Lingard, S.; Choudhury, A.I.; Batterham, R.L.; Claret, M.; Clements, M.; Ramadani, F.; Okkenhaug, K.; Schuster, E.; Blanc, E.; et al. Evidence for lifespan extension and delayed age-related biomarkers in insulin receptor substrate 1 null mice. FASEB J. 2008, 22, 807–818. [Google Scholar] [CrossRef] [Green Version]
  358. Coschigano, K.T.; Holland, A.N.; Riders, M.E.; List, E.O.; Flyvbjerg, A.; Kopchick, J.J. Deletion, but not antagonism, of the mouse growth hormone receptor results in severely decreased body weights, insulin, and insulin-like growth factor I levels and increased life span. Endocrinology 2003, 144, 3799–3810. [Google Scholar] [CrossRef] [Green Version]
  359. Bonkowski, M.S.; Rocha, J.S.; Masternak, M.M.; Al Regaiey, K.A.; Bartke, A. Targeted disruption of growth hormone receptor interferes with the beneficial actions of calorie restriction. Proc. Natl. Acad. Sci. USA 2006, 103, 7901–7905. [Google Scholar] [CrossRef] [Green Version]
  360. Kurosu, H.; Yamamoto, M.; Clark, J.D.; Pastor, J.V.; Nandi, A.; Gurnani, P.; McGuinness, O.P.; Chikuda, H.; Yamaguchi, M.; Kawaguchi, H.; et al. Suppression of aging in mice by the hormone Klotho. Science 2005, 309, 1829–1833. [Google Scholar] [CrossRef] [Green Version]
  361. Erol, A. Insulin resistance is an evolutionarily conserved physiological mechanism at the cellular level for protection against increased oxidative stress. Bioessays 2007, 29, 811–818. [Google Scholar] [CrossRef]
  362. Dobbins, R.L.; Chester, M.W.; Stevenson, B.E.; Daniels, M.B.; Stein, D.T.; McGarry, J.D. A fatty acid-dependent step is critically important for both glucose- and non-glucose-stimulated insulin secretion. J. Clin. Investig. 1998, 101, 2370–2376. [Google Scholar] [CrossRef]
  363. Gremlich, S.; Nolan, C.; Roduit, R.; Burcelin, R.; Peyot, M.L.; Delghingaro-Augusto, V.; Desvergne, B.; Michalik, L.; Prentki, M.; Wahli, W. Pancreatic islet adaptation to fasting is dependent on peroxisome proliferator-activated receptor alpha transcriptional up-regulation of fatty acid oxidation. Endocrinology 2005, 146, 375–382. [Google Scholar] [CrossRef]
  364. Ravnskjaer, K.; Boergesen, M.; Rubi, B.; Larsen, J.K.; Nielsen, T.; Fridriksson, J.; Maechler, P.; Mandrup, S. Peroxisome proliferator-activated receptor alpha (PPARalpha) potentiates, whereas PPARgamma attenuates, glucose-stimulated insulin secretion in pancreatic beta-cells. Endocrinology 2005, 146, 3266–3276. [Google Scholar] [CrossRef] [Green Version]
  365. Holness, M.J.; Smith, N.D.; Greenwood, G.K.; Sugden, M.C. Acute (24 h) activation of peroxisome proliferator-activated receptor-alpha (PPARalpha) reverses high-fat feeding-induced insulin hypersecretion in vivo and in perifused pancreatic islets. J. Endocrinol. 2003, 177, 197–205. [Google Scholar] [CrossRef] [Green Version]
  366. Lalloyer, F.; Vandewalle, B.; Percevault, F.; Torpier, G.; Kerr-Conte, J.; Oosterveer, M.; Paumelle, R.; Fruchart, J.C.; Kuipers, F.; Pattou, F.; et al. Peroxisome proliferator-activated receptor alpha improves pancreatic adaptation to insulin resistance in obese mice and reduces lipotoxicity in human islets. Diabetes 2006, 55, 1605–1613. [Google Scholar] [CrossRef] [Green Version]
  367. Maida, A.; Lamont, B.J.; Cao, X.; Drucker, D.J. Metformin regulates the incretin receptor axis via a pathway dependent on peroxisome proliferator-activated receptor-alpha in mice. Diabetologia 2011, 54, 339–349. [Google Scholar] [CrossRef]
  368. Patsouris, D.; Mandard, S.; Voshol, P.J.; Escher, P.; Tan, N.S.; Havekes, L.M.; Koenig, W.; Marz, W.; Tafuri, S.; Wahli, W.; et al. PPARalpha governs glycerol metabolism. J. Clin. Investig. 2004, 114, 94–103. [Google Scholar] [CrossRef]
  369. Wu, P.; Peters, J.M.; Harris, R.A. Adaptive increase in pyruvate dehydrogenase kinase 4 during starvation is mediated by peroxisome proliferator-activated receptor alpha. Biochem. Biophys. Res. Commun. 2001, 287, 391–396. [Google Scholar] [CrossRef]
  370. Sugden, M.C.; Greenwood, G.K.; Smith, N.D.; Holness, M.J. Peroxisome proliferator-activated receptor-alpha activation during pregnancy attenuates glucose-stimulated insulin hypersecretion in vivo by increasing insulin sensitivity, without impairing pregnancy-induced increases in beta-cell glucose sensing and responsiveness. Endocrinology 2003, 144, 146–153. [Google Scholar] [CrossRef] [Green Version]
  371. Ferrari, C.; Frezzati, S.; Romussi, M.; Bertazzoni, A.; Testori, G.P.; Antonini, S.; Paracchi, A. Effects of short-term clofibrate administration on glucose tolerance and insulin secretion in patients with chemical diabetes or hypertriglyceridemia. Metabolism 1977, 26, 129–139. [Google Scholar] [CrossRef]
  372. Murakami, K.; Nambu, S.; Koh, H.; Kobayashi, M.; Shigeta, Y. Clofibrate enhances the affinity of insulin receptors in non-insulin dependent diabetes mellitus. Br. J. Clin. Pharmacol. 1984, 17, 89–91. [Google Scholar] [CrossRef] [Green Version]
  373. Kobayashi, M.; Shigeta, Y.; Hirata, Y.; Omori, Y.; Sakamoto, N.; Nambu, S.; Baba, S. Improvement of glucose tolerance in NIDDM by clofibrate. Randomized double-blind study. Diabetes Care 1988, 11, 495–499. [Google Scholar] [CrossRef]
  374. Koh, E.H.; Kim, M.S.; Park, J.Y.; Kim, H.S.; Youn, J.Y.; Park, H.S.; Youn, J.H.; Lee, K.U. Peroxisome proliferator-activated receptor (PPAR)-alpha activation prevents diabetes in OLETF rats: Comparison with PPAR-gamma activation. Diabetes 2003, 52, 2331–2337. [Google Scholar] [CrossRef] [Green Version]
  375. Guerre-Millo, M.; Gervois, P.; Raspe, E.; Madsen, L.; Poulain, P.; Derudas, B.; Herbert, J.M.; Winegar, D.A.; Willson, T.M.; Fruchart, J.C.; et al. Peroxisome proliferator-activated receptor alpha activators improve insulin sensitivity and reduce adiposity. J. Biol. Chem. 2000, 275, 16638–16642. [Google Scholar] [CrossRef] [Green Version]
  376. Aasum, E.; Belke, D.D.; Severson, D.L.; Riemersma, R.A.; Cooper, M.; Andreassen, M.; Larsen, T.S. Cardiac function and metabolism in Type 2 diabetic mice after treatment with BM 17.0744, a novel PPAR-alpha activator. Am. J. Physiol. Heart Circ. Physiol. 2002, 283, H949–H957. [Google Scholar] [CrossRef]
  377. Park, C.W.; Zhang, Y.; Zhang, X.; Wu, J.; Chen, L.; Cha, D.R.; Su, D.; Hwang, M.T.; Fan, X.; Davis, L.; et al. PPARalpha agonist fenofibrate improves diabetic nephropathy in db/db mice. Kidney Int. 2006, 69, 1511–1517. [Google Scholar] [CrossRef] [Green Version]
  378. Sugden, M.C.; Bulmer, K.; Gibbons, G.F.; Knight, B.L.; Holness, M.J. Peroxisome-proliferator-activated receptor-alpha (PPARalpha) deficiency leads to dysregulation of hepatic lipid and carbohydrate metabolism by fatty acids and insulin. Biochem. J. 2002, 364, 361–368. [Google Scholar] [CrossRef] [Green Version]
  379. Guerre-Millo, M.; Rouault, C.; Poulain, P.; Andre, J.; Poitout, V.; Peters, J.M.; Gonzalez, F.J.; Fruchart, J.C.; Reach, G.; Staels, B. PPAR-alpha-null mice are protected from high-fat diet-induced insulin resistance. Diabetes 2001, 50, 2809–2814. [Google Scholar] [CrossRef] [Green Version]
  380. Jo, Y.S.; Ryu, D.; Maida, A.; Wang, X.; Evans, R.M.; Schoonjans, K.; Auwerx, J. Phosphorylation of the nuclear receptor corepressor 1 by protein kinase B switches its corepressor targets in the liver in mice. Hepatology 2015, 62, 1606–1618. [Google Scholar] [CrossRef] [Green Version]
  381. Roduit, R.; Morin, J.; Masse, F.; Segall, L.; Roche, E.; Newgard, C.B.; Assimacopoulos-Jeannet, F.; Prentki, M. Glucose down-regulates the expression of the peroxisome proliferator-activated receptor-alpha gene in the pancreatic beta-cell. J. Biol. Chem. 2000, 275, 35799–35806. [Google Scholar] [CrossRef] [Green Version]
  382. Blanquart, C.; Mansouri, R.; Paumelle, R.; Fruchart, J.C.; Staels, B.; Glineur, C. The protein kinase C signaling pathway regulates a molecular switch between transactivation and transrepression activity of the peroxisome proliferator-activated receptor alpha. Mol. Endocrinol. 2004, 18, 1906–1918. [Google Scholar] [CrossRef] [Green Version]
  383. Hostetler, H.A.; Huang, H.; Kier, A.B.; Schroeder, F. Glucose directly links to lipid metabolism through high affinity interaction with peroxisome proliferator-activated receptor alpha. J. Biol. Chem. 2008, 283, 2246–2254. [Google Scholar] [CrossRef] [Green Version]
  384. Cohen, G.; Riahi, Y.; Shamni, O.; Guichardant, M.; Chatgilialoglu, C.; Ferreri, C.; Kaiser, N.; Sasson, S. Role of lipid peroxidation and PPAR-delta in amplifying glucose-stimulated insulin secretion. Diabetes 2011, 60, 2830–2842. [Google Scholar] [CrossRef] [Green Version]
  385. Iglesias, J.; Barg, S.; Vallois, D.; Lahiri, S.; Roger, C.; Yessoufou, A.; Pradevand, S.; McDonald, A.; Bonal, C.; Reimann, F.; et al. PPARbeta/delta affects pancreatic beta cell mass and insulin secretion in mice. J. Clin. Investig. 2012, 122, 4105–4117. [Google Scholar] [CrossRef] [Green Version]
  386. Serrano-Marco, L.; Rodriguez-Calvo, R.; El Kochairi, I.; Palomer, X.; Michalik, L.; Wahli, W.; Vazquez-Carrera, M. Activation of peroxisome proliferator-activated receptor-beta/-delta (PPAR-beta/-delta) ameliorates insulin signaling and reduces SOCS3 levels by inhibiting STAT3 in interleukin-6-stimulated adipocytes. Diabetes 2011, 60, 1990–1999. [Google Scholar] [CrossRef] [Green Version]
  387. Serrano-Marco, L.; Barroso, E.; El Kochairi, I.; Palomer, X.; Michalik, L.; Wahli, W.; Vazquez-Carrera, M. The peroxisome proliferator-activated receptor (PPAR) beta/delta agonist GW501516 inhibits IL-6-induced signal transducer and activator of transcription 3 (STAT3) activation and insulin resistance in human liver cells. Diabetologia 2012, 55, 743–751. [Google Scholar] [CrossRef] [Green Version]
  388. Salvado, L.; Barroso, E.; Gomez-Foix, A.M.; Palomer, X.; Michalik, L.; Wahli, W.; Vazquez-Carrera, M. PPARbeta/delta prevents endoplasmic reticulum stress-associated inflammation and insulin resistance in skeletal muscle cells through an AMPK-dependent mechanism. Diabetologia 2014, 57, 2126–2135. [Google Scholar] [CrossRef]
  389. Lee, C.H.; Olson, P.; Hevener, A.; Mehl, I.; Chong, L.W.; Olefsky, J.M.; Gonzalez, F.J.; Ham, J.; Kang, H.; Peters, J.M.; et al. PPARdelta regulates glucose metabolism and insulin sensitivity. Proc. Natl. Acad. Sci. USA 2006, 103, 3444–3449. [Google Scholar] [CrossRef] [Green Version]
  390. Oliver, W.R., Jr.; Shenk, J.L.; Snaith, M.R.; Russell, C.S.; Plunket, K.D.; Bodkin, N.L.; Lewis, M.C.; Winegar, D.A.; Sznaidman, M.L.; Lambert, M.H.; et al. A selective peroxisome proliferator-activated receptor delta agonist promotes reverse cholesterol transport. Proc. Natl. Acad. Sci. USA 2001, 98, 5306–5311. [Google Scholar] [CrossRef] [Green Version]
  391. Tanaka, T.; Yamamoto, J.; Iwasaki, S.; Asaba, H.; Hamura, H.; Ikeda, Y.; Watanabe, M.; Magoori, K.; Ioka, R.X.; Tachibana, K.; et al. Activation of peroxisome proliferator-activated receptor delta induces fatty acid beta-oxidation in skeletal muscle and attenuates metabolic syndrome. Proc. Natl. Acad. Sci. USA 2003, 100, 15924–15929. [Google Scholar] [CrossRef] [Green Version]
  392. Riserus, U.; Sprecher, D.; Johnson, T.; Olson, E.; Hirschberg, S.; Liu, A.; Fang, Z.; Hegde, P.; Richards, D.; Sarov-Blat, L.; et al. Activation of peroxisome proliferator-activated receptor (PPAR)delta promotes reversal of multiple metabolic abnormalities, reduces oxidative stress, and increases fatty acid oxidation in moderately obese men. Diabetes 2008, 57, 332–339. [Google Scholar] [CrossRef] [Green Version]
  393. Cariou, B.; Hanf, R.; Lambert-Porcheron, S.; Zair, Y.; Sauvinet, V.; Noel, B.; Flet, L.; Vidal, H.; Staels, B.; Laville, M. Dual peroxisome proliferator-activated receptor alpha/delta agonist GFT505 improves hepatic and peripheral insulin sensitivity in abdominally obese subjects. Diabetes Care 2013, 36, 2923–2930. [Google Scholar] [CrossRef] [Green Version]
  394. Buchanan, T.A.; Xiang, A.H.; Peters, R.K.; Kjos, S.L.; Marroquin, A.; Goico, J.; Ochoa, C.; Tan, S.; Berkowitz, K.; Hodis, H.N.; et al. Preservation of pancreatic beta-cell function and prevention of type 2 diabetes by pharmacological treatment of insulin resistance in high-risk hispanic women. Diabetes 2002, 51, 2796–2803. [Google Scholar] [CrossRef] [Green Version]
  395. Knowler, W.C.; Hamman, R.F.; Edelstein, S.L.; Barrett-Connor, E.; Ehrmann, D.A.; Walker, E.A.; Fowler, S.E.; Nathan, D.M.; Kahn, S.E.; Diabetes Prevention Program Research, G. Prevention of type 2 diabetes with troglitazone in the Diabetes Prevention Program. Diabetes 2005, 54, 1150–1156. [Google Scholar] [CrossRef]
  396. Moitra, J.; Mason, M.M.; Olive, M.; Krylov, D.; Gavrilova, O.; Marcus-Samuels, B.; Feigenbaum, L.; Lee, E.; Aoyama, T.; Eckhaus, M.; et al. Life without white fat: A transgenic mouse. Genes Dev. 1998, 12, 3168–3181. [Google Scholar]
  397. Larsen, T.M.; Toubro, S.; Astrup, A. PPARgamma agonists in the treatment of type II diabetes: Is increased fatness commensurate with long-term efficacy? Int. J. Obes. Relat. Metab. Disord. 2003, 27, 147–161. [Google Scholar] [CrossRef] [Green Version]
  398. Ye, J.M.; Doyle, P.J.; Iglesias, M.A.; Watson, D.G.; Cooney, G.J.; Kraegen, E.W. Peroxisome proliferator-activated receptor (PPAR)-alpha activation lowers muscle lipids and improves insulin sensitivity in high fat-fed rats: Comparison with PPAR-gamma activation. Diabetes 2001, 50, 411–417. [Google Scholar] [CrossRef] [Green Version]
  399. Collino, M.; Aragno, M.; Castiglia, S.; Miglio, G.; Tomasinelli, C.; Boccuzzi, G.; Thiemermann, C.; Fantozzi, R. Pioglitazone improves lipid and insulin levels in overweight rats on a high cholesterol and fructose diet by decreasing hepatic inflammation. Br. J. Pharmacol. 2010, 160, 1892–1902. [Google Scholar] [CrossRef] [Green Version]
  400. Iwata, M.; Haruta, T.; Usui, I.; Takata, Y.; Takano, A.; Uno, T.; Kawahara, J.; Ueno, E.; Sasaoka, T.; Ishibashi, O.; et al. Pioglitazone ameliorates tumor necrosis factor-alpha-induced insulin resistance by a mechanism independent of adipogenic activity of peroxisome proliferator-activated receptor-gamma. Diabetes 2001, 50, 1083–1092. [Google Scholar] [CrossRef] [Green Version]
  401. Smith, U.; Gogg, S.; Johansson, A.; Olausson, T.; Rotter, V.; Svalstedt, B. Thiazolidinediones (PPARgamma agonists) but not PPARalpha agonists increase IRS-2 gene expression in 3T3-L1 and human adipocytes. FASEB J. 2001, 15, 215–220. [Google Scholar] [CrossRef]
  402. Rieusset, J.; Auwerx, J.; Vidal, H. Regulation of gene expression by activation of the peroxisome proliferator-activated receptor gamma with rosiglitazone (BRL 49653) in human adipocytes. Biochem. Biophys. Res. Commun. 1999, 265, 265–271. [Google Scholar] [CrossRef]
  403. Baumann, C.A.; Chokshi, N.; Saltiel, A.R.; Ribon, V. Cloning and characterization of a functional peroxisome proliferator activator receptor-gamma-responsive element in the promoter of the CAP gene. J. Biol. Chem. 2000, 275, 9131–9135. [Google Scholar] [CrossRef] [Green Version]
  404. Ribon, V.; Johnson, J.H.; Camp, H.S.; Saltiel, A.R. Thiazolidinediones and insulin resistance: Peroxisome proliferatoractivated receptor gamma activation stimulates expression of the CAP gene. Proc. Natl. Acad. Sci. USA 1998, 95, 14751–14756. [Google Scholar] [CrossRef]
  405. Liu, J.; DeYoung, S.M.; Hwang, J.B.; O’Leary, E.E.; Saltiel, A.R. The roles of Cbl-b and c-Cbl in insulin-stimulated glucose transport. J. Biol. Chem. 2003, 278, 36754–36762. [Google Scholar] [CrossRef] [Green Version]
  406. Kramer, D.; Shapiro, R.; Adler, A.; Bush, E.; Rondinone, C.M. Insulin-sensitizing effect of rosiglitazone (BRL-49653) by regulation of glucose transporters in muscle and fat of Zucker rats. Metabolism 2001, 50, 1294–1300. [Google Scholar] [CrossRef]
  407. Standaert, M.L.; Kanoh, Y.; Sajan, M.P.; Bandyopadhyay, G.; Farese, R.V. Cbl, IRS-1, and IRS-2 mediate effects of rosiglitazone on PI3K, PKC-lambda, and glucose transport in 3T3/L1 adipocytes. Endocrinology 2002, 143, 1705–1716. [Google Scholar] [CrossRef]
  408. Kim, H.I.; Kim, J.W.; Kim, S.H.; Cha, J.Y.; Kim, K.S.; Ahn, Y.H. Identification and functional characterization of the peroxisomal proliferator response element in rat GLUT2 promoter. Diabetes 2000, 49, 1517–1524. [Google Scholar] [CrossRef] [Green Version]
  409. Samaha, F.F.; Szapary, P.O.; Iqbal, N.; Williams, M.M.; Bloedon, L.T.; Kochar, A.; Wolfe, M.L.; Rader, D.J. Effects of rosiglitazone on lipids, adipokines, and inflammatory markers in nondiabetic patients with low high-density lipoprotein cholesterol and metabolic syndrome. Arterioscler. Thromb. Vasc. Biol. 2006, 26, 624–630. [Google Scholar] [CrossRef] [Green Version]
  410. Rui, L.; Yuan, M.; Frantz, D.; Shoelson, S.; White, M.F. SOCS-1 and SOCS-3 block insulin signaling by ubiquitin-mediated degradation of IRS1 and IRS2. J. Biol. Chem. 2002, 277, 42394–42398. [Google Scholar] [CrossRef] [Green Version]
  411. Farrell, G.C. Signalling links in the liver: Knitting SOCS with fat and inflammation. J. Hepatol. 2005, 43, 193–196. [Google Scholar] [CrossRef]
  412. Szalkowski, D.; White-Carrington, S.; Berger, J.; Zhang, B. Antidiabetic thiazolidinediones block the inhibitory effect of tumor necrosis factor-alpha on differentiation, insulin-stimulated glucose uptake, and gene expression in 3T3-L1 cells. Endocrinology 1995, 136, 1474–1481. [Google Scholar] [CrossRef]
  413. Pajvani, U.B.; Hawkins, M.; Combs, T.P.; Rajala, M.W.; Doebber, T.; Berger, J.P.; Wagner, J.A.; Wu, M.; Knopps, A.; Xiang, A.H.; et al. Complex distribution, not absolute amount of adiponectin, correlates with thiazolidinedione-mediated improvement in insulin sensitivity. J. Biol. Chem. 2004, 279, 12152–12162. [Google Scholar] [CrossRef] [Green Version]
  414. Yamauchi, T.; Kamon, J.; Waki, H.; Terauchi, Y.; Kubota, N.; Hara, K.; Mori, Y.; Ide, T.; Murakami, K.; Tsuboyama-Kasaoka, N.; et al. The fat-derived hormone adiponectin reverses insulin resistance associated with both lipoatrophy and obesity. Nat. Med. 2001, 7, 941–946. [Google Scholar] [CrossRef]
  415. Armoni, M.; Harel, C.; Karni, S.; Chen, H.; Bar-Yoseph, F.; Ver, M.R.; Quon, M.J.; Karnieli, E. FOXO1 represses peroxisome proliferator-activated receptor-gamma1 and -gamma2 gene promoters in primary adipocytes. A novel paradigm to increase insulin sensitivity. J. Biol. Chem. 2006, 281, 19881–19891. [Google Scholar] [CrossRef] [Green Version]
  416. Fan, W.; Imamura, T.; Sonoda, N.; Sears, D.D.; Patsouris, D.; Kim, J.J.; Olefsky, J.M. FOXO1 transrepresses peroxisome proliferator-activated receptor gamma transactivation, coordinating an insulin-induced feed-forward response in adipocytes. J. Biol. Chem. 2009, 284, 12188–12197. [Google Scholar] [CrossRef] [Green Version]
  417. Dowell, P.; Otto, T.C.; Adi, S.; Lane, M.D. Convergence of peroxisome proliferator-activated receptor gamma and Foxo1 signaling pathways. J. Biol. Chem. 2003, 278, 45485–45491. [Google Scholar] [CrossRef] [Green Version]
  418. Nolan, J.J.; Ludvik, B.; Beerdsen, P.; Joyce, M.; Olefsky, J. Improvement in glucose tolerance and insulin resistance in obese subjects treated with troglitazone. N. Engl. J. Med. 1994, 331, 1188–1193. [Google Scholar] [CrossRef]
  419. Barroso, I.; Gurnell, M.; Crowley, V.E.; Agostini, M.; Schwabe, J.W.; Soos, M.A.; Maslen, G.L.; Williams, T.D.; Lewis, H.; Schafer, A.J.; et al. Dominant negative mutations in human PPARgamma associated with severe insulin resistance, diabetes mellitus and hypertension. Nature 1999, 402, 880–883. [Google Scholar] [CrossRef]
  420. Savage, D.B.; Tan, G.D.; Acerini, C.L.; Jebb, S.A.; Agostini, M.; Gurnell, M.; Williams, R.L.; Umpleby, A.M.; Thomas, E.L.; Bell, J.D.; et al. Human metabolic syndrome resulting from dominant-negative mutations in the nuclear receptor peroxisome proliferator-activated receptor-gamma. Diabetes 2003, 52, 910–917. [Google Scholar] [CrossRef] [Green Version]
  421. Agarwal, A.K.; Garg, A. A novel heterozygous mutation in peroxisome proliferator-activated receptor-gamma gene in a patient with familial partial lipodystrophy. J. Clin. Endocrinol. Metab. 2002, 87, 408–411. [Google Scholar] [CrossRef]
  422. Hegele, R.A.; Cao, H.; Frankowski, C.; Mathews, S.T.; Leff, T. PPARG F388L, a transactivation-deficient mutant, in familial partial lipodystrophy. Diabetes 2002, 51, 3586–3590. [Google Scholar] [CrossRef] [Green Version]
  423. Demir, T.; Onay, H.; Savage, D.B.; Temeloglu, E.; Uzum, A.K.; Kadioglu, P.; Altay, C.; Ozen, S.; Demir, L.; Cavdar, U.; et al. Familial partial lipodystrophy linked to a novel peroxisome proliferator activator receptor -gamma (PPARG) mutation, H449L: A comparison of people with this mutation and those with classic codon 482 Lamin A/C (LMNA) mutations. Diabet Med. 2016, 33, 1445–1450. [Google Scholar] [CrossRef] [Green Version]
  424. Ludtke, A.; Buettner, J.; Schmidt, H.H.; Worman, H.J. New PPARG mutation leads to lipodystrophy and loss of protein function that is partially restored by a synthetic ligand. J. Med. Genet. 2007, 44, e88. [Google Scholar] [CrossRef] [Green Version]
  425. Ludtke, A.; Buettner, J.; Wu, W.; Muchir, A.; Schroeter, A.; Zinn-Justin, S.; Spuler, S.; Schmidt, H.H.; Worman, H.J. Peroxisome proliferator-activated receptor-gamma C190S mutation causes partial lipodystrophy. J. Clin. Endocrinol. Metab. 2007, 92, 2248–2255. [Google Scholar] [CrossRef] [Green Version]
  426. Reitman, M.L.; Arioglu, E.; Gavrilova, O.; Taylor, S.I. Lipoatrophy revisited. Trends Endocrinol. Metab. 2000, 11, 410–416. [Google Scholar]
  427. Jones, J.R.; Barrick, C.; Kim, K.A.; Lindner, J.; Blondeau, B.; Fujimoto, Y.; Shiota, M.; Kesterson, R.A.; Kahn, B.B.; Magnuson, M.A. Deletion of PPARgamma in adipose tissues of mice protects against high fat diet-induced obesity and insulin resistance. Proc. Natl. Acad. Sci. USA 2005, 102, 6207–6212. [Google Scholar] [CrossRef] [Green Version]
  428. Corton, J.C.; Brown-Borg, H.M. Peroxisome proliferator-activated receptor gamma coactivator 1 in caloric restriction and other models of longevity. J. Gerontol. Biol. Sci. Med. Sci. 2005, 60, 1494–1509. [Google Scholar] [CrossRef]
  429. Carvalho, C.R.; Brenelli, S.L.; Silva, A.C.; Nunes, A.L.; Velloso, L.A.; Saad, M.J. Effect of aging on insulin receptor, insulin receptor substrate-1, and phosphatidylinositol 3-kinase in liver and muscle of rats. Endocrinology 1996, 137, 151–159. [Google Scholar] [CrossRef] [Green Version]
  430. Rocchi, S.; Picard, F.; Vamecq, J.; Gelman, L.; Potier, N.; Zeyer, D.; Dubuquoy, L.; Bac, P.; Champy, M.F.; Plunket, K.D.; et al. A unique PPARgamma ligand with potent insulin-sensitizing yet weak adipogenic activity. Mol. Cell 2001, 8, 737–747. [Google Scholar]
  431. Dunn, F.L.; Higgins, L.S.; Fredrickson, J.; De Paoli, A.M.; group, I.N.T.S. Selective modulation of PPARgamma activity can lower plasma glucose without typical thiazolidinedione side-effects in patients with Type 2 diabetes. J. Diabetes Complicat. 2011, 25, 151–158. [Google Scholar] [CrossRef]
  432. Beekmann, K.; Rubio, L.; de Haan, L.H.; Actis-Goretta, L.; van der Burg, B.; van Bladeren, P.J.; Rietjens, I.M. The effect of quercetin and kaempferol aglycones and glucuronides on peroxisome proliferator-activated receptor-gamma (PPAR-gamma). Food Funct. 2015, 6, 1098–1107. [Google Scholar] [CrossRef]
  433. Fang, X.K.; Gao, J.; Zhu, D.N. Kaempferol and quercetin isolated from Euonymus alatus improve glucose uptake of 3T3-L1 cells without adipogenesis activity. Life Sci. 2008, 82, 615–622. [Google Scholar] [CrossRef]
  434. Takahashi, H.; Hara, H.; Goto, T.; Kamakari, K.; Wataru, N.; Mohri, S.; Takahashi, N.; Suzuki, H.; Shibata, D.; Kawada, T. 13-Oxo-9(Z),11(E),15(Z)-octadecatrienoic acid activates peroxisome proliferator-activated receptor gamma in adipocytes. Lipids 2015, 50, 3–12. [Google Scholar] [CrossRef]
  435. Zhang, Y.; Gu, M.; Cai, W.; Yu, L.; Feng, L.; Zhang, L.; Zang, Q.; Wang, Y.; Wang, D.; Chen, H.; et al. Dietary component isorhamnetin is a PPARgamma antagonist and ameliorates metabolic disorders induced by diet or leptin deficiency. Sci. Rep. 2016, 6, 19288. [Google Scholar] [CrossRef]
  436. Yamauchi, T.; Waki, H.; Kamon, J.; Murakami, K.; Motojima, K.; Komeda, K.; Miki, H.; Kubota, N.; Terauchi, Y.; Tsuchida, A.; et al. Inhibition of RXR and PPARgamma ameliorates diet-induced obesity and type 2 diabetes. J. Clin. Investig. 2001, 108, 1001–1013. [Google Scholar] [CrossRef]
  437. Kubota, N.; Terauchi, Y.; Miki, H.; Tamemoto, H.; Yamauchi, T.; Komeda, K.; Satoh, S.; Nakano, R.; Ishii, C.; Sugiyama, T.; et al. PPAR gamma mediates high-fat diet-induced adipocyte hypertrophy and insulin resistance. Mol. Cell 1999, 4, 597–609. [Google Scholar] [CrossRef]
  438. Miles, P.D.; Barak, Y.; He, W.; Evans, R.M.; Olefsky, J.M. Improved insulin-sensitivity in mice heterozygous for PPAR-gamma deficiency. J. Clin. Investig. 2000, 105, 287–292. [Google Scholar] [CrossRef] [Green Version]
  439. Yamauchi, T.; Kamon, J.; Waki, H.; Murakami, K.; Motojima, K.; Komeda, K.; Ide, T.; Kubota, N.; Terauchi, Y.; Tobe, K.; et al. The mechanisms by which both heterozygous peroxisome proliferator-activated receptor gamma (PPARgamma) deficiency and PPARgamma agonist improve insulin resistance. J. Biol. Chem. 2001, 276, 41245–41254. [Google Scholar] [CrossRef] [Green Version]
  440. Heikkinen, S.; Argmann, C.; Feige, J.N.; Koutnikova, H.; Champy, M.F.; Dali-Youcef, N.; Schadt, E.E.; Laakso, M.; Auwerx, J. The Pro12Ala PPARgamma2 variant determines metabolism at the gene-environment interface. Cell Metab. 2009, 9, 88–98. [Google Scholar] [CrossRef]
  441. Deeb, S.S.; Fajas, L.; Nemoto, M.; Pihlajamaki, J.; Mykkanen, L.; Kuusisto, J.; Laakso, M.; Fujimoto, W.; Auwerx, J. A Pro12Ala substitution in PPARgamma2 associated with decreased receptor activity, lower body mass index and improved insulin sensitivity. Nat. Genet. 1998, 20, 284–287. [Google Scholar] [CrossRef]
  442. Altshuler, D.; Hirschhorn, J.N.; Klannemark, M.; Lindgren, C.M.; Vohl, M.C.; Nemesh, J.; Lane, C.R.; Schaffner, S.F.; Bolk, S.; Brewer, C.; et al. The common PPARgamma Pro12Ala polymorphism is associated with decreased risk of type 2 diabetes. Nat. Genet. 2000, 26, 76–80. [Google Scholar] [CrossRef]
  443. Rieusset, J.; Touri, F.; Michalik, L.; Escher, P.; Desvergne, B.; Niesor, E.; Wahli, W. A new selective peroxisome proliferator-activated receptor gamma antagonist with antiobesity and antidiabetic activity. Mol. Endocrinol. 2002, 16, 2628–2644. [Google Scholar] [CrossRef] [Green Version]
  444. Mukherjee, R.; Hoener, P.A.; Jow, L.; Bilakovics, J.; Klausing, K.; Mais, D.E.; Faulkner, A.; Croston, G.E.; Paterniti, J.R., Jr. A selective peroxisome proliferator-activated receptor-gamma (PPARgamma) modulator blocks adipocyte differentiation but stimulates glucose uptake in 3T3-L1 adipocytes. Mol. Endocrinol. 2000, 14, 1425–1433. [Google Scholar] [CrossRef]
  445. Ferrara, N.; Rinaldi, B.; Corbi, G.; Conti, V.; Stiuso, P.; Boccuti, S.; Rengo, G.; Rossi, F.; Filippelli, A. Exercise training promotes SIRT1 activity in aged rats. Rejuvenation Res. 2008, 11, 139–150. [Google Scholar] [CrossRef]
  446. Corbi, G.; Conti, V.; Scapagnini, G.; Filippelli, A.; Ferrara, N. Role of sirtuins, calorie restriction and physical activity in aging. Front. Biosci. (Elite Ed.) 2012, 4, 768–778. [Google Scholar]
  447. Rack, J.G.; Morra, R.; Barkauskaite, E.; Kraehenbuehl, R.; Ariza, A.; Qu, Y.; Ortmayer, M.; Leidecker, O.; Cameron, D.R.; Matic, I.; et al. Identification of a Class of Protein ADP-Ribosylating Sirtuins in Microbial Pathogens. Mol. Cell 2015, 59, 309–320. [Google Scholar] [CrossRef] [Green Version]
  448. Liou, G.G.; Tanny, J.C.; Kruger, R.G.; Walz, T.; Moazed, D. Assembly of the SIR complex and its regulation by O-acetyl-ADP-ribose, a product of NAD-dependent histone deacetylation. Cell 2005, 121, 515–527. [Google Scholar] [CrossRef] [Green Version]
  449. Vaquero, A.; Scher, M.; Lee, D.; Erdjument-Bromage, H.; Tempst, P.; Reinberg, D. Human SirT1 interacts with histone H1 and promotes formation of facultative heterochromatin. Mol. Cell 2004, 16, 93–105. [Google Scholar] [CrossRef]
  450. Dang, W. The controversial world of sirtuins. Drug Discov. Today Technol. 2014, 12, e9–e17. [Google Scholar] [CrossRef] [Green Version]
  451. Houtkooper, R.H.; Pirinen, E.; Auwerx, J. Sirtuins as regulators of metabolism and healthspan. Nat. Rev. Mol. Cell Biol. 2012, 13, 225–238. [Google Scholar] [CrossRef] [Green Version]
  452. Shimazu, T.; Hirschey, M.D.; Hua, L.; Dittenhafer-Reed, K.E.; Schwer, B.; Lombard, D.B.; Li, Y.; Bunkenborg, J.; Alt, F.W.; Denu, J.M.; et al. SIRT3 deacetylates mitochondrial 3-hydroxy-3-methylglutaryl CoA synthase 2 and regulates ketone body production. Cell Metab. 2010, 12, 654–661. [Google Scholar] [CrossRef] [Green Version]
  453. Hirschey, M.D.; Shimazu, T.; Goetzman, E.; Jing, E.; Schwer, B.; Lombard, D.B.; Grueter, C.A.; Harris, C.; Biddinger, S.; Ilkayeva, O.R.; et al. SIRT3 regulates mitochondrial fatty-acid oxidation by reversible enzyme deacetylation. Nature 2010, 464, 121–125. [Google Scholar] [CrossRef] [Green Version]
  454. Qiu, X.; Brown, K.; Hirschey, M.D.; Verdin, E.; Chen, D. Calorie restriction reduces oxidative stress by SIRT3-mediated SOD2 activation. Cell Metab. 2010, 12, 662–667. [Google Scholar] [CrossRef] [Green Version]
  455. Preyat, N.; Leo, O. Sirtuin deacylases: A molecular link between metabolism and immunity. J. Leukoc. Biol. 2013, 93, 669–680. [Google Scholar] [CrossRef]
  456. Satoh, A.; Brace, C.S.; Ben-Josef, G.; West, T.; Wozniak, D.F.; Holtzman, D.M.; Herzog, E.D.; Imai, S. SIRT1 promotes the central adaptive response to diet restriction through activation of the dorsomedial and lateral nuclei of the hypothalamus. J. Neurosci. 2010, 30, 10220–10232. [Google Scholar] [CrossRef]
  457. Kaeberlein, M.; McVey, M.; Guarente, L. The SIR2/3/4 complex and SIR2 alone promote longevity in Saccharomyces cerevisiae by two different mechanisms. Genes Dev. 1999, 13, 2570–2580. [Google Scholar] [CrossRef] [Green Version]
  458. Whitaker, R.; Faulkner, S.; Miyokawa, R.; Burhenn, L.; Henriksen, M.; Wood, J.G.; Helfand, S.L. Increased expression of Drosophila Sir2 extends life span in a dose-dependent manner. Aging (Albany N. Y.) 2013, 5, 682–691. [Google Scholar] [CrossRef] [Green Version]
  459. Lin, S.J.; Defossez, P.A.; Guarente, L. Requirement of NAD and SIR2 for life-span extension by calorie restriction in Saccharomyces cerevisiae. Science 2000, 289, 2126–2128. [Google Scholar] [CrossRef] [Green Version]
  460. Rogina, B.; Helfand, S.L. Sir2 mediates longevity in the fly through a pathway related to calorie restriction. Proc. Natl. Acad. Sci. USA 2004, 101, 15998–16003. [Google Scholar] [CrossRef] [Green Version]
  461. Tissenbaum, H.A.; Guarente, L. Increased dosage of a sir-2 gene extends lifespan in Caenorhabditis elegans. Nature 2001, 410, 227–230. [Google Scholar] [CrossRef]
  462. Oberdoerffer, P.; Michan, S.; McVay, M.; Mostoslavsky, R.; Vann, J.; Park, S.K.; Hartlerode, A.; Stegmuller, J.; Hafner, A.; Loerch, P.; et al. SIRT1 redistribution on chromatin promotes genomic stability but alters gene expression during aging. Cell 2008, 135, 907–918. [Google Scholar] [CrossRef] [Green Version]
  463. Rodgers, J.T.; Lerin, C.; Haas, W.; Gygi, S.P.; Spiegelman, B.M.; Puigserver, P. Nutrient control of glucose homeostasis through a complex of PGC-1alpha and SIRT1. Nature 2005, 434, 113–118. [Google Scholar] [CrossRef]
  464. Nemoto, S.; Fergusson, M.M.; Finkel, T. SIRT1 functionally interacts with the metabolic regulator and transcriptional coactivator PGC-1{alpha}. J. Biol. Chem. 2005, 280, 16456–16460. [Google Scholar] [CrossRef] [Green Version]
  465. Lagouge, M.; Argmann, C.; Gerhart-Hines, Z.; Meziane, H.; Lerin, C.; Daussin, F.; Messadeq, N.; Milne, J.; Lambert, P.; Elliott, P.; et al. Resveratrol improves mitochondrial function and protects against metabolic disease by activating SIRT1 and PGC-1alpha. Cell 2006, 127, 1109–1122. [Google Scholar] [CrossRef]
  466. Vaziri, H.; Dessain, S.K.; Ng Eaton, E.; Imai, S.I.; Frye, R.A.; Pandita, T.K.; Guarente, L.; Weinberg, R.A. hSIR2(SIRT1) functions as an NAD-dependent p53 deacetylase. Cell 2001, 107, 149–159. [Google Scholar] [CrossRef] [Green Version]
  467. Luo, J.; Nikolaev, A.Y.; Imai, S.; Chen, D.; Su, F.; Shiloh, A.; Guarente, L.; Gu, W. Negative control of p53 by Sir2alpha promotes cell survival under stress. Cell 2001, 107, 137–148. [Google Scholar] [CrossRef] [Green Version]
  468. Daitoku, H.; Hatta, M.; Matsuzaki, H.; Aratani, S.; Ohshima, T.; Miyagishi, M.; Nakajima, T.; Fukamizu, A. Silent information regulator 2 potentiates Foxo1-mediated transcription through its deacetylase activity. Proc. Natl. Acad. Sci. USA 2004, 101, 10042–10047. [Google Scholar] [CrossRef] [Green Version]
  469. van der Horst, A.; Tertoolen, L.G.; de Vries-Smits, L.M.; Frye, R.A.; Medema, R.H.; Burgering, B.M. FOXO4 is acetylated upon peroxide stress and deacetylated by the longevity protein hSir2(SIRT1). J. Biol. Chem. 2004, 279, 28873–28879. [Google Scholar] [CrossRef] [Green Version]
  470. Accili, D.; Arden, K.C. FoxOs at the crossroads of cellular metabolism, differentiation, and transformation. Cell 2004, 117, 421–426. [Google Scholar] [CrossRef] [Green Version]
  471. Martins, R.; Lithgow, G.J.; Link, W. Long live FOXO: Unraveling the role of FOXO proteins in aging and longevity. Aging Cell 2016, 15, 196–207. [Google Scholar] [CrossRef]
  472. Gross, D.N.; van den Heuvel, A.P.; Birnbaum, M.J. The role of FoxO in the regulation of metabolism. Oncogene 2008, 27, 2320–2336. [Google Scholar] [CrossRef] [Green Version]
  473. Frescas, D.; Valenti, L.; Accili, D. Nuclear trapping of the forkhead transcription factor FoxO1 via Sirt-dependent deacetylation promotes expression of glucogenetic genes. J. Biol. Chem. 2005, 280, 20589–20595. [Google Scholar] [CrossRef] [Green Version]
  474. Brunet, A.; Sweeney, L.B.; Sturgill, J.F.; Chua, K.F.; Greer, P.L.; Lin, Y.; Tran, H.; Ross, S.E.; Mostoslavsky, R.; Cohen, H.Y.; et al. Stress-dependent regulation of FOXO transcription factors by the SIRT1 deacetylase. Science 2004, 303, 2011–2015. [Google Scholar] [CrossRef] [Green Version]
  475. Bordone, L.; Motta, M.C.; Picard, F.; Robinson, A.; Jhala, U.S.; Apfeld, J.; McDonagh, T.; Lemieux, M.; McBurney, M.; Szilvasi, A.; et al. Sirt1 regulates insulin secretion by repressing UCP2 in pancreatic beta cells. PLoS Biol. 2006, 4, e31. [Google Scholar] [CrossRef]
  476. Moynihan, K.A.; Grimm, A.A.; Plueger, M.M.; Bernal-Mizrachi, E.; Ford, E.; Cras-Meneur, C.; Permutt, M.A.; Imai, S. Increased dosage of mammalian Sir2 in pancreatic beta cells enhances glucose-stimulated insulin secretion in mice. Cell Metab. 2005, 2, 105–117. [Google Scholar] [CrossRef] [Green Version]
  477. Haigis, M.C.; Mostoslavsky, R.; Haigis, K.M.; Fahie, K.; Christodoulou, D.C.; Murphy, A.J.; Valenzuela, D.M.; Yancopoulos, G.D.; Karow, M.; Blander, G.; et al. SIRT4 inhibits glutamate dehydrogenase and opposes the effects of calorie restriction in pancreatic beta cells. Cell 2006, 126, 941–954. [Google Scholar] [CrossRef] [Green Version]
  478. Bae, N.S.; Swanson, M.J.; Vassilev, A.; Howard, B.H. Human histone deacetylase SIRT2 interacts with the homeobox transcription factor HOXA10. J. Biochem. 2004, 135, 695–700. [Google Scholar] [CrossRef]
  479. North, B.J.; Marshall, B.L.; Borra, M.T.; Denu, J.M.; Verdin, E. The human Sir2 ortholog, SIRT2, is an NAD+-dependent tubulin deacetylase. Mol. Cell 2003, 11, 437–444. [Google Scholar] [CrossRef]
  480. Wang, F.; Nguyen, M.; Qin, F.X.; Tong, Q. SIRT2 deacetylates FOXO3a in response to oxidative stress and caloric restriction. Aging Cell 2007, 6, 505–514. [Google Scholar] [CrossRef]
  481. Jing, E.; Gesta, S.; Kahn, C.R. SIRT2 regulates adipocyte differentiation through FoxO1 acetylation/deacetylation. Cell Metab. 2007, 6, 105–114. [Google Scholar] [CrossRef] [Green Version]
  482. Dryden, S.C.; Nahhas, F.A.; Nowak, J.E.; Goustin, A.S.; Tainsky, M.A. Role for human SIRT2 NAD-dependent deacetylase activity in control of mitotic exit in the cell cycle. Mol. Cell Biol. 2003, 23, 3173–3185. [Google Scholar] [CrossRef] [Green Version]
  483. Sinclair, D.A.; Guarente, L. Extrachromosomal rDNA circles--a cause of aging in yeast. Cell 1997, 91, 1033–1042. [Google Scholar] [CrossRef] [Green Version]
  484. Starai, V.J.; Celic, I.; Cole, R.N.; Boeke, J.D.; Escalante-Semerena, J.C. Sir2-dependent activation of acetyl-CoA synthetase by deacetylation of active lysine. Science 2002, 298, 2390–2392. [Google Scholar] [CrossRef]
  485. Onyango, P.; Celic, I.; McCaffery, J.M.; Boeke, J.D.; Feinberg, A.P. SIRT3, a human SIR2 homologue, is an NAD-dependent deacetylase localized to mitochondria. Proc. Natl. Acad. Sci. USA 2002, 99, 13653–13658. [Google Scholar] [CrossRef] [Green Version]
  486. Shi, T.; Wang, F.; Stieren, E.; Tong, Q. SIRT3, a mitochondrial sirtuin deacetylase, regulates mitochondrial function and thermogenesis in brown adipocytes. J. Biol. Chem. 2005, 280, 13560–13567. [Google Scholar] [CrossRef] [Green Version]
  487. Rose, G.; Dato, S.; Altomare, K.; Bellizzi, D.; Garasto, S.; Greco, V.; Passarino, G.; Feraco, E.; Mari, V.; Barbi, C.; et al. Variability of the SIRT3 gene, human silent information regulator Sir2 homologue, and survivorship in the elderly. Exp. Gerontol. 2003, 38, 1065–1070. [Google Scholar] [CrossRef]
  488. Nasrin, N.; Wu, X.; Fortier, E.; Feng, Y.; Bare, O.C.; Chen, S.; Ren, X.; Wu, Z.; Streeper, R.S.; Bordone, L. SIRT4 regulates fatty acid oxidation and mitochondrial gene expression in liver and muscle cells. J. Biol. Chem. 2010, 285, 31995–32002. [Google Scholar] [CrossRef] [Green Version]
  489. Michishita, E.; Park, J.Y.; Burneskis, J.M.; Barrett, J.C.; Horikawa, I. Evolutionarily conserved and nonconserved cellular localizations and functions of human SIRT proteins. Mol. Biol. Cell 2005, 16, 4623–4635. [Google Scholar] [CrossRef] [Green Version]
  490. McCord, R.A.; Michishita, E.; Hong, T.; Berber, E.; Boxer, L.D.; Kusumoto, R.; Guan, S.; Shi, X.; Gozani, O.; Burlingame, A.L.; et al. SIRT6 stabilizes DNA-dependent protein kinase at chromatin for DNA double-strand break repair. Aging (Albany N. Y.) 2009, 1, 109–121. [Google Scholar] [CrossRef] [Green Version]
  491. Mostoslavsky, R.; Chua, K.F.; Lombard, D.B.; Pang, W.W.; Fischer, M.R.; Gellon, L.; Liu, P.; Mostoslavsky, G.; Franco, S.; Murphy, M.M.; et al. Genomic instability and aging-like phenotype in the absence of mammalian SIRT6. Cell 2006, 124, 315–329. [Google Scholar] [CrossRef] [Green Version]
  492. Kanfi, Y.; Naiman, S.; Amir, G.; Peshti, V.; Zinman, G.; Nahum, L.; Bar-Joseph, Z.; Cohen, H.Y. The sirtuin SIRT6 regulates lifespan in male mice. Nature 2012, 483, 218–221. [Google Scholar] [CrossRef]
  493. Luo, L.L.; Chen, X.C.; Fu, Y.C.; Xu, J.J.; Li, L.; Lin, X.H.; Xiang, Y.F.; Zhang, X.M. The effects of caloric restriction and a high-fat diet on ovarian lifespan and the expression of SIRT1 and SIRT6 proteins in rats. Aging Clin. Exp. Res. 2012, 24, 125–133. [Google Scholar] [CrossRef]
  494. Ford, E.; Voit, R.; Liszt, G.; Magin, C.; Grummt, I.; Guarente, L. Mammalian Sir2 homolog SIRT7 is an activator of RNA polymerase I transcription. Genes Dev. 2006, 20, 1075–1080. [Google Scholar] [CrossRef] [Green Version]
  495. Vazquez, B.N.; Thackray, J.K.; Simonet, N.G.; Kane-Goldsmith, N.; Martinez-Redondo, P.; Nguyen, T.; Bunting, S.; Vaquero, A.; Tischfield, J.A.; Serrano, L. SIRT7 promotes genome integrity and modulates non-homologous end joining DNA repair. EMBO J. 2016, 35, 1488–1503. [Google Scholar] [CrossRef]
  496. Geng, Y.Q.; Li, T.T.; Liu, X.Y.; Li, Z.H.; Fu, Y.C. SIRT1 and SIRT5 activity expression and behavioral responses to calorie restriction. J. Cell Biochem. 2011, 112, 3755–3761. [Google Scholar] [CrossRef]
  497. Ran, M.; Li, Z.; Yang, L.; Tong, L.; Zhang, L.; Dong, H. Calorie restriction attenuates cerebral ischemic injury via increasing SIRT1 synthesis in the rat. Brain Res. 2015, 1610, 61–68. [Google Scholar] [CrossRef]
  498. Nisoli, E.; Tonello, C.; Cardile, A.; Cozzi, V.; Bracale, R.; Tedesco, L.; Falcone, S.; Valerio, A.; Cantoni, O.; Clementi, E.; et al. Calorie restriction promotes mitochondrial biogenesis by inducing the expression of eNOS. Science 2005, 310, 314–317. [Google Scholar] [CrossRef]
  499. Yu, W.; Zhou, H.F.; Lin, R.B.; Fu, Y.C.; Wang, W. Shortterm calorie restriction activates SIRT14 and 7 in cardiomyocytes in vivo and in vitro. Mol. Med. Rep. 2014, 9, 1218–1224. [Google Scholar] [CrossRef] [Green Version]
  500. Yu, W.; Qin, J.; Chen, C.; Fu, Y.; Wang, W. Moderate calorie restriction attenuates ageassociated alterations and improves cardiac function by increasing SIRT1 and SIRT3 expression. Mol. Med. Rep. 2018, 18, 4087–4094. [Google Scholar] [CrossRef] [Green Version]
  501. Chen, L.L.; Deng, X.Q.; Li, N.X. Effects of calorie restriction on SIRT1 expression in liver of nonalcoholic fatty liver disease: Experiment with rats. Zhonghua Yi Xue Za Zhi 2007, 87, 1434–1437. [Google Scholar]
  502. Boily, G.; Seifert, E.L.; Bevilacqua, L.; He, X.H.; Sabourin, G.; Estey, C.; Moffat, C.; Crawford, S.; Saliba, S.; Jardine, K.; et al. SirT1 regulates energy metabolism and response to caloric restriction in mice. PLoS ONE 2008, 3, e1759. [Google Scholar] [CrossRef]
  503. Bordone, L.; Cohen, D.; Robinson, A.; Motta, M.C.; van Veen, E.; Czopik, A.; Steele, A.D.; Crowe, H.; Marmor, S.; Luo, J.; et al. SIRT1 transgenic mice show phenotypes resembling calorie restriction. Aging Cell 2007, 6, 759–767. [Google Scholar] [CrossRef]
  504. Bhullar, K.S.; Hubbard, B.P. Lifespan and healthspan extension by resveratrol. Biochim. Biophys. Acta 2015, 1852, 1209–1218. [Google Scholar] [CrossRef] [Green Version]
  505. Howitz, K.T.; Bitterman, K.J.; Cohen, H.Y.; Lamming, D.W.; Lavu, S.; Wood, J.G.; Zipkin, R.E.; Chung, P.; Kisielewski, A.; Zhang, L.L.; et al. Small molecule activators of sirtuins extend Saccharomyces cerevisiae lifespan. Nature 2003, 425, 191–196. [Google Scholar] [CrossRef]
  506. Wood, J.G.; Rogina, B.; Lavu, S.; Howitz, K.; Helfand, S.L.; Tatar, M.; Sinclair, D. Sirtuin activators mimic caloric restriction and delay ageing in metazoans. Nature 2004, 430, 686–689. [Google Scholar] [CrossRef]
  507. Baur, J.A.; Pearson, K.J.; Price, N.L.; Jamieson, H.A.; Lerin, C.; Kalra, A.; Prabhu, V.V.; Allard, J.S.; Lopez-Lluch, G.; Lewis, K.; et al. Resveratrol improves health and survival of mice on a high-calorie diet. Nature 2006, 444, 337–342. [Google Scholar] [CrossRef]
  508. Timmers, S.; Konings, E.; Bilet, L.; Houtkooper, R.H.; van de Weijer, T.; Goossens, G.H.; Hoeks, J.; van der Krieken, S.; Ryu, D.; Kersten, S.; et al. Calorie restriction-like effects of 30 days of resveratrol supplementation on energy metabolism and metabolic profile in obese humans. Cell Metab. 2011, 14, 612–622. [Google Scholar] [CrossRef] [Green Version]
  509. Barger, J.L.; Kayo, T.; Vann, J.M.; Arias, E.B.; Wang, J.; Hacker, T.A.; Wang, Y.; Raederstorff, D.; Morrow, J.D.; Leeuwenburgh, C.; et al. A low dose of dietary resveratrol partially mimics caloric restriction and retards aging parameters in mice. PLoS ONE 2008, 3, e2264. [Google Scholar] [CrossRef]
  510. Laurent, G.; de Boer, V.C.; Finley, L.W.; Sweeney, M.; Lu, H.; Schug, T.T.; Cen, Y.; Jeong, S.M.; Li, X.; Sauve, A.A.; et al. SIRT4 represses peroxisome proliferator-activated receptor alpha activity to suppress hepatic fat oxidation. Mol. Cell Biol. 2013, 33, 4552–4561. [Google Scholar] [CrossRef] [Green Version]
  511. Purushotham, A.; Schug, T.T.; Xu, Q.; Surapureddi, S.; Guo, X.; Li, X. Hepatocyte-specific deletion of SIRT1 alters fatty acid metabolism and results in hepatic steatosis and inflammation. Cell Metab. 2009, 9, 327–338. [Google Scholar] [CrossRef] [Green Version]
  512. Oka, S.; Zhai, P.; Yamamoto, T.; Ikeda, Y.; Byun, J.; Hsu, C.P.; Sadoshima, J. Peroxisome Proliferator Activated Receptor-alpha Association With Silent Information Regulator 1 Suppresses Cardiac Fatty Acid Metabolism in the Failing Heart. Circ. Heart Fail. 2015, 8, 1123–1132. [Google Scholar] [CrossRef] [Green Version]
  513. Oka, S.; Alcendor, R.; Zhai, P.; Park, J.Y.; Shao, D.; Cho, J.; Yamamoto, T.; Tian, B.; Sadoshima, J. PPARalpha-Sirt1 complex mediates cardiac hypertrophy and failure through suppression of the ERR transcriptional pathway. Cell Metab. 2011, 14, 598–611. [Google Scholar] [CrossRef] [Green Version]
  514. Oka, S.; Zhai, P.; Alcendor, R.; Park, J.Y.; Tian, B.; Sadoshima, J. Suppression of ERR targets by a PPARalpha/Sirt1 complex in the failing heart. Cell Cycle 2012, 11, 856–864. [Google Scholar] [CrossRef] [Green Version]
  515. Planavila, A.; Iglesias, R.; Giralt, M.; Villarroya, F. Sirt1 acts in association with PPARalpha to protect the heart from hypertrophy, metabolic dysregulation, and inflammation. Cardiovasc. Res. 2011, 90, 276–284. [Google Scholar] [CrossRef] [Green Version]
  516. Okazaki, M.; Iwasaki, Y.; Nishiyama, M.; Taguchi, T.; Tsugita, M.; Nakayama, S.; Kambayashi, M.; Hashimoto, K.; Terada, Y. PPARbeta/delta regulates the human SIRT1 gene transcription via Sp1. Endocr. J. 2010, 57, 403–413. [Google Scholar] [CrossRef] [Green Version]
  517. Barroso, E.; Eyre, E.; Palomer, X.; Vazquez-Carrera, M. The peroxisome proliferator-activated receptor beta/delta (PPARbeta/delta) agonist GW501516 prevents TNF-alpha-induced NF-kappaB activation in human HaCaT cells by reducing p65 acetylation through AMPK and SIRT1. Biochem. Pharmacol. 2011, 81, 534–543. [Google Scholar] [CrossRef]
  518. Gong, K.; Qu, B.; Wang, C.; Zhou, J.; Liao, D.; Zheng, W.; Pan, X. Peroxisome Proliferator-Activated Receptor alpha Facilitates Osteogenic Differentiation in MC3T3-E1 Cells via the Sirtuin 1-Dependent Signaling Pathway. Mol. Cells 2017, 40, 393–400. [Google Scholar] [CrossRef] [Green Version]
  519. Gong, K.; Qu, B.; Liao, D.; Liu, D.; Wang, C.; Zhou, J.; Pan, X. MiR-132 regulates osteogenic differentiation via downregulating Sirtuin1 in a peroxisome proliferator-activated receptor beta/delta-dependent manner. Biochem. Biophys. Res. Commun. 2016, 478, 260–267. [Google Scholar] [CrossRef]
  520. Qu, B.; Ma, Y.; Yan, M.; Gong, K.; Liang, F.; Deng, S.; Jiang, K.; Ma, Z.; Pan, X. Sirtuin1 promotes osteogenic differentiation through downregulation of peroxisome proliferator-activated receptor gamma in MC3T3-E1 cells. Biochem. Biophys. Res. Commun. 2016, 478, 439–445. [Google Scholar] [CrossRef]
  521. Kim, M.Y.; Kang, E.S.; Ham, S.A.; Hwang, J.S.; Yoo, T.S.; Lee, H.; Paek, K.S.; Park, C.; Lee, H.T.; Kim, J.H.; et al. The PPARdelta-mediated inhibition of angiotensin II-induced premature senescence in human endothelial cells is SIRT1-dependent. Biochem. Pharmacol. 2012, 84, 1627–1634. [Google Scholar] [CrossRef]
  522. Han, L.; Zhou, R.; Niu, J.; McNutt, M.A.; Wang, P.; Tong, T. SIRT1 is regulated by a PPAR{gamma}-SIRT1 negative feedback loop associated with senescence. Nucleic Acids Res. 2010, 38, 7458–7471. [Google Scholar] [CrossRef] [Green Version]
  523. Zhang, J.; Zhang, Y.; Xiao, F.; Liu, Y.; Wang, J.; Gao, H.; Rong, S.; Yao, Y.; Li, J.; Xu, G. The peroxisome proliferator-activated receptor gamma agonist pioglitazone prevents NF-kappaB activation in cisplatin nephrotoxicity through the reduction of p65 acetylation via the AMPK-SIRT1/p300 pathway. Biochem. Pharmacol. 2016, 101, 100–111. [Google Scholar] [CrossRef] [Green Version]
  524. Picard, F.; Kurtev, M.; Chung, N.; Topark-Ngarm, A.; Senawong, T.; Machado De Oliveira, R.; Leid, M.; McBurney, M.W.; Guarente, L. Sirt1 promotes fat mobilization in white adipocytes by repressing PPAR-gamma. Nature 2004, 429, 771–776. [Google Scholar] [CrossRef]
  525. Floyd, Z.E.; Wang, Z.Q.; Kilroy, G.; Cefalu, W.T. Modulation of peroxisome proliferator-activated receptor gamma stability and transcriptional activity in adipocytes by resveratrol. Metabolism 2008, 57, S32–S38. [Google Scholar] [CrossRef] [Green Version]
  526. Hagen, T.M. Oxidative stress, redox imbalance, and the aging process. Antioxid. Redox Signal. 2003, 5, 503–506. [Google Scholar] [CrossRef]
  527. Navarro, A.; Boveris, A. The mitochondrial energy transduction system and the aging process. Am. J. Physiol. Cell Physiol. 2007, 292, C670–C686. [Google Scholar] [CrossRef]
  528. Harman, D. Aging: A theory based on free radical and radiation chemistry. J. Gerontol. 1956, 11, 298–300. [Google Scholar] [CrossRef] [Green Version]
  529. Merry, B.J. Molecular mechanisms linking calorie restriction and longevity. Int. J. Biochem. Cell Biol. 2002, 34, 1340–1354. [Google Scholar]
  530. Sanz, A.; Caro, P.; Ibanez, J.; Gomez, J.; Gredilla, R.; Barja, G. Dietary restriction at old age lowers mitochondrial oxygen radical production and leak at complex I and oxidative DNA damage in rat brain. J. Bioenerg. Biomembr. 2005, 37, 83–90. [Google Scholar] [CrossRef]
  531. Yu, B.P. Aging and oxidative stress: Modulation by dietary restriction. Free Radic. Biol. Med. 1996, 21, 651–668. [Google Scholar] [CrossRef]
  532. Barja, G. Aging in vertebrates, and the effect of caloric restriction: A mitochondrial free radical production-DNA damage mechanism? Biol. Rev. Cambr. Philos. Soc. 2004, 79, 235–251. [Google Scholar]
  533. Barja, G. Free radicals and aging. Trends Neurosci. 2004, 27, 595–600. [Google Scholar] [CrossRef]
  534. Forster, M.J.; Sohal, B.H.; Sohal, R.S. Reversible effects of long-term caloric restriction on protein oxidative damage. J. Gerontol. Ser. A Biol. Sci. Med. Sci. 2000, 55, B522–B529. [Google Scholar]
  535. Lambert, A.J.; Merry, B.J. Lack of effect of caloric restriction on bioenergetics and reactive oxygen species production in intact rat hepatocytes. J. Gerontol. Ser. A Biol. Sci. Med. Sci. 2005, 60, 175–180. [Google Scholar]
  536. Lambert, A.J.; Portero-Otin, M.; Pamplona, R.; Merry, B.J. Effect of ageing and caloric restriction on specific markers of protein oxidative damage and membrane peroxidizability in rat liver mitochondria. Mech. Ageing Dev. 2004, 125, 529–538. [Google Scholar] [CrossRef]
  537. Bagi, Z.; Koller, A.; Kaley, G. PPARgamma activation, by reducing oxidative stress, increases NO bioavailability in coronary arterioles of mice with Type 2 diabetes. Am. J. Physiol. Heart Circ. Physiol. 2004, 286, H742–H748. [Google Scholar] [CrossRef] [Green Version]
  538. Meerarani, P.; Reiterer, G.; Toborek, M.; Hennig, B. Zinc modulates PPARgamma signaling and activation of porcine endothelial cells. J. Nutr. 2003, 133, 3058–3064. [Google Scholar] [CrossRef] [Green Version]
  539. Iemitsu, M.; Miyauchi, T.; Maeda, S.; Tanabe, T.; Takanashi, M.; Irukayama-Tomobe, Y.; Sakai, S.; Ohmori, H.; Matsuda, M.; Yamaguchi, I. Aging-induced decrease in the PPAR-alpha level in hearts is improved by exercise training. Am. J. Physiol. Heart Circ. Physiol. 2002, 283, H1750–H1760. [Google Scholar] [CrossRef]
  540. Patterson, A.D.; Shah, Y.M.; Matsubara, T.; Krausz, K.W.; Gonzalez, F.J. Peroxisome proliferator-activated receptor alpha induction of uncoupling protein 2 protects against acetaminophen-induced liver toxicity. Hepatology 2012, 56, 281–290. [Google Scholar] [CrossRef] [Green Version]
  541. Sekulic-Jablanovic, M.; Petkovic, V.; Wright, M.B.; Kucharava, K.; Huerzeler, N.; Levano, S.; Brand, Y.; Leitmeyer, K.; Glutz, A.; Bausch, A.; et al. Effects of peroxisome proliferator activated receptors (PPAR)-gamma and -alpha agonists on cochlear protection from oxidative stress. PLoS ONE 2017, 12, e0188596. [Google Scholar] [CrossRef]
  542. Girnun, G.D.; Domann, F.E.; Moore, S.A.; Robbins, M.E. Identification of a functional peroxisome proliferator-activated receptor response element in the rat catalase promoter. Mol. Endocrinol. 2002, 16, 2793–2801. [Google Scholar] [CrossRef] [Green Version]
  543. Okuno, Y.; Matsuda, M.; Miyata, Y.; Fukuhara, A.; Komuro, R.; Shimabukuro, M.; Shimomura, I. Human catalase gene is regulated by peroxisome proliferator activated receptor-gamma through a response element distinct from that of mouse. Endocr. J. 2010, 57, 303–309. [Google Scholar] [CrossRef] [Green Version]
  544. Ding, G.; Fu, M.; Qin, Q.; Lewis, W.; Kim, H.W.; Fukai, T.; Bacanamwo, M.; Chen, Y.E.; Schneider, M.D.; Mangelsdorf, D.J.; et al. Cardiac peroxisome proliferator-activated receptor gamma is essential in protecting cardiomyocytes from oxidative damage. Cardiovasc. Res. 2007, 76, 269–279. [Google Scholar] [CrossRef] [Green Version]
  545. Chung, S.S.; Kim, M.; Youn, B.S.; Lee, N.S.; Park, J.W.; Lee, I.K.; Lee, Y.S.; Kim, J.B.; Cho, Y.M.; Lee, H.K.; et al. Glutathione peroxidase 3 mediates the antioxidant effect of peroxisome proliferator-activated receptor gamma in human skeletal muscle cells. Mol. Cell Biol. 2009, 29, 20–30. [Google Scholar] [CrossRef] [Green Version]
  546. Zhang, F.; Liu, F.; Yan, M.; Ji, H.; Hu, L.; Li, X.; Qian, J.; He, X.; Zhang, L.; Shen, A.; et al. Peroxisome proliferator-activated receptor-gamma agonists suppress iNOS expression induced by LPS in rat primary Schwann cells. J. Neuroimmunol. 2010, 218, 36–47. [Google Scholar] [CrossRef]
  547. Vandewalle, B.; Moerman, E.; Lefebvre, B.; Defrance, F.; Gmyr, V.; Lukowiak, B.; Kerr Conte, J.; Pattou, F. PPARgamma-dependent and -independent effects of rosiglitazone on lipotoxic human pancreatic islets. Biochem. Biophys. Res. Commun. 2008, 366, 1096–1101. [Google Scholar] [CrossRef]
  548. Kleinhenz, J.M.; Kleinhenz, D.J.; You, S.; Ritzenthaler, J.D.; Hansen, J.M.; Archer, D.R.; Sutliff, R.L.; Hart, C.M. Disruption of endothelial peroxisome proliferator-activated receptor-gamma reduces vascular nitric oxide production. Am. J. Physiol. Heart Circ. Physiol. 2009, 297, H1647–H1654. [Google Scholar] [CrossRef] [Green Version]
  549. Ricote, M.; Li, A.C.; Willson, T.M.; Kelly, C.J.; Glass, C.K. The peroxisome proliferator-activated receptor-gamma is a negative regulator of macrophage activation. Nature 1998, 391, 79–82. [Google Scholar] [CrossRef]
  550. Li, M.; Pascual, G.; Glass, C.K. Peroxisome proliferator-activated receptor gamma-dependent repression of the inducible nitric oxide synthase gene. Mol. Cell Biol. 2000, 20, 4699–4707. [Google Scholar] [CrossRef] [Green Version]
  551. Ghosh, N.; Chaki, R.; Mandal, V.; Mandal, S.C. COX-2 as a target for cancer chemotherapy. Pharmacol. Rep. 2010, 62, 233–244. [Google Scholar]
  552. Morgan, M.J.; Liu, Z.G. Crosstalk of reactive oxygen species and NF-kappaB signaling. Cell Res. 2011, 21, 103–115. [Google Scholar] [CrossRef] [Green Version]
  553. Wang, X.; Wang, Z.; Liu, J.Z.; Hu, J.X.; Chen, H.L.; Li, W.L.; Hai, C.X. Double antioxidant activities of rosiglitazone against high glucose-induced oxidative stress in hepatocyte. Toxicol. In Vitro 2011, 25, 839–847. [Google Scholar] [CrossRef]
  554. Aleshin, S.; Grabeklis, S.; Hanck, T.; Sergeeva, M.; Reiser, G. Peroxisome proliferator-activated receptor (PPAR)-gamma positively controls and PPARalpha negatively controls cyclooxygenase-2 expression in rat brain astrocytes through a convergence on PPARbeta/delta via mutual control of PPAR expression levels. Mol. Pharmacol. 2009, 76, 414–424. [Google Scholar] [CrossRef]
  555. Collino, M.; Aragno, M.; Mastrocola, R.; Gallicchio, M.; Rosa, A.C.; Dianzani, C.; Danni, O.; Thiemermann, C.; Fantozzi, R. Modulation of the oxidative stress and inflammatory response by PPAR-gamma agonists in the hippocampus of rats exposed to cerebral ischemia/reperfusion. Eur. J. Pharmacol. 2006, 530, 70–80. [Google Scholar] [CrossRef]
  556. Zhao, Y.; Patzer, A.; Herdegen, T.; Gohlke, P.; Culman, J. Activation of cerebral peroxisome proliferator-activated receptors gamma promotes neuroprotection by attenuation of neuronal cyclooxygenase-2 overexpression after focal cerebral ischemia in rats. FASEB J. 2006, 20, 1162–1175. [Google Scholar] [CrossRef] [Green Version]
  557. Chan, S.H.; Wu, K.L.; Kung, P.S.; Chan, J.Y. Oral intake of rosiglitazone promotes a central antihypertensive effect via upregulation of peroxisome proliferator-activated receptor-gamma and alleviation of oxidative stress in rostral ventrolateral medulla of spontaneously hypertensive rats. Hypertension 2010, 55, 1444–1453. [Google Scholar] [CrossRef] [Green Version]
  558. Paradis, E.; Clavel, S.; Bouillaud, F.; Ricquier, D.; Richard, D. Uncoupling protein 2: A novel player in neuroprotection. Trends Mol. Med. 2003, 9, 522–525. [Google Scholar]
  559. Nagy, L.; Tontonoz, P.; Alvarez, J.G.; Chen, H.; Evans, R.M. Oxidized LDL regulates macrophage gene expression through ligand activation of PPARgamma. Cell 1998, 93, 229–240. [Google Scholar] [CrossRef] [Green Version]
  560. Ishii, T.; Itoh, K.; Ruiz, E.; Leake, D.S.; Unoki, H.; Yamamoto, M.; Mann, G.E. Role of Nrf2 in the regulation of CD36 and stress protein expression in murine macrophages: Activation by oxidatively modified LDL and 4-hydroxynonenal. Circ. Res. 2004, 94, 609–616. [Google Scholar] [CrossRef] [Green Version]
  561. Febbraio, M.; Hajjar, D.P.; Silverstein, R.L. CD36: A class B scavenger receptor involved in angiogenesis, atherosclerosis, inflammation, and lipid metabolism. J. Clin. Investig. 2001, 108, 785–791. [Google Scholar] [CrossRef]
  562. Fuenzalida, K.; Quintanilla, R.; Ramos, P.; Piderit, D.; Fuentealba, R.A.; Martinez, G.; Inestrosa, N.C.; Bronfman, M. Peroxisome proliferator-activated receptor gamma up-regulates the Bcl-2 anti-apoptotic protein in neurons and induces mitochondrial stabilization and protection against oxidative stress and apoptosis. J. Biol. Chem. 2007, 282, 37006–37015. [Google Scholar] [CrossRef] [Green Version]
  563. Ren, Y.; Sun, C.; Sun, Y.; Tan, H.; Wu, Y.; Cui, B.; Wu, Z. PPAR gamma protects cardiomyocytes against oxidative stress and apoptosis via Bcl-2 upregulation. Vasc. Pharmacol. 2009, 51, 169–174. [Google Scholar] [CrossRef]
  564. Chung, S.W.; Kang, B.Y.; Kim, S.H.; Pak, Y.K.; Cho, D.; Trinchieri, G.; Kim, T.S. Oxidized low density lipoprotein inhibits interleukin-12 production in lipopolysaccharide-activated mouse macrophages via direct interactions between peroxisome proliferator-activated receptor-gamma and nuclear factor-kappa B. J. Biol. Chem. 2000, 275, 32681–32687. [Google Scholar] [CrossRef] [Green Version]
  565. Liu, J.; Xia, Q.; Zhang, Q.; Li, H.; Zhang, J.; Li, A.; Xiu, R. Peroxisome proliferator-activated receptor-gamma ligands 15-deoxy-delta(12,14)-prostaglandin J2 and pioglitazone inhibit hydroxyl peroxide-induced TNF-alpha and lipopolysaccharide-induced CXC chemokine expression in neonatal rat cardiac myocytes. Shock 2009, 32, 317–324. [Google Scholar] [CrossRef]
  566. Goto, M.; Katayama, K.I.; Shirakawa, F.; Tanaka, I. Involvement of NF-kappaB p50/p65 heterodimer in activation of the human pro-interleukin-1beta gene at two subregions of the upstream enhancer element. Cytokine 1999, 11, 16–28. [Google Scholar] [CrossRef]
  567. Kim, H.J.; Kim, K.W.; Yu, B.P.; Chung, H.Y. The effect of age on cyclooxygenase-2 gene expression: NF-kappaB activation and IkappaBalpha degradation. Free Radic. Biol. Med. 2000, 28, 683–692. [Google Scholar] [CrossRef]
  568. Libermann, T.A.; Baltimore, D. Activation of interleukin-6 gene expression through the NF-kappa B transcription factor. Mol. Cell Biol. 1990, 10, 2327–2334. [Google Scholar] [CrossRef]
  569. Yao, J.; Mackman, N.; Edgington, T.S.; Fan, S.T. Lipopolysaccharide induction of the tumor necrosis factor-alpha promoter in human monocytic cells. Regulation by Egr-1, c-Jun, and NF-kappaB transcription factors. J. Biol. Chem. 1997, 272, 17795–17801. [Google Scholar] [CrossRef] [Green Version]
  570. Taylor, B.S.; de Vera, M.E.; Ganster, R.W.; Wang, Q.; Shapiro, R.A.; Morris, S.M., Jr.; Billiar, T.R.; Geller, D.A. Multiple NF-kappaB enhancer elements regulate cytokine induction of the human inducible nitric oxide synthase gene. J. Biol. Chem. 1998, 273, 15148–15156. [Google Scholar] [CrossRef] [Green Version]
  571. Oliveira-Marques, V.; Marinho, H.S.; Cyrne, L.; Antunes, F. Role of hydrogen peroxide in NF-kappaB activation: From inducer to modulator. Antioxid. Redox Signal. 2009, 11, 2223–2243. [Google Scholar] [CrossRef]
  572. Ye, J. Regulation of PPARgamma function by TNF-alpha. Biochem. Biophys. Res. Commun. 2008, 374, 405–408. [Google Scholar] [CrossRef] [Green Version]
  573. Napolitano, M.; Costa, L.; Palermo, R.; Giovenco, A.; Vacca, A.; Gulino, A. Protective effect of pioglitazone, a PPARgamma ligand, in a 3 nitropropionic acid model of Huntington’s disease. Brain Res. Bull. 2011, 85, 231–237. [Google Scholar] [CrossRef]
  574. Osburn, W.O.; Kensler, T.W. Nrf2 signaling: An adaptive response pathway for protection against environmental toxic insults. Mutat. Res. 2008, 659, 31–39. [Google Scholar] [CrossRef] [Green Version]
  575. Lee, J.M.; Calkins, M.J.; Chan, K.; Kan, Y.W.; Johnson, J.A. Identification of the NF-E2-related factor-2-dependent genes conferring protection against oxidative stress in primary cortical astrocytes using oligonucleotide microarray analysis. J. Biol. Chem. 2003, 278, 12029–12038. [Google Scholar] [CrossRef] [Green Version]
  576. Park, E.Y.; Cho, I.J.; Kim, S.G. Transactivation of the PPAR-responsive enhancer module in chemopreventive glutathione S-transferase gene by the peroxisome proliferator-activated receptor-gamma and retinoid X receptor heterodimer. Cancer Res. 2004, 64, 3701–3713. [Google Scholar] [CrossRef] [Green Version]
  577. D’Archivio, M.; Scazzocchio, B.; Filesi, C.; Vari, R.; Maggiorella, M.T.; Sernicola, L.; Santangelo, C.; Giovannini, C.; Masella, R. Oxidised LDL up-regulate CD36 expression by the Nrf2 pathway in 3T3-L1 preadipocytes. FEBS Lett. 2008, 582, 2291–2298. [Google Scholar] [CrossRef] [Green Version]
  578. Gong, P.; Stewart, D.; Hu, B.; Li, N.; Cook, J.; Nel, A.; Alam, J. Activation of the mouse heme oxygenase-1 gene by 15-deoxy-Delta(12,14)-prostaglandin J(2) is mediated by the stress response elements and transcription factor Nrf2. Antioxid. Redox Signal. 2002, 4, 249–257. [Google Scholar] [CrossRef]
  579. Nakaso, K.; Yano, H.; Fukuhara, Y.; Takeshima, T.; Wada-Isoe, K.; Nakashima, K. PI3K is a key molecule in the Nrf2-mediated regulation of antioxidative proteins by hemin in human neuroblastoma cells. FEBS Lett. 2003, 546, 181–184. [Google Scholar] [CrossRef] [Green Version]
  580. 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] [Green Version]
  581. Cho, H.Y.; Gladwell, W.; Wang, X.; Chorley, B.; Bell, D.; Reddy, S.P.; Kleeberger, S.R. Nrf2-regulated PPAR{gamma} expression is critical to protection against acute lung injury in mice. Am. J. Respir. Crit. Care Med. 2010, 182, 170–182. [Google Scholar] [CrossRef]
  582. Chan, D.C. Mitochondria: Dynamic organelles in disease, aging, and development. Cell 2006, 125, 1241–1252. [Google Scholar] [CrossRef] [Green Version]
  583. Kregel, K.C.; Zhang, H.J. An integrated view of oxidative stress in aging: Basic mechanisms, functional effects, and pathological considerations. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2007, 292, R18–R36. [Google Scholar] [CrossRef]
  584. Short, K.R.; Bigelow, M.L.; Kahl, J.; Singh, R.; Coenen-Schimke, J.; Raghavakaimal, S.; Nair, K.S. Decline in skeletal muscle mitochondrial function with aging in humans. Proc. Natl. Acad. Sci. USA 2005, 102, 5618–5623. [Google Scholar] [CrossRef] [Green Version]
  585. Chen, J.C.; Warshaw, J.B.; Sanadi, D.R. Regulation of mitochondrial respiration in senescence. J. Cell Physiol. 1972, 80, 141–148. [Google Scholar] [CrossRef]
  586. Hansford, R.G. Lipid oxidation by heart mitochondria from young adult and senescent rats. Biochem. J. 1978, 170, 285–295. [Google Scholar] [CrossRef] [Green Version]
  587. Trounce, I.; Byrne, E.; Marzuki, S. Decline in skeletal muscle mitochondrial respiratory chain function: Possible factor in ageing. Lancet 1989, 1, 637–639. [Google Scholar]
  588. Cooper, J.M.; Mann, V.M.; Schapira, A.H. Analyses of mitochondrial respiratory chain function and mitochondrial DNA deletion in human skeletal muscle: Effect of ageing. J. Neurol. Sci. 1992, 113, 91–98. [Google Scholar]
  589. Boffoli, D.; Scacco, S.C.; Vergari, R.; Solarino, G.; Santacroce, G.; Papa, S. Decline with age of the respiratory chain activity in human skeletal muscle. Biochim. Biophys. Acta 1994, 1226, 73–82. [Google Scholar]
  590. Linnane, A.W.; Marzuki, S.; Ozawa, T.; Tanaka, M. Mitochondrial DNA mutations as an important contributor to ageing and degenerative diseases. Lancet 1989, 1, 642–645. [Google Scholar]
  591. Hancock, C.R.; Han, D.H.; Higashida, K.; Kim, S.H.; Holloszy, J.O. Does calorie restriction induce mitochondrial biogenesis? A reevaluation. FASEB J. 2011, 25, 785–791. [Google Scholar] [CrossRef] [Green Version]
  592. Zhu, M.; de Cabo, R.; Lane, M.A.; Ingram, D.K. Caloric restriction modulates early events in insulin signaling in liver and skeletal muscle of rat. Ann. N. Y. Acad. Sci. 2004, 1019, 448–452. [Google Scholar] [CrossRef]
  593. Ranhotra, H.S. Long-term caloric restriction up-regulates PPAR gamma co-activator 1 alpha (PGC-1alpha) expression in mice. Indian J. Biochem. Biophys. 2010, 47, 272–277. [Google Scholar]
  594. Puigserver, P.; Wu, Z.; Park, C.W.; Graves, R.; Wright, M.; Spiegelman, B.M. A cold-inducible coactivator of nuclear receptors linked to adaptive thermogenesis. Cell 1998, 92, 829–839. [Google Scholar]
  595. Handschin, C.; Spiegelman, B.M. Peroxisome proliferator-activated receptor gamma coactivator 1 coactivators, energy homeostasis, and metabolism. Endocr. Rev. 2006, 27, 728–735. [Google Scholar] [CrossRef]
  596. Wu, Z.; Puigserver, P.; Andersson, U.; Zhang, C.; Adelmant, G.; Mootha, V.; Troy, A.; Cinti, S.; Lowell, B.; Scarpulla, R.C.; et al. Mechanisms controlling mitochondrial biogenesis and respiration through the thermogenic coactivator PGC-1. Cell 1999, 98, 115–124. [Google Scholar] [CrossRef] [Green Version]
  597. St-Pierre, J.; Drori, S.; Uldry, M.; Silvaggi, J.M.; Rhee, J.; Jager, S.; Handschin, C.; Zheng, K.; Lin, J.; Yang, W.; et al. Suppression of reactive oxygen species and neurodegeneration by the PGC-1 transcriptional coactivators. Cell 2006, 127, 397–408. [Google Scholar] [CrossRef] [Green Version]
  598. Lanza, I.R.; Zabielski, P.; Klaus, K.A.; Morse, D.M.; Heppelmann, C.J.; Bergen, H.R., 3rd; Dasari, S.; Walrand, S.; Short, K.R.; Johnson, M.L.; et al. Chronic caloric restriction preserves mitochondrial function in senescence without increasing mitochondrial biogenesis. Cell Metab. 2012, 16, 777–788. [Google Scholar] [CrossRef] [Green Version]
  599. Hepple, R.T.; Baker, D.J.; McConkey, M.; Murynka, T.; Norris, R. Caloric restriction protects mitochondrial function with aging in skeletal and cardiac muscles. Rejuvenation Res. 2006, 9, 219–222. [Google Scholar] [CrossRef]
  600. Miller, B.F.; Robinson, M.M.; Bruss, M.D.; Hellerstein, M.; Hamilton, K.L. A comprehensive assessment of mitochondrial protein synthesis and cellular proliferation with age and caloric restriction. Aging Cell 2012, 11, 150–161. [Google Scholar] [CrossRef] [Green Version]
  601. Lopez-Lluch, G.; Hunt, N.; Jones, B.; Zhu, M.; Jamieson, H.; Hilmer, S.; Cascajo, M.V.; Allard, J.; Ingram, D.K.; Navas, P.; et al. Calorie restriction induces mitochondrial biogenesis and bioenergetic efficiency. Proc. Natl. Acad. Sci. USA 2006, 103, 1768–1773. [Google Scholar] [CrossRef] [Green Version]
  602. Gudiksen, A.; Pilegaard, H. PGC-1alpha and fasting-induced PDH regulation in mouse skeletal muscle. Physiol. Rep. 2017, 5. [Google Scholar] [CrossRef]
  603. Geng, T.; Li, P.; Okutsu, M.; Yin, X.; Kwek, J.; Zhang, M.; Yan, Z. PGC-1alpha plays a functional role in exercise-induced mitochondrial biogenesis and angiogenesis but not fiber-type transformation in mouse skeletal muscle. Am. J. Physiol. Cell Physiol. 2010, 298, C572–C579. [Google Scholar] [CrossRef] [Green Version]
  604. Leick, L.; Wojtaszewski, J.F.; Johansen, S.T.; Kiilerich, K.; Comes, G.; Hellsten, Y.; Hidalgo, J.; Pilegaard, H. PGC-1alpha is not mandatory for exercise- and training-induced adaptive gene responses in mouse skeletal muscle. Am. J. Physiol. Endocrinol. Metab. 2008, 294, E463–E474. [Google Scholar] [CrossRef] [Green Version]
  605. Vega, R.B.; Huss, J.M.; Kelly, D.P. The coactivator PGC-1 cooperates with peroxisome proliferator-activated receptor alpha in transcriptional control of nuclear genes encoding mitochondrial fatty acid oxidation enzymes. Mol. Cell Biol. 2000, 20, 1868–1876. [Google Scholar] [CrossRef] [Green Version]
  606. Perez-Schindler, J.; Summermatter, S.; Salatino, S.; Zorzato, F.; Beer, M.; Balwierz, P.J.; van Nimwegen, E.; Feige, J.N.; Auwerx, J.; Handschin, C. The corepressor NCoR1 antagonizes PGC-1alpha and estrogen-related receptor alpha in the regulation of skeletal muscle function and oxidative metabolism. Mol. Cell Biol. 2012, 32, 4913–4924. [Google Scholar] [CrossRef] [Green Version]
  607. Zolezzi, J.M.; Silva-Alvarez, C.; Ordenes, D.; Godoy, J.A.; Carvajal, F.J.; Santos, M.J.; Inestrosa, N.C. Peroxisome proliferator-activated receptor (PPAR) gamma and PPARalpha agonists modulate mitochondrial fusion-fission dynamics: Relevance to reactive oxygen species (ROS)-related neurodegenerative disorders? PLoS ONE 2013, 8, e64019. [Google Scholar] [CrossRef] [Green Version]
  608. Watanabe, K.; Fujii, H.; Takahashi, T.; Kodama, M.; Aizawa, Y.; Ohta, Y.; Ono, T.; Hasegawa, G.; Naito, M.; Nakajima, T.; et al. Constitutive regulation of cardiac fatty acid metabolism through peroxisome proliferator-activated receptor alpha associated with age-dependent cardiac toxicity. J. Biol. Chem. 2000, 275, 22293–22299. [Google Scholar] [CrossRef] [Green Version]
  609. Minnich, A.; Tian, N.; Byan, L.; Bilder, G. A potent PPARalpha agonist stimulates mitochondrial fatty acid beta-oxidation in liver and skeletal muscle. Am. J. Physiol. Endocrinol. Metab. 2001, 280, E270–E279. [Google Scholar] [CrossRef] [Green Version]
  610. 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 PPARalpha. Nat. Commun. 2019, 10, 1566. [Google Scholar] [CrossRef] [Green Version]
  611. Cree, M.G.; Zwetsloot, J.J.; Herndon, D.N.; Qian, T.; Morio, B.; Fram, R.; Sanford, A.P.; Aarsland, A.; Wolfe, R.R. Insulin sensitivity and mitochondrial function are improved in children with burn injury during a randomized controlled trial of fenofibrate. Ann. Surg. 2007, 245, 214–221. [Google Scholar] [CrossRef]
  612. Hong, M.; Song, K.D.; Lee, H.K.; Yi, S.; Lee, Y.S.; Heo, T.H.; Jun, H.S.; Kim, S.J. Fibrates inhibit the apoptosis of Batten disease lymphoblast cells via autophagy recovery and regulation of mitochondrial membrane potential. In Vitro Cell Dev. Biol. Anim. 2016, 52, 349–355. [Google Scholar] [CrossRef]
  613. Mohagheghi, F.; Ahmadiani, A.; Rahmani, B.; Moradi, F.; Romond, N.; Khalaj, L. Gemfibrozil pretreatment resulted in a sexually dimorphic outcome in the rat models of global cerebral ischemia-reperfusion via modulation of mitochondrial pro-survival and apoptotic cell death factors as well as MAPKs. J. Mol. Neurosci. 2013, 50, 379–393. [Google Scholar] [CrossRef]
  614. Brunmair, B.; Lest, A.; Staniek, K.; Gras, F.; Scharf, N.; Roden, M.; Nohl, H.; Waldhausl, W.; Furnsinn, C. Fenofibrate impairs rat mitochondrial function by inhibition of respiratory complex I. J. Pharmacol. Exp. Ther. 2004, 311, 109–114. [Google Scholar] [CrossRef] [Green Version]
  615. Zhou, S.; Wallace, K.B. The effect of peroxisome proliferators on mitochondrial bioenergetics. Toxicol. Sci. 1999, 48, 82–89. [Google Scholar] [CrossRef] [Green Version]
  616. Zungu, M.; Young, M.E.; Stanley, W.C.; Essop, M.F. Chronic treatment with the peroxisome proliferator-activated receptor alpha agonist Wy-14,643 attenuates myocardial respiratory capacity and contractile function. Mol. Cell Biochem. 2009, 330, 55–62. [Google Scholar] [CrossRef]
  617. Koh, J.H.; Hancock, C.R.; Terada, S.; Higashida, K.; Holloszy, J.O.; Han, D.H. PPARbeta Is Essential for Maintaining Normal Levels of PGC-1alpha and Mitochondria and for the Increase in Muscle Mitochondria Induced by Exercise. Cell Metab. 2017, 25, 1176–1185. [Google Scholar] [CrossRef]
  618. Dressel, U.; Allen, T.L.; Pippal, J.B.; Rohde, P.R.; Lau, P.; Muscat, G.E. The peroxisome proliferator-activated receptor beta/delta agonist, GW501516, regulates the expression of genes involved in lipid catabolism and energy uncoupling in skeletal muscle cells. Mol. Endocrinol. 2003, 17, 2477–2493. [Google Scholar] [CrossRef] [Green Version]
  619. Dulloo, A.G.; Samec, S. Uncoupling proteins: Their roles in adaptive thermogenesis and substrate metabolism reconsidered. Br. J. Nutr. 2001, 86, 123–139. [Google Scholar] [CrossRef] [Green Version]
  620. Luquet, S.; Lopez-Soriano, J.; Holst, D.; Fredenrich, A.; Melki, J.; Rassoulzadegan, M.; Grimaldi, P.A. Peroxisome proliferator-activated receptor delta controls muscle development and oxidative capability. FASEB J. 2003, 17, 2299–2301. [Google Scholar] [CrossRef]
  621. Lin, J.; Wu, H.; Tarr, P.T.; Zhang, C.Y.; Wu, Z.; Boss, O.; Michael, L.F.; Puigserver, P.; Isotani, E.; Olson, E.N.; et al. Transcriptional co-activator PGC-1 alpha drives the formation of slow-twitch muscle fibres. Nature 2002, 418, 797–801. [Google Scholar] [CrossRef]
  622. Kleiner, S.; Nguyen-Tran, V.; Bare, O.; Huang, X.; Spiegelman, B.; Wu, Z. PPAR{delta} agonism activates fatty acid oxidation via PGC-1{alpha} but does not increase mitochondrial gene expression and function. J. Biol. Chem. 2009, 284, 18624–18633. [Google Scholar] [CrossRef] [Green Version]
  623. Fan, W.; Waizenegger, W.; Lin, C.S.; Sorrentino, V.; He, M.X.; Wall, C.E.; Li, H.; Liddle, C.; Yu, R.T.; Atkins, A.R.; et al. PPARdelta Promotes Running Endurance by Preserving Glucose. Cell Metab. 2017, 25, 1186–1193. [Google Scholar] [CrossRef]
  624. Zhou, Y.; Yang, J.; Huang, J.; Li, T.; Xu, D.; Zuo, B.; Hou, L.; Wu, W.; Zhang, L.; Xia, X.; et al. The formation of brown adipose tissue induced by transgenic over-expression of PPARgamma2. Biochem. Biophys. Res. Commun. 2014, 446, 959–964. [Google Scholar] [CrossRef]
  625. Son, N.H.; Park, T.S.; Yamashita, H.; Yokoyama, M.; Huggins, L.A.; Okajima, K.; Homma, S.; Szabolcs, M.J.; Huang, L.S.; Goldberg, I.J. Cardiomyocyte expression of PPARgamma leads to cardiac dysfunction in mice. J. Clin. Investig. 2007, 117, 2791–2801. [Google Scholar] [CrossRef] [Green Version]
  626. Wilson-Fritch, L.; Nicoloro, S.; Chouinard, M.; Lazar, M.A.; Chui, P.C.; Leszyk, J.; Straubhaar, J.; Czech, M.P.; Corvera, S. Mitochondrial remodeling in adipose tissue associated with obesity and treatment with rosiglitazone. J. Clin. Investig. 2004, 114, 1281–1289. [Google Scholar] [CrossRef]
  627. Hakansson, J.; Eliasson, B.; Smith, U.; Enerback, S. Adipocyte mitochondrial genes and the forkhead factor FOXC2 are decreased in type 2 diabetes patients and normalized in response to rosiglitazone. Diabetol. Metab. Syndr. 2011, 3, 32. [Google Scholar] [CrossRef] [Green Version]
  628. Bogacka, I.; Xie, H.; Bray, G.A.; Smith, S.R. Pioglitazone induces mitochondrial biogenesis in human subcutaneous adipose tissue in vivo. Diabetes 2005, 54, 1392–1399. [Google Scholar] [CrossRef] [Green Version]
  629. Rong, J.X.; Klein, J.L.; Qiu, Y.; Xie, M.; Johnson, J.H.; Waters, K.M.; Zhang, V.; Kashatus, J.A.; Remlinger, K.S.; Bing, N.; et al. Rosiglitazone Induces Mitochondrial Biogenesis in Differentiated Murine 3T3-L1 and C3H/10T1/2 Adipocytes. PPAR Res. 2011, 2011, 179454. [Google Scholar] [CrossRef] [Green Version]
  630. Wang, Y.L.; Frauwirth, K.A.; Rangwala, S.M.; Lazar, M.A.; Thompson, C.B. Thiazolidinedione activation of peroxisome proliferator-activated receptor gamma can enhance mitochondrial potential and promote cell survival. J. Biol. Chem. 2002, 277, 31781–31788. [Google Scholar] [CrossRef] [Green Version]
  631. Chiang, M.C.; Chern, Y.; Huang, R.N. PPARgamma rescue of the mitochondrial dysfunction in Huntington’s disease. Neurobiol. Dis. 2012, 45, 322–328. [Google Scholar] [CrossRef]
  632. Dello Russo, C.; Gavrilyuk, V.; Weinberg, G.; Almeida, A.; Bolanos, J.P.; Palmer, J.; Pelligrino, D.; Galea, E.; Feinstein, D.L. Peroxisome proliferator-activated receptor gamma thiazolidinedione agonists increase glucose metabolism in astrocytes. J. Biol. Chem. 2003, 278, 5828–5836. [Google Scholar] [CrossRef] [Green Version]
  633. Miglio, G.; Rosa, A.C.; Rattazzi, L.; Collino, M.; Lombardi, G.; Fantozzi, R. PPARgamma stimulation promotes mitochondrial biogenesis and prevents glucose deprivation-induced neuronal cell loss. Neurochem. Int. 2009, 55, 496–504. [Google Scholar] [CrossRef]
  634. Ghosh, S.; Patel, N.; Rahn, D.; McAllister, J.; Sadeghi, S.; Horwitz, G.; Berry, D.; Wang, K.X.; Swerdlow, R.H. The thiazolidinedione pioglitazone alters mitochondrial function in human neuron-like cells. Mol. Pharmacol. 2007, 71, 1695–1702. [Google Scholar] [CrossRef] [Green Version]
  635. Feinstein, D.L.; Spagnolo, A.; Akar, C.; Weinberg, G.; Murphy, P.; Gavrilyuk, V.; Dello Russo, C. Receptor-independent actions of PPAR thiazolidinedione agonists: Is mitochondrial function the key? Biochem. Pharmacol. 2005, 70, 177–188. [Google Scholar] [CrossRef]
  636. Colca, J.R.; Tanis, S.P.; McDonald, W.G.; Kletzien, R.F. Insulin sensitizers in 2013: New insights for the development of novel therapeutic agents to treat metabolic diseases. Expert Opin. Investig. Drugs 2014, 23, 1–7. [Google Scholar] [CrossRef] [Green Version]
  637. Hobbs, M.V.; Weigle, W.O.; Noonan, D.J.; Torbett, B.E.; McEvilly, R.J.; Koch, R.J.; Cardenas, G.J.; Ernst, D.N. Patterns of cytokine gene expression by CD4+ T cells from young and old mice. J. Immunol. 1993, 150, 3602–3614. [Google Scholar]
  638. Riancho, J.A.; Zarrabeitia, M.T.; Amado, J.A.; Olmos, J.M.; Gonzalez-Macias, J. Age-related differences in cytokine secretion. Gerontology 1994, 40, 8–12. [Google Scholar] [CrossRef]
  639. Miller, R.A. Aging and immune function. Int. Rev. Cytol. 1991, 124, 187–215. [Google Scholar] [CrossRef]
  640. Kubo, M.; Cinader, B. Polymorphism of age-related changes in interleukin (IL) production: Differential changes of T helper subpopulations, synthesizing IL 2, IL 3 and IL 4. Eur. J. Immunol. 1990, 20, 1289–1296. [Google Scholar] [CrossRef]
  641. Hayek, M.G.; Mura, C.; Wu, D.; Beharka, A.A.; Han, S.N.; Paulson, K.E.; Hwang, D.; Meydani, S.N. Enhanced expression of inducible cyclooxygenase with age in murine macrophages. J. Immunol. 1997, 159, 2445–2451. [Google Scholar]
  642. Erol, A. Interleukin-6 (IL-6) is still the leading biomarker of the metabolic and aging related disorders. Med. Hypotheses 2007, 69, 708. [Google Scholar] [CrossRef]
  643. Ricote, M.; Huang, J.; Fajas, L.; Li, A.; Welch, J.; Najib, J.; Witztum, J.L.; Auwerx, J.; Palinski, W.; Glass, C.K. Expression of the peroxisome proliferator-activated receptor gamma (PPARgamma) in human atherosclerosis and regulation in macrophages by colony stimulating factors and oxidized low density lipoprotein. Proc. Natl. Acad. Sci. USA 1998, 95, 7614–7619. [Google Scholar]
  644. Jiang, C.; Ting, A.T.; Seed, B. PPAR-gamma agonists inhibit production of monocyte inflammatory cytokines. Nature 1998, 391, 82–86. [Google Scholar] [CrossRef]
  645. Okada, M.; Yan, S.F.; Pinsky, D.J. Peroxisome proliferator-activated receptor-gamma (PPAR-gamma) activation suppresses ischemic induction of Egr-1 and its inflammatory gene targets. FASEB J. 2002, 16, 1861–1868. [Google Scholar] [CrossRef]
  646. Barbier, O.; Torra, I.P.; Duguay, Y.; Blanquart, C.; Fruchart, J.C.; Glineur, C.; Staels, B. Pleiotropic actions of peroxisome proliferator-activated receptors in lipid metabolism and atherosclerosis. Arterioscler. Thromb. Vasc. Biol. 2002, 22, 717–726. [Google Scholar]
  647. Narala, V.R.; Adapala, R.K.; Suresh, M.V.; Brock, T.G.; Peters-Golden, M.; Reddy, R.C. Leukotriene B4 is a physiologically relevant endogenous peroxisome proliferator-activated receptor-alpha agonist. J. Biol. Chem. 2010, 285, 22067–22074. [Google Scholar] [CrossRef] [Green Version]
  648. Delerive, P.; De Bosscher, K.; Besnard, S.; Vanden Berghe, W.; Peters, J.M.; Gonzalez, F.J.; Fruchart, J.C.; Tedgui, A.; Haegeman, G.; Staels, B. Peroxisome proliferator-activated receptor alpha negatively regulates the vascular inflammatory gene response by negative cross-talk with transcription factors NF-kappaB and AP-1. J. Biol. Chem. 1999, 274, 32048–32054. [Google Scholar]
  649. Crisafulli, C.; Cuzzocrea, S. The role of endogenous and exogenous ligands for the peroxisome proliferator-activated receptor alpha (PPAR-alpha) in the regulation of inflammation in macrophages. Shock 2009, 32, 62–73. [Google Scholar] [CrossRef]
  650. Peters, J.M.; Hollingshead, H.E.; Gonzalez, F.J. Role of peroxisome-proliferator-activated receptor beta/delta (PPARbeta/delta) in gastrointestinal tract function and disease. Clin. Sci. (Lond.) 2008, 115, 107–127. [Google Scholar] [CrossRef] [Green Version]
  651. Ding, G.; Cheng, L.; Qin, Q.; Frontin, S.; Yang, Q. PPARdelta modulates lipopolysaccharide-induced TNFalpha inflammation signaling in cultured cardiomyocytes. J. Mol. Cell Cardiol. 2006, 40, 821–828. [Google Scholar] [CrossRef]
  652. Iwashita, A.; Muramatsu, Y.; Yamazaki, T.; Muramoto, M.; Kita, Y.; Yamazaki, S.; Mihara, K.; Moriguchi, A.; Matsuoka, N. Neuroprotective efficacy of the peroxisome proliferator-activated receptor delta-selective agonists in vitro and in vivo. J. Pharmacol. Exp. Ther. 2007, 320, 1087–1096. [Google Scholar] [CrossRef]
  653. Benetti, E.; Mastrocola, R.; Rogazzo, M.; Chiazza, F.; Aragno, M.; Fantozzi, R.; Collino, M.; Minetto, M.A. High sugar intake and development of skeletal muscle insulin resistance and inflammation in mice: A protective role for PPAR- delta agonism. Mediat. Inflamm. 2013, 2013, 509502. [Google Scholar] [CrossRef] [Green Version]
  654. Lewis, J.D.; Lichtenstein, G.R.; Deren, J.J.; Sands, B.E.; Hanauer, S.B.; Katz, J.A.; Lashner, B.; Present, D.H.; Chuai, S.; Ellenberg, J.H.; et al. Rosiglitazone for active ulcerative colitis: A randomized placebo-controlled trial. Gastroenterology 2008, 134, 688–695. [Google Scholar] [CrossRef] [Green Version]
  655. Lewis, J.D.; Lichtenstein, G.R.; Stein, R.B.; Deren, J.J.; Judge, T.A.; Fogt, F.; Furth, E.E.; Demissie, E.J.; Hurd, L.B.; Su, C.G.; et al. An open-label trial of the PPAR-gamma ligand rosiglitazone for active ulcerative colitis. Am. J. Gastroenterol. 2001, 96, 3323–3328. [Google Scholar] [CrossRef]
  656. Liang, H.L.; Ouyang, Q. A clinical trial of combined use of rosiglitazone and 5-aminosalicylate for ulcerative colitis. World J. Gastroenterol. WJG 2008, 14, 114–119. [Google Scholar]
  657. Bassaganya-Riera, J.; Hontecillas, R. CLA and n-3 PUFA differentially modulate clinical activity and colonic PPAR-responsive gene expression in a pig model of experimental IBD. Clin. Nutr. 2006, 25, 454–465. [Google Scholar] [CrossRef]
  658. Hontecillas, R.; Wannemeulher, M.J.; Zimmerman, D.R.; Hutto, D.L.; Wilson, J.H.; Ahn, D.U.; Bassaganya-Riera, J. Nutritional regulation of porcine bacterial-induced colitis by conjugated linoleic acid. J. Nutr. 2002, 132, 2019–2027. [Google Scholar]
  659. Sanchez-Hidalgo, M.; Martin, A.R.; Villegas, I.; de la Lastra, C.A. Rosiglitazone, a PPARgamma ligand, modulates signal transduction pathways during the development of acute TNBS-induced colitis in rats. Eur. J. Pharmacol. 2007, 562, 247–258. [Google Scholar] [CrossRef]
  660. Sato, N.; Kozar, R.A.; Zou, L.; Weatherall, J.M.; Attuwaybi, B.; Moore-Olufemi, S.D.; Weisbrodt, N.W.; Moore, F.A. Peroxisome proliferator-activated receptor gamma mediates protection against cyclooxygenase-2-induced gut dysfunction in a rodent model of mesenteric ischemia/reperfusion. Shock 2005, 24, 462–469. [Google Scholar]
  661. Saubermann, L.J.; Nakajima, A.; Wada, K.; Zhao, S.; Terauchi, Y.; Kadowaki, T.; Aburatani, H.; Matsuhashi, N.; Nagai, R.; Blumberg, R.S. Peroxisome proliferator-activated receptor gamma agonist ligands stimulate a Th2 cytokine response and prevent acute colitis. Inflamm. Bowel Dis. 2002, 8, 330–339. [Google Scholar]
  662. Su, C.G.; Wen, X.; Bailey, S.T.; Jiang, W.; Rangwala, S.M.; Keilbaugh, S.A.; Flanigan, A.; Murthy, S.; Lazar, M.A.; Wu, G.D. A novel therapy for colitis utilizing PPAR-gamma ligands to inhibit the epithelial inflammatory response. J. Clin. Investig. 1999, 104, 383–389. [Google Scholar] [CrossRef] [Green Version]
  663. Kundu, P.; Ling, T.W.; Korecka, A.; Li, Y.; D’Arienzo, R.; Bunte, R.M.; Berger, T.; Arulampalam, V.; Chambon, P.; Mak, T.W.; et al. Absence of intestinal PPARgamma aggravates acute infectious colitis in mice through a lipocalin-2-dependent pathway. PLoS Pathog 2014, 10, e1003887. [Google Scholar] [CrossRef] [Green Version]
  664. Michalik, L.; Wahli, W. PPARs Mediate Lipid Signaling in Inflammation and Cancer. PPAR Res. 2008, 2008, 134059. [Google Scholar] [CrossRef] [Green Version]
  665. Shah, Y.M.; Morimura, K.; Gonzalez, F.J. Expression of peroxisome proliferator-activated receptor-gamma in macrophage suppresses experimentally induced colitis. Am. J. Physiol. Gastrointest. Liver Physiol. 2007, 292, G657–G666. [Google Scholar] [CrossRef] [Green Version]
  666. Leung, E.; Hong, J.; Fraser, A.; Merriman, T.; Krissansen, G. PPAR-gamma and Crohn’s disease in New Zealand. Gastroenterology 2006, 130, 2249–2250. [Google Scholar] [CrossRef]
  667. Wada, K.; Nakajima, A.; Blumberg, R.S. PPARgamma and inflammatory bowel disease: A new therapeutic target for ulcerative colitis and Crohn’s disease. Trends Mol. Med. 2001, 7, 329–331. [Google Scholar]
  668. Pascual, G.; Fong, A.L.; Ogawa, S.; Gamliel, A.; Li, A.C.; Perissi, V.; Rose, D.W.; Willson, T.M.; Rosenfeld, M.G.; Glass, C.K. A SUMOylation-dependent pathway mediates transrepression of inflammatory response genes by PPAR-gamma. Nature 2005, 437, 759–763. [Google Scholar] [CrossRef]
  669. Chawla, A.; Barak, Y.; Nagy, L.; Liao, D.; Tontonoz, P.; Evans, R.M. PPAR-gamma dependent and independent effects on macrophage-gene expression in lipid metabolism and inflammation. Nat. Med. 2001, 7, 48–52. [Google Scholar] [CrossRef]
  670. Gosset, P.; Charbonnier, A.S.; Delerive, P.; Fontaine, J.; Staels, B.; Pestel, J.; Tonnel, A.B.; Trottein, F. Peroxisome proliferator-activated receptor gamma activators affect the maturation of human monocyte-derived dendritic cells. Eur. J. Immunol. 2001, 31, 2857–2865. [Google Scholar]
  671. Faveeuw, C.; Fougeray, S.; Angeli, V.; Fontaine, J.; Chinetti, G.; Gosset, P.; Delerive, P.; Maliszewski, C.; Capron, M.; Staels, B.; et al. Peroxisome proliferator-activated receptor gamma activators inhibit interleukin-12 production in murine dendritic cells. FEBS Lett. 2000, 486, 261–266. [Google Scholar]
  672. Odegaard, J.I.; Ricardo-Gonzalez, R.R.; Goforth, M.H.; Morel, C.R.; Subramanian, V.; Mukundan, L.; Red Eagle, A.; Vats, D.; Brombacher, F.; Ferrante, A.W.; et al. Macrophage-specific PPARgamma controls alternative activation and improves insulin resistance. Nature 2007, 447, 1116–1120. [Google Scholar] [CrossRef] [Green Version]
  673. Klotz, L.; Schmidt, S.; Heun, R.; Klockgether, T.; Kolsch, H. Association of the PPARgamma gene polymorphism Pro12Ala with delayed onset of multiple sclerosis. Neurosci. Lett. 2009, 449, 81–83. [Google Scholar] [CrossRef]
  674. Regieli, J.J.; Jukema, J.W.; Doevendans, P.A.; Zwinderman, A.H.; van der Graaf, Y.; Kastelein, J.J.; Grobbee, D.E. PPAR gamma variant influences angiographic outcome and 10-year cardiovascular risk in male symptomatic coronary artery disease patients. Diabetes Care 2009, 32, 839–844. [Google Scholar] [CrossRef] [Green Version]
  675. Jiang, J.; Xie, Z.; Guo, J.; Wang, Y.; Liu, C.; Zhang, S.; Tang, W.; Chen, Y. Association of PPARG rs 1801282 C>G polymorphism with risk of colorectal cancer: From a case-control study to a meta-analysis. Oncotarget 2017, 8, 100558–100569. [Google Scholar] [CrossRef] [Green Version]
  676. Lin, J.; Chen, Y.; Tang, W.F.; Liu, C.; Zhang, S.; Guo, Z.Q.; Chen, G.; Zheng, X.W. PPARG rs3856806 C>T Polymorphism Increased the Risk of Colorectal Cancer: A Case-Control Study in Eastern Chinese Han Population. Front. Oncol. 2019, 9, 63. [Google Scholar] [CrossRef] [Green Version]
  677. Bassaganya-Riera, J.; Reynolds, K.; Martino-Catt, S.; Cui, Y.; Hennighausen, L.; Gonzalez, F.; Rohrer, J.; Benninghoff, A.U.; Hontecillas, R. Activation of PPAR gamma and delta by conjugated linoleic acid mediates protection from experimental inflammatory bowel disease. Gastroenterology 2004, 127, 777–791. [Google Scholar]
  678. Yasui, Y.; Hosokawa, M.; Sahara, T.; Suzuki, R.; Ohgiya, S.; Kohno, H.; Tanaka, T.; Miyashita, K. Bitter gourd seed fatty acid rich in 9c,11t,13t-conjugated linolenic acid induces apoptosis and up-regulates the GADD45, p53 and PPARgamma in human colon cancer Caco-2 cells. Prostaglandins Leukot. Essent. Fatty Acids 2005, 73, 113–119. [Google Scholar] [CrossRef]
  679. Akinyeke, T.O.; Stewart, L.V. Troglitazone suppresses c-Myc levels in human prostate cancer cells via a PPARgamma-independent mechanism. Cancer Biol. Ther. 2011, 11, 1046–1058. [Google Scholar]
  680. Yang, W.L.; Frucht, H. Activation of the PPAR pathway induces apoptosis and COX-2 inhibition in HT-29 human colon cancer cells. Carcinogenesis 2001, 22, 1379–1383. [Google Scholar]
  681. Subbaramaiah, K.; Lin, D.T.; Hart, J.C.; Dannenberg, A.J. Peroxisome proliferator-activated receptor gamma ligands suppress the transcriptional activation of cyclooxygenase-2. Evidence for involvement of activator protein-1 and CREB-binding protein/p300. J. Biol. Chem. 2001, 276, 12440–12448. [Google Scholar] [CrossRef] [Green Version]
  682. Vandoros, G.P.; Konstantinopoulos, P.A.; Sotiropoulou-Bonikou, G.; Kominea, A.; Papachristou, G.I.; Karamouzis, M.V.; Gkermpesi, M.; Varakis, I.; Papavassiliou, A.G. PPAR-gamma is expressed and NF-kB pathway is activated and correlates positively with COX-2 expression in stromal myofibroblasts surrounding colon adenocarcinomas. J. Cancer Res. Clin. Oncol. 2006, 132, 76–84. [Google Scholar] [CrossRef]
  683. Heneka, M.T.; Klockgether, T.; Feinstein, D.L. Peroxisome proliferator-activated receptor-gamma ligands reduce neuronal inducible nitric oxide synthase expression and cell death in vivo. J. Neurosci. 2000, 20, 6862–6867. [Google Scholar]
  684. Girnun, G.D.; Smith, W.M.; Drori, S.; Sarraf, P.; Mueller, E.; Eng, C.; Nambiar, P.; Rosenberg, D.W.; Bronson, R.T.; Edelmann, W.; et al. APC-dependent suppression of colon carcinogenesis by PPARgamma. Proc. Natl. Acad. Sci. USA 2002, 99, 13771–13776. [Google Scholar] [CrossRef] [Green Version]
  685. Fujisawa, T.; Nakajima, A.; Fujisawa, N.; Takahashi, H.; Ikeda, I.; Tomimoto, A.; Yonemitsu, K.; Nakajima, N.; Kudo, C.; Wada, K.; et al. Peroxisome proliferator-activated receptor gamma (PPARgamma) suppresses colonic epithelial cell turnover and colon carcinogenesis through inhibition of the beta-catenin/T cell factor (TCF) pathway. J. Pharmacol. Sci. 2008, 106, 627–638. [Google Scholar]
  686. Peeters, A.; Baes, M. Role of PPAR in Hepatic Carbohydrate Metabolism. PPAR Res. 2010, 2010. [Google Scholar] [CrossRef] [Green Version]
  687. Xu, J.; Xiao, G.; Trujillo, C.; Chang, V.; Blanco, L.; Joseph, S.B.; Bassilian, S.; Saad, M.F.; Tontonoz, P.; Lee, W.N.; et al. Peroxisome proliferator-activated receptor alpha (PPARalpha) influences substrate utilization for hepatic glucose production. J. Biol. Chem. 2002, 277, 50237–50244. [Google Scholar] [CrossRef] [Green Version]
  688. Mandard, S.; Stienstra, R.; Escher, P.; Tan, N.S.; Kim, I.; Gonzalez, F.J.; Wahli, W.; Desvergne, B.; Muller, M.; Kersten, S. Glycogen synthase 2 is a novel target gene of peroxisome proliferator-activated receptors. Cell Mol. Life Sci. 2007, 64, 1145–1157. [Google Scholar] [CrossRef] [Green Version]
  689. Finck, B.N.; Lehman, J.J.; Leone, T.C.; Welch, M.J.; Bennett, M.J.; Kovacs, A.; Han, X.; Gross, R.W.; Kozak, R.; Lopaschuk, G.D.; et al. The cardiac phenotype induced by PPARalpha overexpression mimics that caused by diabetes mellitus. J. Clin. Investig. 2002, 109, 121–130. [Google Scholar] [CrossRef] [Green Version]
  690. Philp, A.; MacKenzie, M.G.; Belew, M.Y.; Towler, M.C.; Corstorphine, A.; Papalamprou, A.; Hardie, D.G.; Baar, K. Glycogen content regulates peroxisome proliferator activated receptor- partial differential (PPAR- partial differential) activity in rat skeletal muscle. PLoS ONE 2013, 8, e77200. [Google Scholar] [CrossRef] [Green Version]
  691. Rodriguez, J.C.; Gil-Gomez, G.; Hegardt, F.G.; Haro, D. Peroxisome proliferator-activated receptor mediates induction of the mitochondrial 3-hydroxy-3-methylglutaryl-CoA synthase gene by fatty acids. J. Biol. Chem. 1994, 269, 18767–18772. [Google Scholar]
  692. Rakhshandehroo, M.; Hooiveld, G.; Muller, M.; Kersten, S. Comparative analysis of gene regulation by the transcription factor PPARalpha between mouse and human. PLoS ONE 2009, 4, e6796. [Google Scholar] [CrossRef] [Green Version]
  693. Louet, J.F.; Chatelain, F.; Decaux, J.F.; Park, E.A.; Kohl, C.; Pineau, T.; Girard, J.; Pegorier, J.P. Long-chain fatty acids regulate liver carnitine palmitoyltransferase I gene (L-CPT I) expression through a peroxisome-proliferator-activated receptor alpha (PPARalpha)-independent pathway. Biochem. J. 2001, 354, 189–197. [Google Scholar] [CrossRef]
  694. Mascaro, C.; Acosta, E.; Ortiz, J.A.; Marrero, P.F.; Hegardt, F.G.; Haro, D. Control of human muscle-type carnitine palmitoyltransferase I gene transcription by peroxisome proliferator-activated receptor. J. Biol. Chem. 1998, 273, 8560–8563. [Google Scholar] [CrossRef] [Green Version]
  695. Barrero, M.J.; Camarero, N.; Marrero, P.F.; Haro, D. Control of human carnitine palmitoyltransferase II gene transcription by peroxisome proliferator-activated receptor through a partially conserved peroxisome proliferator-responsive element. Biochem. J. 2003, 369, 721–729. [Google Scholar] [CrossRef]
  696. Gulick, T.; Cresci, S.; Caira, T.; Moore, D.D.; Kelly, D.P. The peroxisome proliferator-activated receptor regulates mitochondrial fatty acid oxidative enzyme gene expression. Proc. Natl. Acad. Sci. USA 1994, 91, 11012–11016. [Google Scholar] [CrossRef] [Green Version]
  697. Aoyama, T.; Peters, J.M.; Iritani, N.; Nakajima, T.; Furihata, K.; Hashimoto, T.; Gonzalez, F.J. Altered constitutive expression of fatty acid-metabolizing enzymes in mice lacking the peroxisome proliferator-activated receptor alpha (PPARalpha). J. Biol. Chem. 1998, 273, 5678–5684. [Google Scholar] [CrossRef] [Green Version]
  698. Badman, M.K.; Pissios, P.; Kennedy, A.R.; Koukos, G.; Flier, J.S.; Maratos-Flier, E. Hepatic fibroblast growth factor 21 is regulated by PPARalpha and is a key mediator of hepatic lipid metabolism in ketotic states. Cell Metab. 2007, 5, 426–437. [Google Scholar] [CrossRef] [Green Version]
  699. Galman, C.; Lundasen, T.; Kharitonenkov, A.; Bina, H.A.; Eriksson, M.; Hafstrom, I.; Dahlin, M.; Amark, P.; Angelin, B.; Rudling, M. The circulating metabolic regulator FGF21 is induced by prolonged fasting and PPARalpha activation in man. Cell Metab. 2008, 8, 169–174. [Google Scholar] [CrossRef] [Green Version]
  700. Iroz, A.; Montagner, A.; Benhamed, F.; Levavasseur, F.; Polizzi, A.; Anthony, E.; Regnier, M.; Fouche, E.; Lukowicz, C.; Cauzac, M.; et al. A Specific ChREBP and PPARalpha Cross-Talk Is Required for the Glucose-Mediated FGF21 Response. Cell Rep. 2017, 21, 403–416. [Google Scholar] [CrossRef] [Green Version]
  701. Li, G.; Brocker, C.N.; Yan, T.; Xie, C.; Krausz, K.W.; Xiang, R.; Gonzalez, F.J. Corrigendum to “Metabolic adaptation to intermittent fasting is independent of peroxisome proliferator-activated receptor alpha” [Mol Metab 7 (2018) 80–89]. Mol. Metab. 2018, 9, 217–219. [Google Scholar] [CrossRef]
  702. Chaix, A.; Lin, T.; Le, H.D.; Chang, M.W.; Panda, S. Time-Restricted Feeding Prevents Obesity and Metabolic Syndrome in Mice Lacking a Circadian Clock. Cell Metab. 2019, 29, 303–319. [Google Scholar] [CrossRef]
  703. Adamo, K.B.; Dent, R.; Langefeld, C.D.; Cox, M.; Williams, K.; Carrick, K.M.; Stuart, J.S.; Sundseth, S.S.; Harper, M.E.; McPherson, R.; et al. Peroxisome proliferator-activated receptor gamma 2 and acyl-CoA synthetase 5 polymorphisms influence diet response. Obesity (Silver Spring) 2007, 15, 1068–1075. [Google Scholar] [CrossRef]
  704. Matsuo, T.; Nakata, Y.; Katayama, Y.; Iemitsu, M.; Maeda, S.; Okura, T.; Kim, M.K.; Ohkubo, H.; Hotta, K.; Tanaka, K. PPARG genotype accounts for part of individual variation in body weight reduction in response to calorie restriction. Obesity (Silver Spring) 2009, 17, 1924–1931. [Google Scholar] [CrossRef]
  705. Stryjecki, C.; Peralta-Romero, J.; Alyass, A.; Karam-Araujo, R.; Suarez, F.; Gomez-Zamudio, J.; Burguete-Garcia, A.; Cruz, M.; Meyre, D. Association between PPAR-gamma2 Pro12Ala genotype and insulin resistance is modified by circulating lipids in Mexican children. Sci. Rep. 2016, 6, 24472. [Google Scholar] [CrossRef] [Green Version]
  706. Masugi, J.; Tamori, Y.; Mori, H.; Koike, T.; Kasuga, M. Inhibitory effect of a proline-to-alanine substitution at codon 12 of peroxisome proliferator-activated receptor-gamma 2 on thiazolidinedione-induced adipogenesis. Biochem. Biophys. Res. Commun. 2000, 268, 178–182. [Google Scholar] [CrossRef]
  707. Memisoglu, A.; Hu, F.B.; Hankinson, S.E.; Manson, J.E.; De Vivo, I.; Willett, W.C.; Hunter, D.J. Interaction between a peroxisome proliferator-activated receptor gamma gene polymorphism and dietary fat intake in relation to body mass. Hum. Mol. Genet. 2003, 12, 2923–2929. [Google Scholar] [CrossRef] [Green Version]
  708. Robitaille, J.; Despres, J.P.; Perusse, L.; Vohl, M.C. The PPAR-gamma P12A polymorphism modulates the relationship between dietary fat intake and components of the metabolic syndrome: Results from the Quebec Family Study. Clin. Genet. 2003, 63, 109–116. [Google Scholar] [CrossRef]
  709. Luan, J.; Browne, P.O.; Harding, A.H.; Halsall, D.J.; O’Rahilly, S.; Chatterjee, V.K.; Wareham, N.J. Evidence for gene-nutrient interaction at the PPARgamma locus. Diabetes 2001, 50, 686–689. [Google Scholar] [CrossRef] [Green Version]
  710. Lindi, V.; Schwab, U.; Louheranta, A.; Laakso, M.; Vessby, B.; Hermansen, K.; Storlien, L.; Riccardi, G.; Matti IJ Uusitupa for the KANWU Study Group. Impact of the Pro12Ala polymorphism of the PPAR-gamma2 gene on serum triacylglycerol response to n-3 fatty acid supplementation. Mol. Genet. Metab. 2003, 79, 52–60. [Google Scholar] [CrossRef]
  711. Nicklas, B.J.; van Rossum, E.F.; Berman, D.M.; Ryan, A.S.; Dennis, K.E.; Shuldiner, A.R. Genetic variation in the peroxisome proliferator-activated receptor-gamma2 gene (Pro12Ala) affects metabolic responses to weight loss and subsequent weight regain. Diabetes 2001, 50, 2172–2176. [Google Scholar] [CrossRef] [Green Version]
  712. Ballor, D.L.; Poehlman, E.T. Exercise intensity does not affect depression of resting metabolic rate during severe diet restriction in male Sprague-Dawley rats. J. Nutr. 1993, 123, 1270–1276. [Google Scholar]
  713. Koizumi, A.; Wada, Y.; Tuskada, M.; Kayo, T.; Naruse, M.; Horiuchi, K.; Mogi, T.; Yoshioka, M.; Sasaki, M.; Miyamaura, Y.; et al. A tumor preventive effect of dietary restriction is antagonized by a high housing temperature through deprivation of torpor. Mech. Ageing Dev. 1996, 92, 67–82. [Google Scholar]
  714. Tabarean, I.; Morrison, B.; Marcondes, M.C.; Bartfai, T.; Conti, B. Hypothalamic and dietary control of temperature-mediated longevity. Ageing Res. Rev. 2010, 9, 41–50. [Google Scholar] [CrossRef] [Green Version]
  715. Rachid, T.L.; Penna-de-Carvalho, A.; Bringhenti, I.; Aguila, M.B.; Mandarim-de-Lacerda, C.A.; Souza-Mello, V. Fenofibrate (PPARalpha agonist) induces beige cell formation in subcutaneous white adipose tissue from diet-induced male obese mice. Mol. Cell Endocrinol. 2015, 402, 86–94. [Google Scholar] [CrossRef]
  716. Rachid, T.L.; Silva-Veiga, F.M.; Graus-Nunes, F.; Bringhenti, I.; Mandarim-de-Lacerda, C.A.; Souza-Mello, V. Differential actions of PPAR-alpha and PPAR-beta/delta on beige adipocyte formation: A study in the subcutaneous white adipose tissue of obese male mice. PLoS ONE 2018, 13, e0191365. [Google Scholar] [CrossRef]
  717. Lu, C.; Cheng, S.Y. Thyroid hormone receptors regulate adipogenesis and carcinogenesis via crosstalk signaling with peroxisome proliferator-activated receptors. J. Mol. Endocrinol. 2010, 44, 143–154. [Google Scholar] [CrossRef]
  718. Liu, Y.Y.; Brent, G.A. Thyroid hormone crosstalk with nuclear receptor signaling in metabolic regulation. Trends Endocrinol. Metab. 2010, 21, 166–173. [Google Scholar] [CrossRef] [Green Version]
  719. Colbert, L.H.; Mai, V.; Tooze, J.A.; Perkins, S.N.; Berrigan, D.; Hursting, S.D. Negative energy balance induced by voluntary wheel running inhibits polyp development in APCMin mice. Carcinogenesis 2006, 27, 2103–2107. [Google Scholar] [CrossRef] [Green Version]
  720. Holloszy, J.O. Mortality rate and longevity of food-restricted exercising male rats: A reevaluation. J. Appl. Physiol. (1985) 1997, 82, 399–403. [Google Scholar] [CrossRef] [Green Version]
  721. Manini, T.M.; Everhart, J.E.; Patel, K.V.; Schoeller, D.A.; Colbert, L.H.; Visser, M.; Tylavsky, F.; Bauer, D.C.; Goodpaster, B.H.; Harris, T.B. Daily activity energy expenditure and mortality among older adults. JAMA 2006, 296, 171–179. [Google Scholar] [CrossRef] [Green Version]
  722. Warburton, D.E.; Nicol, C.W.; Bredin, S.S. Health benefits of physical activity: The evidence. CMAJ 2006, 174, 801–809. [Google Scholar] [CrossRef] [Green Version]
  723. Laaksonen, D.E.; Lindstrom, J.; Lakka, T.A.; Eriksson, J.G.; Niskanen, L.; Wikstrom, K.; Aunola, S.; Keinanen-Kiukaanniemi, S.; Laakso, M.; Valle, T.T.; et al. Physical activity in the prevention of type 2 diabetes: The Finnish diabetes prevention study. Diabetes 2005, 54, 158–165. [Google Scholar] [CrossRef]
  724. Holloszy, J.O. Exercise and longevity: Studies on rats. J. Gerontol. 1988, 43, B149–B151. [Google Scholar] [CrossRef]
  725. Samorajski, T.; Delaney, C.; Durham, L.; Ordy, J.M.; Johnson, J.A.; Dunlap, W.P. Effect of exercise on longevity, body weight, locomotor performance, and passive-avoidance memory of C57BL/6J mice. Neurobiol. Aging 1985, 6, 17–24. [Google Scholar] [CrossRef]
  726. Pekkanen, J.; Marti, B.; Nissinen, A.; Tuomilehto, J.; Punsar, S.; Karvonen, M.J. Reduction of premature mortality by high physical activity: A 20-year follow-up of middle-aged Finnish men. Lancet 1987, 1, 1473–1477. [Google Scholar] [CrossRef]
  727. Fontana, L.; Meyer, T.E.; Klein, S.; Holloszy, J.O. Long-term low-calorie low-protein vegan diet and endurance exercise are associated with low cardiometabolic risk. Rejuvenation Res. 2007, 10, 225–234. [Google Scholar] [CrossRef] [Green Version]
  728. Fontana, L.; Klein, S.; Holloszy, J.O. Long-term low-protein, low-calorie diet and endurance exercise modulate metabolic factors associated with cancer risk. Am. J. Clin. Nutr. 2006, 84, 1456–1462. [Google Scholar] [CrossRef]
  729. Huffman, D.M.; Moellering, D.R.; Grizzle, W.E.; Stockard, C.R.; Johnson, M.S.; Nagy, T.R. Effect of exercise and calorie restriction on biomarkers of aging in mice. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2008, 294, R1618–R1627. [Google Scholar] [CrossRef]
  730. Thomas, A.W.; Davies, N.A.; Moir, H.; Watkeys, L.; Ruffino, J.S.; Isa, S.A.; Butcher, L.R.; Hughes, M.G.; Morris, K.; Webb, R. Exercise-associated generation of PPARgamma ligands activates PPARgamma signaling events and upregulates genes related to lipid metabolism. J. Appl. Physiol. 2012, 112, 806–815. [Google Scholar] [CrossRef] [Green Version]
  731. Zhang, S.; Liu, Y.; Li, Q.; Dong, X.; Hu, H.; Hu, R.; Ye, H.; Wu, Y.; Hu, R.; Li, Y. Exercise improved rat metabolism by raising PPAR-alpha. Int. J. Sports Med. 2011, 32, 568–573. [Google Scholar] [CrossRef]
  732. Santos, M.H.; Higuchi Mde, L.; Tucci, P.J.; Garavelo, S.M.; Reis, M.M.; Antonio, E.L.; Serra, A.J.; Maranhao, R.C. Previous exercise training increases levels of PPAR-alpha in long-term post-myocardial infarction in rats, which is correlated with better inflammatory response. Clinics (Sao Paulo) 2016, 71, 163–168. [Google Scholar] [CrossRef]
  733. Chen, W.; Gao, R.; Xie, X.; Zheng, Z.; Li, H.; Li, S.; Dong, F.; Wang, L. A metabolomic study of the PPARdelta agonist GW501516 for enhancing running endurance in Kunming mice. Sci. Rep. 2015, 5, 9884. [Google Scholar] [CrossRef] [Green Version]
  734. Butcher, L.R.; Thomas, A.; Backx, K.; Roberts, A.; Webb, R.; Morris, K. Low-intensity exercise exerts beneficial effects on plasma lipids via PPARgamma. Med. Sci. Sports Exerc. 2008, 40, 1263–1270. [Google Scholar] [CrossRef]
  735. Ruschke, K.; Fishbein, L.; Dietrich, A.; Kloting, N.; Tonjes, A.; Oberbach, A.; Fasshauer, M.; Jenkner, J.; Schon, M.R.; Stumvoll, M.; et al. Gene expression of PPARgamma and PGC-1alpha in human omental and subcutaneous adipose tissues is related to insulin resistance markers and mediates beneficial effects of physical training. Eur. J. Endocrinol. 2010, 162, 515–523. [Google Scholar] [CrossRef] [Green Version]
  736. Sasaki, T.; Nakata, R.; Inoue, H.; Shimizu, M.; Inoue, J.; Sato, R. Role of AMPK and PPARgamma1 in exercise-induced lipoprotein lipase in skeletal muscle. Am. J. Physiol. Endocrinol. Metab. 2014, 306, E1085–E1092. [Google Scholar] [CrossRef] [Green Version]
  737. Weiss, E.P.; Kulaputana, O.; Ghiu, I.A.; Brandauer, J.; Wohn, C.R.; Phares, D.A.; Shuldiner, A.R.; Hagberg, J.M. Endurance training-induced changes in the insulin response to oral glucose are associated with the peroxisome proliferator-activated receptor-gamma2 Pro12Ala genotype in men but not in women. Metabolism 2005, 54, 97–102. [Google Scholar] [CrossRef]
  738. Kahara, T.; Takamura, T.; Hayakawa, T.; Nagai, Y.; Yamaguchi, H.; Katsuki, T.; Katsuki, K.; Katsuki, M.; Kobayashi, K. PPARgamma gene polymorphism is associated with exercise-mediated changes of insulin resistance in healthy men. Metabolism 2003, 52, 209–212. [Google Scholar] [CrossRef] [Green Version]
  739. Adamo, K.B.; Sigal, R.J.; Williams, K.; Kenny, G.; Prud’homme, D.; Tesson, F. Influence of Pro12Ala peroxisome proliferator-activated receptor gamma2 polymorphism on glucose response to exercise training in type 2 diabetes. Diabetologia 2005, 48, 1503–1509. [Google Scholar] [CrossRef] [Green Version]
  740. Ostergard, T.; Ek, J.; Hamid, Y.; Saltin, B.; Pedersen, O.B.; Hansen, T.; Schmitz, O. Influence of the PPAR-gamma2 Pro12Ala and ACE I/D polymorphisms on insulin sensitivity and training effects in healthy offspring of type 2 diabetic subjects. Horm. Metab. Res. 2005, 37, 99–105. [Google Scholar] [CrossRef]
  741. Pan, W.; Liu, C.; Zhang, J.; Gao, X.; Yu, S.; Tan, H.; Yu, J.; Qian, D.; Li, J.; Bian, S.; et al. Association Between Single Nucleotide Polymorphisms in PPARA and EPAS1 Genes and High-Altitude Appetite Loss in Chinese Young Men. Front. Physiol. 2019, 10, 59. [Google Scholar] [CrossRef]
  742. Fu, J.; Gaetani, S.; Oveisi, F.; Lo Verme, J.; Serrano, A.; Rodriguez De Fonseca, F.; Rosengarth, A.; Luecke, H.; Di Giacomo, B.; Tarzia, G.; et al. Oleylethanolamide regulates feeding and body weight through activation of the nuclear receptor PPAR-alpha. Nature 2003, 425, 90–93. [Google Scholar] [CrossRef]
  743. Fu, J.; Dipatrizio, N.V.; Guijarro, A.; Schwartz, G.J.; Li, X.; Gaetani, S.; Astarita, G.; Piomelli, D. Sympathetic activity controls fat-induced oleoylethanolamide signaling in small intestine. J. Neurosci. 2011, 31, 5730–5736. [Google Scholar] [CrossRef]
  744. Magotti, P.; Bauer, I.; Igarashi, M.; Babagoli, M.; Marotta, R.; Piomelli, D.; Garau, G. Structure of human N-acylphosphatidylethanolamine-hydrolyzing phospholipase D: Regulation of fatty acid ethanolamide biosynthesis by bile acids. Structure 2015, 23, 598–604. [Google Scholar] [CrossRef] [Green Version]
  745. Gaetani, S.; Oveisi, F.; Piomelli, D. Modulation of meal pattern in the rat by the anorexic lipid mediator oleoylethanolamide. Neuropsychopharmacology 2003, 28, 1311–1316. [Google Scholar] [CrossRef]
  746. Fu, J.; Oveisi, F.; Gaetani, S.; Lin, E.; Piomelli, D. Oleoylethanolamide, an endogenous PPAR-alpha agonist, lowers body weight and hyperlipidemia in obese rats. Neuropharmacology 2005, 48, 1147–1153. [Google Scholar] [CrossRef]
  747. Caillon, A.; Duszka, K.; Wahli, W.; Rohner-Jeanrenaud, F.; Altirriba, J. The OEA effect on food intake is independent from the presence of PPARalpha in the intestine and the nodose ganglion, while the impact of OEA on energy expenditure requires the presence of PPARalpha in mice. Metabolism, 2018. [Google Scholar] [CrossRef]
  748. Lo Verme, J.; Gaetani, S.; Fu, J.; Oveisi, F.; Burton, K.; Piomelli, D. Regulation of food intake by oleoylethanolamide. Cell Mol. Life Sci. 2005, 62, 708–716. [Google Scholar] [CrossRef] [Green Version]
  749. Piomelli, D. A fatty gut feeling. Trends Endocrinol. Metab. 2013, 24, 332–341. [Google Scholar] [CrossRef] [Green Version]
  750. Proulx, K.; Cota, D.; Castaneda, T.R.; Tschop, M.H.; D’Alessio, D.A.; Tso, P.; Woods, S.C.; Seeley, R.J. Mechanisms of oleoylethanolamide-induced changes in feeding behavior and motor activity. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2005, 289, R729–R737. [Google Scholar] [CrossRef] [Green Version]
  751. Rodriguez de Fonseca, F.; Navarro, M.; Gomez, R.; Escuredo, L.; Nava, F.; Fu, J.; Murillo-Rodriguez, E.; Giuffrida, A.; LoVerme, J.; Gaetani, S.; et al. An anorexic lipid mediator regulated by feeding. Nature 2001, 414, 209–212. [Google Scholar] [CrossRef] [Green Version]
  752. Gaetani, S.; Fu, J.; Cassano, T.; Dipasquale, P.; Romano, A.; Righetti, L.; Cianci, S.; Laconca, L.; Giannini, E.; Scaccianoce, S.; et al. The fat-induced satiety factor oleoylethanolamide suppresses feeding through central release of oxytocin. J. Neurosci. 2010, 30, 8096–8101. [Google Scholar] [CrossRef] [Green Version]
  753. Koleva, D.I.; Orbetzova, M.M.; Atanassova, P.K. Adipose tissue hormones and appetite and body weight regulators in insulin resistance. Folia Med. (Plovdiv) 2013, 55, 25–32. [Google Scholar]
  754. Gautron, L.; Laye, S. Neurobiology of inflammation-associated anorexia. Front. Neurosci. 2009, 3, 59. [Google Scholar] [CrossRef] [Green Version]
  755. Shi, X.; Wang, X.; Li, Q.; Su, M.; Chew, E.; Wong, E.T.; Lacza, Z.; Radda, G.K.; Tergaonkar, V.; Han, W. Nuclear factor kappaB (NF-kappaB) suppresses food intake and energy expenditure in mice by directly activating the Pomc promoter. Diabetologia 2013, 56, 925–936. [Google Scholar] [CrossRef] [Green Version]
  756. Jang, P.G.; Namkoong, C.; Kang, G.M.; Hur, M.W.; Kim, S.W.; Kim, G.H.; Kang, Y.; Jeon, M.J.; Kim, E.H.; Lee, M.S.; et al. NF-kappaB activation in hypothalamic pro-opiomelanocortin neurons is essential in illness- and leptin-induced anorexia. J. Biol. Chem. 2010, 285, 9706–9715. [Google Scholar] [CrossRef] [Green Version]
  757. Speakman, J.R. Body size, energy metabolism and lifespan. J. Exp. Biol. 2005, 208, 1717–1730. [Google Scholar] [CrossRef] [Green Version]
  758. Pearl, R. The Rate of Living; University of London Press: London, UK, 1928. [Google Scholar]
  759. Houthoofd, K.; Braeckman, B.P.; Lenaerts, I.; Brys, K.; De Vreese, A.; Van Eygen, S.; Vanfleteren, J.R. Axenic growth up-regulates mass-specific metabolic rate, stress resistance, and extends life span in Caenorhabditis elegans. Exp. Gerontol. 2002, 37, 1371–1378. [Google Scholar] [CrossRef]
  760. Metaxakis, A.; Partridge, L. Dietary restriction extends lifespan in wild-derived populations of Drosophila melanogaster. PLoS ONE 2013, 8, e74681. [Google Scholar] [CrossRef]
  761. Abalan, F.; Mayo, W.; Simon, H.; Le Moal, M. Paradoxical effect of severe dietary restriction on Long-Evans rat life span. Int. J. Vitam. Nutr. Res. 2010, 80, 386–393. [Google Scholar] [CrossRef]
  762. Pifferi, F.; Terrien, J.; Marchal, J.; Dal-Pan, A.; Djelti, F.; Hardy, I.; Chahory, S.; Cordonnier, N.; Desquilbet, L.; Hurion, M.; et al. Caloric restriction increases lifespan but affects brain integrity in grey mouse lemur primates. Commun. Biol. 2018, 1, 1–8. [Google Scholar] [CrossRef] [Green Version]
  763. Colman, R.J.; Beasley, T.M.; Kemnitz, J.W.; Johnson, S.C.; Weindruch, R.; Anderson, R.M. Caloric restriction reduces age-related and all-cause mortality in rhesus monkeys. Nat. Commun. 2014, 5, 3557. [Google Scholar] [CrossRef]
  764. Knight, J.A. The biochemistry of aging. Adv. Clin. Chem. 2000, 35, 1–62. [Google Scholar]
  765. Park, S.Y.; Kim, Y.W.; Kim, J.E.; Kim, J.Y. Age-associated changes in fat metabolism in the rat and its relation to sympathetic activity. Life Sci. 2006, 79, 2228–2233. [Google Scholar] [CrossRef]
  766. Petersen, K.F.; Befroy, D.; Dufour, S.; Dziura, J.; Ariyan, C.; Rothman, D.L.; DiPietro, L.; Cline, G.W.; Shulman, G.I. Mitochondrial dysfunction in the elderly: Possible role in insulin resistance. Science 2003, 300, 1140–1142. [Google Scholar] [CrossRef] [Green Version]
  767. Kenyon, C. The plasticity of aging: Insights from long-lived mutants. Cell 2005, 120, 449–460. [Google Scholar] [CrossRef] [Green Version]
  768. Siersbaek, R.; Nielsen, R.; Mandrup, S. PPARgamma in adipocyte differentiation and metabolism-novel insights from genome-wide studies. FEBS Lett. 2010, 584, 3242–3249. [Google Scholar] [CrossRef] [Green Version]
  769. Tontonoz, P.; Hu, E.; Spiegelman, B.M. Stimulation of adipogenesis in fibroblasts by PPAR gamma 2, a lipid-activated transcription factor. Cell 1994, 79, 1147–1156. [Google Scholar] [CrossRef]
  770. Cock, T.A.; Houten, S.M.; Auwerx, J. Peroxisome proliferator-activated receptor-gamma: Too much of a good thing causes harm. EMBO Rep. 2004, 5, 142–147. [Google Scholar] [CrossRef] [Green Version]
  771. Argmann, C.; Dobrin, R.; Heikkinen, S.; Auburtin, A.; Pouilly, L.; Cock, T.A.; Koutnikova, H.; Zhu, J.; Schadt, E.E.; Auwerx, J. Ppargamma2 is a key driver of longevity in the mouse. PLoS Genet. 2009, 5, e1000752. [Google Scholar] [CrossRef]
  772. Viguerie, N.; Vidal, H.; Arner, P.; Holst, C.; Verdich, C.; Avizou, S.; Astrup, A.; Saris, W.H.; Macdonald, I.A.; Klimcakova, E.; et al. Adipose tissue gene expression in obese subjects during low-fat and high-fat hypocaloric diets. Diabetologia 2005, 48, 123–131. [Google Scholar] [CrossRef]
  773. Karbowska, J.; Kochan, Z. Intermittent fasting up-regulates Fsp27/Cidec gene expression in white adipose tissue. Nutrition 2012, 28, 294–299. [Google Scholar] [CrossRef]
  774. Lee, C.K.; Allison, D.B.; Brand, J.; Weindruch, R.; Prolla, T.A. Transcriptional profiles associated with aging and middle age-onset caloric restriction in mouse hearts. Proc. Natl. Acad. Sci. USA 2002, 99, 14988–14993. [Google Scholar] [CrossRef] [Green Version]
  775. Chao, C.; Youssef, J.; Rezaiekhaleigh, M.; Birnbaum, L.S.; Badr, M. Senescence-associated decline in hepatic peroxisomal enzyme activities corresponds with diminished levels of retinoid X receptor alpha, but not peroxisome proliferator-activated receptor alpha. Mech. Ageing Dev. 2002, 123, 1469–1476. [Google Scholar] [CrossRef]
  776. Long, X.; Boluyt, M.O.; O’Neill, L.; Zheng, J.S.; Wu, G.; Nitta, Y.K.; Crow, M.T.; Lakatta, E.G. Myocardial retinoid X receptor, thyroid hormone receptor, and myosin heavy chain gene expression in the rat during adult aging. J. Gerontol. A Biol. Sci. Med. Sci. 1999, 54, B23–B27. [Google Scholar] [CrossRef] [Green Version]
  777. Enderlin, V.; Pallet, V.; Alfos, S.; Dargelos, E.; Jaffard, R.; Garcin, H.; Higueret, P. Age-related decreases in mRNA for brain nuclear receptors and target genes are reversed by retinoic acid treatment. Neurosci. Lett. 1997, 229, 125–129. [Google Scholar] [CrossRef]
  778. Stauber, A.J.; Brown-Borg, H.; Liu, J.; Waalkes, M.P.; Laughter, A.; Staben, R.A.; Coley, J.C.; Swanson, C.; Voss, K.A.; Kopchick, J.J.; et al. Constitutive expression of peroxisome proliferator-activated receptor alpha-regulated genes in dwarf mice. Mol. Pharmacol. 2005, 67, 681–694. [Google Scholar] [CrossRef] [Green Version]
  779. Zhang, Y.; Xie, Y.; Berglund, E.D.; Coate, K.C.; He, T.T.; Katafuchi, T.; Xiao, G.; Potthoff, M.J.; Wei, W.; Wan, Y.; et al. The starvation hormone, fibroblast growth factor-21, extends lifespan in mice. Elife 2012, 1, e00065. [Google Scholar] [CrossRef]
  780. Olsson, B.; Bohlooly, Y.M.; Brusehed, O.; Isaksson, O.G.; Ahren, B.; Olofsson, S.O.; Oscarsson, J.; Tornell, J. Bovine growth hormone-transgenic mice have major alterations in hepatic expression of metabolic genes. Am. J. Physiol. Endocrinol. Metab. 2003, 285, E504–E511. [Google Scholar] [CrossRef]
  781. Sugiyama, H.; Yamada, J.; Suga, T. Effects of testosterone, hypophysectomy and growth hormone treatment on clofibrate induction of peroxisomal beta-oxidation in female rat liver. Biochem. Pharmacol. 1994, 47, 918–921. [Google Scholar] [CrossRef]
  782. Zhou, Y.C.; Waxman, D.J. Cross-talk between janus kinase-signal transducer and activator of transcription (JAK-STAT) and peroxisome proliferator-activated receptor-alpha (PPARalpha) signaling pathways. Growth hormone inhibition of pparalpha transcriptional activity mediated by stat5b. J. Biol. Chem. 1999, 274, 2672–2681. [Google Scholar] [CrossRef] [Green Version]
  783. Zhou, Y.C.; Waxman, D.J. STAT5b down-regulates peroxisome proliferator-activated receptor alpha transcription by inhibition of ligand-independent activation function region-1 trans-activation domain. J. Biol. Chem. 1999, 274, 29874–29882. [Google Scholar] [CrossRef] [Green Version]
  784. Shipley, J.M.; Waxman, D.J. Down-regulation of STAT5b transcriptional activity by ligand-activated peroxisome proliferator-activated receptor (PPAR) alpha and PPARgamma. Mol. Pharmacol. 2003, 64, 355–364. [Google Scholar] [CrossRef]
  785. Zheng, X.; Wang, S.; Jia, W. Calorie restriction and its impact on gut microbial composition and global metabolism. Front. Med. 2018, 12, 634–644. [Google Scholar] [CrossRef] [Green Version]
  786. Ruiz, A.; Cerdo, T.; Jauregui, R.; Pieper, D.H.; Marcos, A.; Clemente, A.; Garcia, F.; Margolles, A.; Ferrer, M.; Campoy, C.; et al. One-year calorie restriction impacts gut microbial composition but not its metabolic performance in obese adolescents. Environ. Microbiol. 2017, 19, 1536–1551. [Google Scholar] [CrossRef]
  787. Tanca, A.; Abbondio, M.; Palomba, A.; Fraumene, C.; Marongiu, F.; Serra, M.; Pagnozzi, D.; Laconi, E.; Uzzau, S. Caloric restriction promotes functional changes involving short-chain fatty acid biosynthesis in the rat gut microbiota. Sci. Rep. 2018, 8, 14778. [Google Scholar] [CrossRef]
  788. Zhang, C.; Li, S.; Yang, L.; Huang, P.; Li, W.; Wang, S.; Zhao, G.; Zhang, M.; Pang, X.; Yan, Z.; et al. Structural modulation of gut microbiota in life-long calorie-restricted mice. Nat. Commun. 2013, 4, 2163. [Google Scholar] [CrossRef] [Green Version]
  789. Fraumene, C.; Manghina, V.; Cadoni, E.; Marongiu, F.; Abbondio, M.; Serra, M.; Palomba, A.; Tanca, A.; Laconi, E.; Uzzau, S. Caloric restriction promotes rapid expansion and long-lasting increase of Lactobacillus in the rat fecal microbiota. Gut Microbes 2018, 9, 104–114. [Google Scholar] [CrossRef] [Green Version]
  790. Liu, T.; Wu, Y.; Wang, L.; Pang, X.; Zhao, L.; Yuan, H.; Zhang, C. A More Robust Gut Microbiota in Calorie-Restricted Mice Is Associated with Attenuated Intestinal Injury Caused by the Chemotherapy Drug Cyclophosphamide. mBio 2019, 10. [Google Scholar] [CrossRef] [Green Version]
  791. Wang, S.; Huang, M.; You, X.; Zhao, J.; Chen, L.; Wang, L.; Luo, Y.; Chen, Y. Gut microbiota mediates the anti-obesity effect of calorie restriction in mice. Sci. Rep. 2018, 8, 13037. [Google Scholar] [CrossRef] [Green Version]
  792. Crawford, P.A.; Crowley, J.R.; Sambandam, N.; Muegge, B.D.; Costello, E.K.; Hamady, M.; Knight, R.; Gordon, J.I. Regulation of myocardial ketone body metabolism by the gut microbiota during nutrient deprivation. Proc. Natl. Acad. Sci. USA 2009, 106, 11276–11281. [Google Scholar] [CrossRef] [Green Version]
  793. Montagner, A.; Korecka, A.; Polizzi, A.; Lippi, Y.; Blum, Y.; Canlet, C.; Tremblay-Franco, M.; Gautier-Stein, A.; Burcelin, R.; Yen, Y.C.; et al. Hepatic circadian clock oscillators and nuclear receptors integrate microbiome-derived signals. Sci. Rep. 2016, 6, 20127. [Google Scholar] [CrossRef]
  794. Mukherji, A.; Kobiita, A.; Ye, T.; Chambon, P. Homeostasis in intestinal epithelium is orchestrated by the circadian clock and microbiota cues transduced by TLRs. Cell 2013, 153, 812–827. [Google Scholar] [CrossRef] [Green Version]
  795. Manoharan, I.; Suryawanshi, A.; Hong, Y.; Ranganathan, P.; Shanmugam, A.; Ahmad, S.; Swafford, D.; Manicassamy, B.; Ramesh, G.; Koni, P.A.; et al. Homeostatic PPARalpha Signaling Limits Inflammatory Responses to Commensal Microbiota in the Intestine. J. Immunol. 2016, 196, 4739–4749. [Google Scholar] [CrossRef]
  796. Li, G.; Xie, C.; Lu, S.; Nichols, R.G.; Tian, Y.; Li, L.; Patel, D.; Ma, Y.; Brocker, C.N.; Yan, T.; et al. Intermittent Fasting Promotes White Adipose Browning and Decreases Obesity by Shaping the Gut Microbiota. Cell Metab. 2017, 26, 801. [Google Scholar] [CrossRef] [Green Version]
  797. Burdick, A.D.; Kim, D.J.; Peraza, M.A.; Gonzalez, F.J.; Peters, J.M. The role of peroxisome proliferator-activated receptor-beta/delta in epithelial cell growth and differentiation. Cell. Signal. 2006, 18, 9–20. [Google Scholar] [CrossRef]
  798. Murakami, M.; Tognini, P.; Liu, Y.; Eckel-Mahan, K.L.; Baldi, P.; Sassone-Corsi, P. Gut microbiota directs PPARgamma-driven reprogramming of the liver circadian clock by nutritional challenge. EMBO Rep. 2016, 17, 1292–1303. [Google Scholar] [CrossRef] [Green Version]
  799. Peyrin-Biroulet, L.; Beisner, J.; Wang, G.; Nuding, S.; Oommen, S.T.; Kelly, D.; Parmentier-Decrucq, E.; Dessein, R.; Merour, E.; Chavatte, P.; et al. Peroxisome proliferator-activated receptor gamma activation is required for maintenance of innate antimicrobial immunity in the colon. Proc. Natl. Acad. Sci. USA 2010, 107, 8772–8777. [Google Scholar] [CrossRef] [Green Version]
  800. Oh, H.Y.P.; Visvalingam, V.; Wahli, W. The PPAR-microbiota-metabolic organ trilogy to fine-tune physiology. FASEB J. 2019, 33, 9706–9730. [Google Scholar] [CrossRef] [Green Version]
  801. Fann, D.Y.; Ng, G.Y.; Poh, L.; Arumugam, T.V. Positive effects of intermittent fasting in ischemic stroke. Exp. Gerontol. 2017, 89, 93–102. [Google Scholar] [CrossRef]
  802. Razeghi Jahromi, S.; Ghaemi, A.; Alizadeh, A.; Sabetghadam, F.; Moradi Tabriz, H.; Togha, M. Effects of Intermittent Fasting on Experimental Autoimune Encephalomyelitis in C57BL/6 Mice. Iran. J. Allergy Asthma Immunol. 2016, 15, 212–219. [Google Scholar]
  803. Okada, T.; Otsubo, T.; Hagiwara, T.; Inazuka, F.; Kobayashi, E.; Fukuda, S.; Inoue, T.; Higuchi, K.; Kawamura, Y.I.; Dohi, T. Intermittent fasting prompted recovery from dextran sulfate sodium-induced colitis in mice. J. Clin. Biochem. Nutr. 2017, 61, 100–107. [Google Scholar] [CrossRef]
  804. Savendahl, L.; Underwood, L.E.; Haldeman, K.M.; Ulshen, M.H.; Lund, P.K. Fasting prevents experimental murine colitis produced by dextran sulfate sodium and decreases interleukin-1 beta and insulin-like growth factor I messenger ribonucleic acid. Endocrinology 1997, 138, 734–740. [Google Scholar] [CrossRef]
  805. Cheng, C.W.; Villani, V.; Buono, R.; Wei, M.; Kumar, S.; Yilmaz, O.H.; Cohen, P.; Sneddon, J.B.; Perin, L.; Longo, V.D. Fasting-Mimicking Diet Promotes Ngn3-Driven beta-Cell Regeneration to Reverse Diabetes. Cell 2017, 168, 775–788. [Google Scholar] [CrossRef] [Green Version]
  806. Wei, M.; Brandhorst, S.; Shelehchi, M.; Mirzaei, H.; Cheng, C.W.; Budniak, J.; Groshen, S.; Mack, W.J.; Guen, E.; Di Biase, S.; et al. Fasting-mimicking diet and markers/risk factors for aging, diabetes, cancer, and cardiovascular disease. Sci. Transl. Med. 2017, 9. [Google Scholar] [CrossRef]
  807. Rangan, P.; Choi, I.; Wei, M.; Navarrete, G.; Guen, E.; Brandhorst, S.; Enyati, N.; Pasia, G.; Maesincee, D.; Ocon, V.; et al. Fasting-Mimicking Diet Modulates Microbiota and Promotes Intestinal Regeneration to Reduce Inflammatory Bowel Disease Pathology. Cell Rep. 2019, 26, 2704–2719. [Google Scholar] [CrossRef] [Green Version]
  808. Liu, K.; Liu, B.; Heilbronn, L.K. Intermittent fasting: What questions should we be asking? Physiol. Behav. 2020, 218, 112827. [Google Scholar] [CrossRef]
  809. Horne, B.D.; Muhlestein, J.B.; Anderson, J.L. Health effects of intermittent fasting: Hormesis or harm? A systematic review. Am. J. Clin. Nutr. 2015, 102, 464–470. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Activation and main functions of peroxisome proliferator-activated receptors (PPARs) in different tissues. PPARs share fatty acids (FA) as common ligands, peroxisome proliferator response elements (PPRE) as their DNA binding site, and retinoid X receptors (RXR) as their heterodimer partner. However, each PPAR shows distinct expression and function patterns. The dominant role of PPARα is connected to metabolic adjustment in the liver and brown adipose tissue (BAT). PPARβ/δ is primarily associated with muscle and white adipose tissue (WAT) metabolism, as well as with organ development. PPARγ is a master regulator of adipogenesis and WAT maintenance and plays an important anti-inflammatory role. However, this cartoon represents a schematic and simplified view of much more complex patterns.
Figure 1. Activation and main functions of peroxisome proliferator-activated receptors (PPARs) in different tissues. PPARs share fatty acids (FA) as common ligands, peroxisome proliferator response elements (PPRE) as their DNA binding site, and retinoid X receptors (RXR) as their heterodimer partner. However, each PPAR shows distinct expression and function patterns. The dominant role of PPARα is connected to metabolic adjustment in the liver and brown adipose tissue (BAT). PPARβ/δ is primarily associated with muscle and white adipose tissue (WAT) metabolism, as well as with organ development. PPARγ is a master regulator of adipogenesis and WAT maintenance and plays an important anti-inflammatory role. However, this cartoon represents a schematic and simplified view of much more complex patterns.
Cells 09 01708 g001
Figure 2. The main signaling pathways associated with mammalian target of rapamycin (mTOR). mTOR integrates inputs from multiple sources including growth factors, insulin, stress, energy balance, oxygen, and nutrients, and it controls many major downstream processes, including metabolism, macromolecule synthesis, mitochondria function, cell growth, and autophagy.
Figure 2. The main signaling pathways associated with mammalian target of rapamycin (mTOR). mTOR integrates inputs from multiple sources including growth factors, insulin, stress, energy balance, oxygen, and nutrients, and it controls many major downstream processes, including metabolism, macromolecule synthesis, mitochondria function, cell growth, and autophagy.
Cells 09 01708 g002
Figure 3. Interactions between PPARs and mTOR. mTOR interacts with all PPARs, resulting in the modulation of ketogenesis, autophagy, and adipogenesis.
Figure 3. Interactions between PPARs and mTOR. mTOR interacts with all PPARs, resulting in the modulation of ketogenesis, autophagy, and adipogenesis.
Cells 09 01708 g003
Figure 4. The tissue-specific outcomes of the interaction between PPARs and adenosine monophosphate (AMP)-activated protein kinase (AMPK). PPARγ interacts with AMPK in multiple tissues including blood vessels, WAT, pancreas, muscle, heart, and liver, leading to enhanced metabolism as well as reduced oxidative stress and inflammation. PPARα, in cooperation with AMPK, affects metabolism in the liver as well as reducing inflammation and apoptosis in blood vessels, whereas PPAR β/δ with AMPK affects muscle performance.
Figure 4. The tissue-specific outcomes of the interaction between PPARs and adenosine monophosphate (AMP)-activated protein kinase (AMPK). PPARγ interacts with AMPK in multiple tissues including blood vessels, WAT, pancreas, muscle, heart, and liver, leading to enhanced metabolism as well as reduced oxidative stress and inflammation. PPARα, in cooperation with AMPK, affects metabolism in the liver as well as reducing inflammation and apoptosis in blood vessels, whereas PPAR β/δ with AMPK affects muscle performance.
Cells 09 01708 g004
Figure 5. Pathways in which PPARγ activity leads to increased insulin sensitivity. PPARγ affects insulin sensitivity by managing glucose uptake and disposal, enhancing insulin signaling, and maintaining functioning WAT and pancreas.
Figure 5. Pathways in which PPARγ activity leads to increased insulin sensitivity. PPARγ affects insulin sensitivity by managing glucose uptake and disposal, enhancing insulin signaling, and maintaining functioning WAT and pancreas.
Cells 09 01708 g005
Figure 6. The interaction between sirtuin 1 (SIRT1) and PPARs. Caloric restriction (CR)-triggered energy shortage leads to the activation of SIRT1 and its interaction with PPARs. Each of these interactions results in a distinct outcome.
Figure 6. The interaction between sirtuin 1 (SIRT1) and PPARs. Caloric restriction (CR)-triggered energy shortage leads to the activation of SIRT1 and its interaction with PPARs. Each of these interactions results in a distinct outcome.
Cells 09 01708 g006

Share and Cite

MDPI and ACS Style

Duszka, K.; Gregor, A.; Guillou, H.; König, J.; Wahli, W. Peroxisome Proliferator-Activated Receptors and Caloric Restriction—Common Pathways Affecting Metabolism, Health, and Longevity. Cells 2020, 9, 1708. https://doi.org/10.3390/cells9071708

AMA Style

Duszka K, Gregor A, Guillou H, König J, Wahli W. Peroxisome Proliferator-Activated Receptors and Caloric Restriction—Common Pathways Affecting Metabolism, Health, and Longevity. Cells. 2020; 9(7):1708. https://doi.org/10.3390/cells9071708

Chicago/Turabian Style

Duszka, Kalina, András Gregor, Hervé Guillou, Jürgen König, and Walter Wahli. 2020. "Peroxisome Proliferator-Activated Receptors and Caloric Restriction—Common Pathways Affecting Metabolism, Health, and Longevity" Cells 9, no. 7: 1708. https://doi.org/10.3390/cells9071708

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