Skip to content
BY-NC-ND 4.0 license Open Access Published by De Gruyter December 31, 2018

Prospects of Pharmacological Interventions to Organismal Aging

  • Olivia Hillson , Suam Gonzalez and Charalampos Rallis EMAIL logo
From the journal Biomolecular Concepts

Abstract

Intense research in the areas of cellular and organismal aging using diverse laboratory model systems has enriched our knowledge in the processes and the signalling pathways involved in normal and pathological conditions. The field finds itself in a position to take decisive steps towards clinical applications and interventions not only for targeted age-related diseases such as cardiovascular conditions and neurodegeneration but also for the modulation of health span and lifespan of a whole organism. Beyond nutritional interventions such as dietary restriction without malnutrition and various regimes of intermittent fasting, accumulating evidence provides promise for pharmacological interventions. The latter, mimic caloric or dietary restriction, tune cellular and organismal stress responses, affect the metabolism of microbiome with subsequent effects on the host or modulate repair pathways, among others. In this mini review, we summarise some of the evidence on drugs that can alter organismal lifespan and the prospects they might offer for promoting healthspan and delaying age-related diseases.

Introduction

Human life expectancies are steadily rising worldwide, with the demographics of developed countries being shifted towards older ages [1]. This change is accompanied by a disturbing price tag: an increase in occurrence of age-related diseases together with personal, social and financial burdens [2] with 40% of the National Health System’s budget spent on over 65s in the UK. Aging research and biogerontology has long been concentrated on the molecular workings underlying the aging phenomenon towards increasing the healthspan and extending the productive timeframes of an individual. Based on various theories of aging [3, 4, 5, 6], scientists have focused efforts on processes, genetic, epigenetic or environmental factors that could be setting the pace of the biological clock; or alternatively those that are potentially direct targets, mediators or effectors of this clock. For example, the understanding of the role of nutrition and evolutionarily conserved nutrient-responsive pathways, such as the Insulin Growth Factor (IGF) or the Target of rapamycin (TOR) [7], has been pivotal in linking metabolism, bioenergetics and lifespan. Dissection of these pathways has led us to comprehend genetic perturbations that prolong lifespan, and has provided a picture of the molecular players and functional networks towards developing pharmacological interventions for modulation of stress responses and amelioration or even prevention of age-related diseases. In this mini review we will briefly discuss some of these drugs that clearly extend lifespan or reduce aging rates in diverse model systems. We will refer to signalling pathways such as TOR, AMP-activated protein kinase (AMPK) and IGF. We will also discuss compounds such as metformin or aspirin that are widely used in pathologies as well as dietary supplements such as resveratrol. The effects of these drugs on aging and lifespan are without a doubt impressive based on laboratory model systems data. Although some of them are prescribed for treating diseases (such as metformin for diabetes) their ability to modulate accordingly human aging rates, lifespan and more importantly, healthspan still remains to be demonstrated.

Rapamycin, rapalogs and ATP-analogues: inhibiting the TOR signaling pathway

First isolated from the bacteria Streptomyces hygroscopicus by Suren Sehgal in 1972, and identified as an antifungal [8], the macrolide rapamycin has been the subject of research interest for nearly 50 years. Initially, rapamycin showed immunosuppressive potential and gained FDA approval for this purpose [9]. It has since been found to be the first pharmacological agent to extend lifespan in both genders of a mammalian species [10]. As its name suggests, rapamycin works by inhibition of the Target of Rapamycin (TOR) pathway which is an evolutionarily conserved nutrient-responsive pathway serving as a cellular rheostat of energy and a regulator of metabolism and growth [11].

Initial discovery of the TOR pathway came in the early 1990s (much later than the discovery of rapamycin) by several methods. Firstly, a genetic screen for Saccharomyces cerevisiae mutants identified the gene encoding the cellular receptor for rapamycin, FKBP (FPR1) [12] after the same team had previously identified FKBP as the binding protein for rapamycin’s structural homologue FK-506 [13]. While the genes were identified in Heitman’s screen, TOR1 and TOR2, the two TOR kinase homologues more commonly referred to as the targets of Rapamycin, were not fully characterised until 1993 and 1994 when TOR2 was identified as a target of rapamycin [14] and TOR1/2 were found to be structurally and functionally similar but non-identical [15].

The importance of these characterisations can only be fully appreciated when viewed within the context of the conservation of TOR from yeast to man. In humans, there is one TOR kinase, known as mechanistic TOR or mTOR [11], rather than the two homologues found in yeast [16]. The isolation of mTOR came in 1994 and marked the first evidence that yeast could be used as a relevant model organism for TOR-related research in humans. mTOR was initially identified as the FKBP-rapamycin-associated-protein (FRAP) [17] but was referred to as mTOR after it was found to be an orthologue to the yeast TOR homologues [18]. Intense research established that mTOR, as its yeast orthologues, exists and functions within two highly conserved protein complexes termed mTORC1 and mTORC2 [9].

TOR1, TOR2 and mTOR are all considered to be rapamycin targets (Figure 1A) in their respective organisms and, while some key points are known, research into the relationship between TOR inhibition and aging is still ongoing. However, despite the lack of a fully elucidated mechanism, rapamycin and its analogs or ‘rapalogs’ have long been considered a key candidate for the pharmacological intervention of aging, with strong evidence showing that rapamycin increases the lifespan and healthspan in mouse models [19, 20, 21, 22, 23, 24]. Rapalogs hold an advantage over rapamycin itself as a treatment option as they can be developed to have more favourable pharmacological conditions and provide an opportunity for intellectual property, which can be advantageous to the drug development industry [25].

Figure 1 Overview of the signaling pathways controlling cellular growth, metabolism and affecting aging, and some of the drugs mentioned in the manuscript. A. Relationships between IGF/GH axis, TOR and AMPK pathways and sirtuins in aging and lifespan regulation and drugs that either inhibit or boost their functions (look inside text for details). B. General scheme for the inhibitory action of aspirin on COX1 and COX2 leading to beneficial cellular and organismal effects.
Figure 1

Overview of the signaling pathways controlling cellular growth, metabolism and affecting aging, and some of the drugs mentioned in the manuscript. A. Relationships between IGF/GH axis, TOR and AMPK pathways and sirtuins in aging and lifespan regulation and drugs that either inhibit or boost their functions (look inside text for details). B. General scheme for the inhibitory action of aspirin on COX1 and COX2 leading to beneficial cellular and organismal effects.

The two complexes, mTORC1 and mTORC2, are differently affected by rapamycin treatment with mTORC1 inhibition occurring immediately and mTORC2 inhibition occurring only after prolonged treatment with the drug [26]. The two mTOR complexes are not only structurally different from one another; they also have distinct differences in their downstream functions. mTORC1 is associated with the control of anabolic and catabolic processes in response to nutrient availability [27] and is much better understood than mTORC2 but it is believed that both could potentially affect healthy lifespan and aging.

Beyond lifespan extension, rapamycin has also shown promise in age-related disease intervention. Age-related diseases pose a threat in much of the developed world where longer lives, poorer diets and lifestyles are considered to be contributing to the rise in these diseases. According to the World Health Organisation there is a distinct gap between life expectancy and healthy life expectancy across the globe. This is an important point to address with pharmacological interventions since the goal would be to extend both the lifespan and healthspan not just increase lifespan at any cost to life quality.

Inflammation is considered to play an important role in many age-related diseases and rapamycin has been shown to have anti-inflammatory effects [27] to the extent of being used as an immunosuppressive drug [9]. Chronic inflammation has been shown to underpin the mechanism of a number of age-related diseases including cancer, dementia and cardiovascular diseases [28]. Dietary and exercise interventions are often advised in these cases however there is a potential for rapamycin to add to these as a pharmacological option.

It is not just inflammation-related mechanisms that rapamycin has shown promise in such diseases. In cancer, rapamycin has been FDA approved for renal cell carcinoma, mantle cell lymphoma, and pancreatic cancer [27]. It is common to see an upregulation of mTOR signalling in cancer patients, creating an interest in mTOR inhibition as a treatment option. However, a major limitation of using rapamycin or rapalogs in this way is that the drugs are often cytostatic rather than cytotoxic causing tumour growth to slow or stop but not reducing tumour size [25].

In Alzheimer’s disease and vascular dementia, cerebrovascular dysfunction can be one of the early symptoms used to diagnose the disease [29]. It is also suggested to be a major contributor to disease onset and progression [30]. This highlights the importance of recent findings that mTOR could be a target for neuroprotection and vasculoprotection via inhibition with rapamycin [29] such as the inhibition of mTOR with rapamycin preventing ANG II-mediated endothelial vascular dysfunction [31].

Unfortunately, the use of rapamycin as a drug does not come without side effects. Dyslipidaemia is a common and concerning side effect of mTOR inhibition and is observed in 40-75% of patients treated with rapamycin or rapalogs [32, 33]. Inhibition of mTORC1, seen in acute rapamycin treatment, leads to an increase in LDL cholesterol levels, hyperlipidaemia, and increased lipophagy. Chronic rapamycin treatment promoting mTORC2 inhibition also increases lipolysis via an unknown mechanism [33]. Interestingly, despite the high occurrence of dyslipidaemia, rapamycin treatment can be used to combat atherosclerosis, the process by which damage to blood vessel walls leads to blockages and therefore coronary problems such as stroke and heart attack. mTOR inhibition by rapamycin can lead to decreased atherosclerosis and potentially even reversal of disease progression [32]. The combination of rapamycin or rapalogs with stents has also been shown to be successful either as a drug eluting stent [34] or as an oral treatment alongside a base metal stent [35].

Intense research has led to the identification of potent ATP-analog mTOR inhibitors. Developed by AstraZeneca, the ATP-competitive mTOR inhibitors Torin1 [36] and Torin2 [37] can inhibit both mTORC1 and mTORC2 [25] (Figure 1A). Torin1 has been used in a few settings to demonstrate anti-aging properties. Interestingly, Torin1 was shown to be more potent than rapamycin in inhibiting senescent morphology in human cells, suggesting that these processes may rely on rapamycin-insensitive aspects of TOR signaling [38, 39]. Torin1-fed fruit flies exhibited increased lifespan without reduced fertility [40]. Regarding treatment of age-related diseases with ATP-competitive TOR inhibitors there is currently limited data available. Nevertheless, Torin1 has no additional beneficial effects in cancer treatment compared to rapamycin: strong pan-TOR inhibition causes severe cell growth arrest and inhibition of pivotal cellular processes within all, healthy or not cells [25, 41]. Having only been synthesised in 2009, research into the effects of Torin1 and 2 on aging and age-related diseases is still in its early steps. The potency and selectivity of their action encourage further analysis of how they could be utilised in aging and pathologies.

Resveratrol

In the context of biogerontology, resveratrol is one of the most widely studied molecules [42] due its notable ability to counteract different age-related diseases in mammals with apparent lack of toxicity [43]. This polyphenolic natural product is not recommended during pregnancy, even though it is not toxic for animals [44]. It was initially described in 1939 in ethanol extracts of the white hellebore Veratrum grandiflorum and firstly characterized as a phytoalexin [42]. In recent years, it has been shown to be present in grapes, red wine and a few other species under stress conditions [45].

There are two isomeric configurations of this compound (3,5,4’-trihydroxystilbene) named trans-(E) and cis-(Z), which may undergo isomerization upon exposure to ultraviolet radiation [46]. The 4-hydroxystilbene skeleton in resveratrol acts as an antioxidant pharmacophore with the ability to scavenge free radicals [47, 48]. Several studies focused on this antioxidant capacity showed that it inhibits both the formation of copper-catalyzed LDL oxidation [49], and the peroxidation of membrane lipids [50]. Additionally, it restricts the release of inflammatory mediators contributing to cardiovascular disease [51], prevents the formation and growth of multiple types of cancers even by topical application in a model of skin cancer [43, 52] and is neuroprotective [53]. In contrast, a few studies have reported negative effects, such as increased atherosclerosis [54] and DNA damage [55].

Most of the health-enhancing properties of resveratrol are mediated by inhibition of cyclooxygenases enzymes COX-1 and COX-2 [56], various cytochrome P450s [57], NADPH oxidase (the enzyme responsible for reactive oxygen species (ROS) production) [58], PKD (protein kinase D) [59], S6 kinase [60], the transcription factor AP-1 (activator protein 1), and activation of sirtuins (Figure 1A), an evolutionary conserved family of NAD+- dependent (class III) histone/protein deacetylases with a wide range of biological functions [61]. Sirtuins deacetylate a plethora of substrates such as NF-κB, p53, PGC1α (peroxisome proliferator-activated receptor gamma coactivator 1α), FOXOs (forkhead box transcription factors) [62]. All these regulators influence the mitochondrial environment and metabolic diseases [63], and are actively involved in inflammation, carcinogenesis, lipolysis and other pathologies [64, 65, 66].

Besides the antioxidant effects of resveratrol described in humans, its ability to increase lifespan have been shown in vitro and in vivo in several eukaryotic model organisms [67]; in budding yeast this phenotype is thought to be mediated by overexpression of the enzyme Sir2 [68, 69, 70], which consumes NAD+ resembling lifespan extension by caloric restriction [71]. Homologs of this enzyme upregulated by resveratrol in worms [72], fruit flies [73], short-lived fishes [74, 75, 76], obese mice [43], and rhesus monkeys fed a high-fat/high-sucrose diet [77] are also reported to extend lifespan. Nevertheless, relevant experiments in worms and Drosophila have cast doubt on the robustness of the previously reported effects of sirtuins on lifespan control [78]. Mammalian SIRT1 is the most characterized of the seven sirtuins (SIRT1-7) and it regulates various cellular processes through the correct activation of AMPK and inhibition of mTOR signaling. SIRT1 is the main target of resveratrol in vivo (Figure 1A) through a direct allosteric binding that enhances deacetylation of non-tagged peptide substrates [70, 79, 80]. Its overexpression is beneficial in cellular models of Alzheimer’s disease [82], cancer [83], type II diabetes [84], and cardiovascular disease [66]. Despite all these positive data, healthy wild-type mice show no increase in longevity by resveratrol treatment, and there are no large-scale studies on lifespan of healthy humans to date [42, 62]. Therefore, while data from the effects on mammalian lifespan in population studies is limited, it shows a remarkable capacity to enhance health and longevity in the presence of pathologies due to its pleiotropic effects, which exert physiological defence mechanisms that increase survival and make its pharmacological assessment complicated at the same time [85]. Nonetheless, it remains inconclusive whether resveratrol contains life-prolonging properties [86]. The large amount of on-going clinical trials in different pathological contexts and clinical settings is likely to produce valuable data of its effects on human health.

Targeting the GH/IGF axis in aging and age-related diseases

Growth hormone (GH) or somatotropin secretion leads to a plethora of effects connected to cellular and organismal growth and is reduced during aging, a phenomenon known as somatopause [87]. In addition, mutations in the GH axis increase lifespan, while expression of bovine GH in mice reduces it [88]. These findings resulted in proposing and even misusing human GH as an anti-aging therapy (Figure 1A). The effects of such therapies are questionable: revisiting and analysis of clinical trials on GH supplementation in the elderly revealed minor benefits related to adipose tissue composition and an increased risk of insulin resistance development or diabetes mellitus in more than half of the male participants during 26 weeks of GH therapy [89, 90]. While systemic administration of GH is probably detrimental or even dangerous, local treatments might have beneficial effects: GH injections in the knee of older individuals showed improvements in tendon collagen synthesis [91].

An interesting prospect in treatment of aging-related diseases is the utilisation of GH antagonists or somatostatin analogues (Figure 1A) normally used in treatment of acromegaly. Acromegaly, caused by excess GH in humans, is manifested by increased growth of hands, feet, face and megalocardia [92]. Patients with acromegaly are often insulin-resistant, develop diabetes mellitus, have cardiovascular problems and have increased risk of colon cancer occurrence [93]. Treatment with somatostatin analogues that inhibit GH secretion or the GH antagonist pegvisomant (that inhibits endogenous GH binding to its GH receptor), effectively brings mortality rates of acromegaly patients to those of unaffected population [93]. Pegvisomant is more efficient compared to somatostatin analogues for normalising IGF-1 levels and, while the results of combinatorial therapies are very encouraging [94], a large a global safety surveillance study (ACROSTUDY) set in 14 countries (373 sites) has concluded that pegvisomant is safe and effective for patients with acromegaly as monotherapy [95].

IGF-1/insulin signaling is extensively studied in the context of lifespan in invertebrates, with a number of studies in murine models [96]. Notably, a recent study in mice showed that insulin-like growth factor-1 receptor (IGF-1R) monoclonal antibodies (mAb) are a promising anti-aging approach (Figure 1A): IGF-1R mAb preferentially improved female healthspan and increased median female lifespan by 9%. These changes were accompanied by a reduction in both cancer occurrence and general inflammation [97]. The best evidence of improved human healthspan for the IGF-1 pathway derives from studies on centenarians. These individuals are more likely to have IGF-1 receptor variants that are associated with reduced function. In addition, reduced IGF-1 levels are predictive of enhanced survival in female nonagenarians (i.e. women 90-99 years of age) [98].

Metformin

Metformin is the most widely used oral hypoglycemic agent and a first-line treatment for non-insulin-dependent (type 2) diabetes [99]. This compound is derived from a natural product called galegine obtained from the plant French lilac Galega officinalis [100]. Extracts of this plant were used in herbal medicine in medieval times for the treatment of painful/frequent urination accompanying diabetes mellitus. The glucose-lowering activity of galegine was discovered in 1920s, proving to be toxic in humans, but was re-introduced to clinical use in 1957 when was established as a safe and effective therapy for the treatment of type 2 diabetes [101]. Considering that this drug was not designed to target a specific pathway or disease and, despite its clinical use for over 60 years, the primary molecular mechanism of this dimethylbiguanide has remained unclear [99]. Several studies have established that the major effects of the compound are the suppression of ˃60% gluconeogenesis/lipogenesis in the liver and the increase in insulin-mediated uptake of glucose in muscles [102, 103]. These have been attributed to different pathways including AMPK (Figure 1A).

The adenosine monophosphate-activated protein kinase (AMPK) pathway is an evolutionarily conserved signaling pathway that senses cellular energy status through AMP/ATP and ADP/ATP ratios promoting catabolism and inhibiting anabolism [104]. It is not a surprise that dietary restriction activates AMPK and nutritional or caloric overload inhibits it, leading to increased occurrence of metabolic syndrome [105, 106]. AMPK modulates multiple processes such as gluconeogenesis, lipid oxidation, mitophagy and protein synthesis through TOR regulation [104]. AMP-activated protein serine/threonine kinases are heretotrimeric consisting of one enzymatic and two regulatory subunits. AMPK is tightly regulated by upstream kinases and phosphatases such as liver kinase B (LKB), Ca2+/calmodulin-dependent protein kinase kinase β (CaMKKβ) and transforming growth factor-β-activated kinase 1 (TAK1) [107, 108, 109]. Intense studies have provided insights on the structure and function of AMPK and how small molecules can inhibit or enhance its actions [110, 111].

AMPK is directly linked to aging. Beyond cellular energy status AMPK coordinates and controls repair and housekeeping mechanisms linked with maintenance, senescence and lifespan, such as autophagocytosis [112], ER stress suppression [113], oxidative stress alleviation [114] or suppression of inflammation [115]. Overexpression of AMPK can extend lifespan in worms [116] and Drosophila [117]. In mammals the situation is complicated, with gene deletions having severely detrimental effects [118]. AMPK activation through physical exercise has beneficial effects [119], while activation in pathological conditions such as stroke can enhance pathologies and induce additional cellular and tissue damage [120, 121]. The ability of AMPK to be activated upon certain stimuli declines during lifespan. For example, AMPK activity increases within muscles of young but not old rats [122]. Decline in the AMPK activation ability with aging is now linked with age-related diseases such as cardiovascular diseases and metabolic syndrome [123, 124]. As already mentioned, physiological regulation of AMPK through nutrition and exercise is possible. Nevertheless, increasing data strongly suggests potent activation of the pathway with the use of drugs, some of them as common as metformin and aspirin.

Metformin-mediated actions that are AMPK-dependent involve phosphorylation/activation of AMPK in its catalytic α1/α2 subunits at threorine-172, mediated by LKB1 (liver kinase B1) [125]. This is triggered by the inhibition of the mitochondrial respiratory chain complex I, which produces a decrease in ATP levels together with an increase in the ratios of ADP/ATP [126] and AMP/ATP that activate AMPK [127]. The latter results leading in subsequent phosphorylation of CRTC2 (CREB-regulated transcription coactivator 2) and CBP (CREB-binding protein), that downregulate gluconeogenic gene expression [128, 129]. Interestingly, another study reported an AMPK-independent pathway based on results from AMPK α1/α2 knockout mice, indicating that the decrease in glucose production occurs through the regulation of gluconeogenesis flux in response to a decrease of the hepatic energy availability instead of direct suppression of gluconeogenic gene expression [130].

Lifespan extension is among the therapeutic opportunities being explored for metformin. They have been attributed to transcription, genome stability, epigenome marking, metal-interactive regulation of protein function that inhibit pro-inflammatory proteases and anti-apoptotic pathways independent of its glucose reducing effects [131, 132, 133, 134]. This was proved in various systems including S. cerevisiae where extension of chronological lifespan was reported due to glycation inhibition (mimics calorie restriction) restoring deregulated proteins involved in mitochondrial respiration and facilitating the shift of metabolism from fermentation to respiration [135]. In C. elegans metformin works via impairment of folate and methionine metabolism in the intestinal microbiome [136], while in mice reduces adipose tissue inflammation [137]. Within fly intestinal stem cells, metformin inhibits aging phenotypes in an Atg6 (autophagy-related 6)-dependent manner [138]. Additionally, metformin is reported to inhibit differentiation of monocytes into macrophages through disruption of NF-κB (nuclear factor kappa-light-chain-enhancer of activated B cells) signaling [139]. In human plasma it suppresses proinflammatory cytokines such as CCL11 (CC motif chemokine 11) involved in age-related cellular and tissue dysfunction and affects the neutrophil to lymphocyte ratio, which is a marker of inflammation linked to mortality [140].

Clinical trials in humans confirmed the significant reduction of CRP (C-reactive protein) levels, which is a marker of systemic inflammation. Interestingly, such an effect is not observed with resveratrol treatment [99]. In cells derived from patients with Hutchinson-Gilford progeria syndrome, metformin lowers progerin production (a toxic protein also present in normal subjects at lower levels). This is done through inhibition of SRSF1 (serine-arginine rich splicing factor 1) alleviating pathological defects. In this scenario, cellular stress-induced AMPK activation promotes both the activity of the splicing factor-associated protein p32. This protein acts as endogenous inhibitor of SRSF1 and binds to mitochondrial mRNAs in vivo, regulating their efficient translation. As these RNAs code for proteins involved in oxidative phosphorylation, p32 is pivotal in maintaining this essential cellular function. In addition, p32 affects the transcriptional activity of NFE2L2 (Nuclear factor erythroid-derived 2-like 2), a main regulator of the antioxidant response disrupted in progeria cells [134]. Another possibility to explain the health span-extension effects of metformin is through the inhibition of nutrient/energy-sensing metabolic pathways such as the insulin/IGF-1 (insulin-like growth factor 1) and mTOR [133, 141]. Other studies have reported neuroprotective effects against cognitive dysfunction by hippocampal neurogenesis/differentiation and inhibition of aging-related neuroinflammation in mice [142, 143, 144]. The overall favourable effects of metformin in physiological functions through multiple modest, but substantial, effects, combined with its well-characterized profile, suggest that it may be beneficial for the treatment of normal or pathological aging [134].

Aspirin

Acetylsalicylic acid is the most common nonsteroidal anti-inflammatory drug and globally used medication. Considered against its potential side effects of producing stomach ulcers and bleeding, is currently widely used for the treatment of pain, fever, inflammation, platelet aggregation, prevention of cardiovascular pathologies and cancer [145]. Its active component salicylate is chemically synthesized from the bark of the willow tree Salix alba. Its therapeutic use dates back to ancient times but pure acetyl salicylate has been manufactured and commercialized since 1899 [146, 147].

Aspirin affects multiple signal transduction pathways; the principal mode of action discovered in 1971 is through the irreversible inactivation of COX-1 (prostaglandin-endoperoxide synthase, PTGS1) and the inhibition of COX-2 (PTGS2), suppressing prostanoid biosynthesis [148] (Figure 1B). The drug is rapidly broken down in vivo to salicylate. The latter has multiple effects such as uncoupling oxidative respiration through proton transport on the inner mitochondrial membrane [149]; acetylation and inhibition of G6PD (glucose-6-phosphate dehydrogenase), which catalyzes the first reaction in the pentose phosphate pathway involved in the regulation of oxidative stress [150]; activation of AMPK via Thr-172 phosphorylation; and activation of protein kinase IKKβ (inhibitor of nuclear factor kappa-B) that arrests the pro-inflammatory transcription factor NF-κB [151]. Additionally, salicylate competitively inhibits the binding of acetyl coenzyme A (the sole acetyl group donor) to acetyltransferases such as EP300 (E1A-associated protein p300) thus inhibiting its activity and inducing the autophagic cascade that enhances longevity, still observed in the absence of AMPK [152, 153].

The pro-health benefits of aspirin also include the delay in the onset of various age-related diseases and an increase in the maximum and mean lifespan of different organisms through pleiotropic molecular mechanisms [154]. Lifespan extension have been reported in worms via attenuation of endogenous levels of ROS as well as upregulation of antioxidant genes [155], activation of the transcription factors DAF-12 and DAF-16 that increase lipid hydrolysis and inhibit the proliferation of germline stem cells without alterations in the number of offspring [154]. In the fruit fly, it has been associated to a decrease in female fecundity produced by the inhibition of the heme peroxidase Pxt, a COX-like facilitator of follicle maturation [156]. These effects are accompanied by increased resistance to stress and improved locomotor activity that are overall mediated by the Pkh2-ypk1-lem3-tat2 signaling pathway [157]. Interestingly, in mice, aspirin increases the survival of males but not females [158]. Different trials have shown protective effects against Alzheimer’s and Parkinson’s disease, low risk of cancer incidence and mortality[ 159, 160] with metabolic and immunity functions [154]. However, the successful use of aspirin in humans to actually increase healthspan and lifespan has yet to be tested [161].

Future prospects

The current challenges for the treatment of multiple age-related diseases are focused on the identification of molecules that could increase human lifespan and the development of strategies to safely assess them in the population. The common mechanisms of action observed between aspirin, resveratrol and metformin could bring light to this matter by pharmacologically inducing the benefits of caloric restriction. Statins are inhibitors of the HMG-CoA-competitive reductase and exhibit a plethora of effects. They affect cholesterol production and protein prenylation resulting in improvement of endothelial and immune function as well as in other cardiovascular benefits [162]. Despite the heterogeneity and inconclusive results from clinical trials, statins present some potential in terms of their anti-inflammatory properties and in the possibility of enhancing or improving present therapies for some types of cancer [162].

Angiotensin-converting-enzyme (ACE) inhibitors and angiotensin II receptor blockers are used for hypertension treatment. Nevertheless, Angiotensin II is shown to be implicated in age-related cardiovascular disease in rats [163]. The renin-angiotensin system is shown to play an important role in aging kidneys [164], while angiotensin II inhibitors increase the lifespan of hypertensive rats [165, 166]. In humans, there is no large study for the effects on lifespan and aging rates. However, beneficial effects on cardiac hypertrophy [163] and lower incidence of cancer [167] are very promising.

Recent data has shown that the Ras-Erk-ETS pathway is a drug target of longevity [168] with trametinib, a specific inhibitor of the Ras pathway being able to extend lifespan via adult-onset administration [168] in the fruitfly. The ETS transcriptional repressor, Anterior open (Aop) is central to the lifespan extension downstream of reduced insulin/IGF-1 signaling or Ras attenuation [168]. The Ras pathway is highly conserved and these data show promise for anti-aging targets in humans. Finally, inhibition of the TORC1-regulated RNA Polymerase III, an essential enzyme responsible for transcribing tRNAs and 5S ribosomal RNA, extends lifespan in yeast, worms and fruitfly [169]. Pol III inhibition within adult worm or fly guts is sufficient to extend lifespan and results in amelioration of age-related gut deterioration and pathology. These results, place Pol III as an interesting potential target for pharmacological interventions.

Dietary restriction, activation of AMPK through metformin and inhibition of TOR through rapamycin and Torins lead to increase in autophagy. Autophagy is a degradation pathway leading to recycling of cellular material and removal of damaged macromolecules and organelles that may pose a burden to the cell. Autophagy plays a crucial role in cellular homeostasis, development, immunity, tumour suppression and cellular metabolism, prevention of neurodegeneration and lifespan extension. Therefore, pharmacological stimulation of autophagy may be an effective approach for preventing or ameliorating certain human diseases and reducing aging symptoms. Towards identifying new autophagy inducers, high-throughput screens with chemical compound libraries containing around 300,000 compounds are utilised. In a recent screen three candidate molecules have been identified that may be clinically useful as autophagy-inducing agents [170]. Interestingly, a cell-based quantitative high-throughput image screening (qHTS) for autophagy modulators using mouse embryonic fibroblasts (MEFs) has identified (apart from a number of novel autophagy inducers, inhibitors, and modulators) a group of compounds related to dopamine receptors. These compounds are commonly used as clinical psychiatric drugs. These include indatraline hydrochloride (IND), a dopamine inhibitor, and chlorpromazine hydrochloride (CPZ) and fluphenazine dihydrochloride (FPZ), two dopamine receptor antagonists. FPZ-induced autophagy happens through mTOR inhibition but IND and CPZ can induce autophagy via a TOR-independent manner [171]. These results underline once again the importance of revisiting and repurposing already tested and trialled drugs in biogerontology. Further studies on the same theme focus on natural products [172].

While intense biogerontology research is focused on genetic factors it is now evident that epigenetic regulators impact greatly on lifespan. Increasing evidence shows that histone deacetylase (HDAC) inhibitors, able to reverse the deacetylation of histone tails and activate the expression of particular genes, are a promising class of anti-aging compounds that can play major roles for combating age-related diseases [173]. HDAC inhibitors have lifespan-extending effects to preclinical animal models such as fruitflies, C. elegans and rodents [173]. Importantly, preclinical and clinical studies using HDAC inhibitors for age-related conditions have generated very positive outcomes: a wide range of these molecules have emerged as anticancer drugs [174]. A number of them are already approved for specific lymphomas and haematological cancers while others are currently on different stages of clinical development and trials [175, 176]. HDAC inhibitors have also been shown to be beneficial in the contexts of neurodegenerative disease, metabolic and cardiovascular problems as well as inflammatory conditions [173].

Organismal aging has a cellular component and can be viewed as the combination of limited chronological and replicative lifespan of the cells that the organism is composed of. While chronological lifespan is defined as the time that a postmitotic population is viable, replicative lifespan is the number of mitotic divisions that a mother cell can give rise to until senescence. Cellular senescence, first described by Hayflick and Moorhead in the 1960s, is directly linked with decreased telomere length and decreased telomerase activity. During aging and exposure to various intra- and extracellular stresses a rise in senescent cells is observed. If senescent cells are not removed, the senescence state is induced in neighbouring young cells leading to tissue dysfunction [177]. Telomerase activator therapies are currently undergoing with at least one product, TA-65, already available. TA-65 is able to increase telomerase levels and healthspan in mice. Nevertheless, mean and maximal lifespan was not increased [178]. TA-65 has also been reported to decrease senescent immune system cells in patients [179]. Telomerase-based interventions although not fully understood will have no effects on the chronological lifespan of cells within the body (such as postmitotic brain cells). In addition, as telomerase favours tumorigenesis, its long-term efficiency and safety as anti-aging therapy is still questionable [177].

An area of intense interest in drugs with rejuvenating potential related to senescent cells is senolytics. Senolytic drugs (term originating from ‘senescence’ and ‘lysis’) selectively destroy senescent cells from tissues leading to improved health markers in mouse models. The first molecules to be identified as senolytics in 2015 were Dasatinib and Quercetin. Dasatinib, a small molecule targeting BCR/Abl (the Philadelphia chromosome), Src, c-Kit, ephrin receptors, and several other tyrosine kinases. Dasatinib is sold under the brand name Sprycel and is a chemotherapy medication used to treat certain cases of chronic myelogenous leukemia (CML) and acute lymphoblastic leukemia (ALL). Quercetin is a plant flavonol from the group of polyphenols and is found in fruit, vegetables, leaves, and grains. Dasatinib is able to eliminate senescent human fat cell progenitors, while

quercetin is more effective against senescent human endothelial cells. The combination of Dasatinib and Quercetin is effective in eliminating senescent mouse embryonic fibroblasts. Combinations of Dasatinib and Quercetin are shown to reduce senescent cells in aged, and progeroid mice [180]. Navitoclax (ABT-263) was the third senolytic to be identified [181]. Navitoclax is a protein– protein interaction inhibitor targeting the BCL-2 family of apoptotic proteins. The finding that BCL-2 inhibitors can act as senolytics helped towards the identification of analogues A1331852, A1155463 and ABT-737 that inhibit BCL-2 family members, as senolytics [182]. In addition, the action of the flavonol fisetin and the alkaloid piperlongumine as senolytics or senotherapeutics is now established [182, 183, 184]. A FOXO4 peptide able to perturb the FOXO4 interaction with p53 was identified [185]. In senescent cells, this selectively causes p53 nuclear exclusion and cell-intrinsic apoptosis. Under conditions where it was well tolerated in vivo, this FOXO4 peptide neutralized doxorubicin-induced chemotoxicity. Moreover, it restored fitness, fur density, and renal function in both fast aging and naturally aged mice [185]. This is an exciting focus in biogerontology studies with new data amounting on the beneficial effects of senolytics. Nevertheless, their effects on young tissues are not quite clear yet [186]. A characteristic of this research is again the repurposing of known and trialled drugs.

It is more than evident that the field rapidly progresses through the identification of lifespan-increasing genetic and environmental factors. However, the need now is the identification of those interventions that increase human healthspan while not enhancing morbidity periods through clinical trials. Focused drug design studies as well as pre-clinical and clinical trials are of essence to realise the treatments that could delay and ameliorate or even prevent age-related ailments (Table 1).

Table 1

Drugs discussed in the manuscript, their effects on the lifespan of animal aging models and clinical studies in humans related aging and aging phenotype. The effects of somatostatin and pegvisomant on animal aging rates, median and maximal lifespans is unclear. The bar sign (–) within the clinical studies column of the table designates that there are no studies for direct effect on human lifespan to date (data from clinicaltrials.gov as of 30.11.18).

DrugLongevity effects on animalClinical Trials for aging or age-related disease1
modelsTotalCompleted
MetforminYes, positive237
RapamycinYes, positive123
Torin1Yes, positive0-
ResveratrolMixed results1610
Somatostatin-22
Pegvisomant-0-
AspirinYes, positive85

Acknowledgements

We apologize to those not cited due to space limitations. We thank members of the Rallis lab for useful discussions. SG is funded from a UEL PhD Studentship to CR.

References

1 GBD 2015 Mortality and Causes of Death Collaborators (2016). Global, regional, and national life expectancy, all-cause mortality, and cause-specific mortality for 249 causes of death, 1980-2015: a systematic analysis for the Global Burden of Disease Study 2015. Lancet (London, England) 388, 1459-1544.10.1016/S0140-6736(16)31012-1Search in Google Scholar

2 Caley, M., and Sidhu, K. (2011). Estimating the future healthcare costs of an aging population in the UK: expansion of morbidity and the need for preventative care. Journal of public health (Oxford, England) 33, 117-122.10.1093/pubmed/fdq044Search in Google Scholar PubMed

3 Blagosklonny, M.V. (2013). Aging is not programmed: genetic pseudo-program is a shadow of developmental growth. Cell cycle (Georgetown, Tex.) 12, 3736-3742.10.4161/cc.27188Search in Google Scholar PubMed PubMed Central

4 Freitas, A.A., and de Magalhaes, J.P. (2011). A review and appraisal of the DNA damage theory of ageing. Mutation research 728, 12-22.10.1016/j.mrrev.2011.05.001Search in Google Scholar PubMed

5 Liguori, I., Russo, G., Curcio, F., Bulli, G., Aran, L., Della-Morte, D., Gargiulo, G., Testa, G., Cacciatore, F., Bonaduce, D., et al. (2018). Oxidative stress, aging, and diseases. Clinical interventions in aging 13, 757-772.10.2147/CIA.S158513Search in Google Scholar PubMed PubMed Central

6 Lipsky, M.S., and King, M. (2015). Biological theories of aging. Disease-a-month : DM 61, 460-466.10.1016/j.disamonth.2015.09.005Search in Google Scholar PubMed

7 Mazucanti, C.H., Cabral-Costa, J.V., Vasconcelos, A.R., Andreotti, D.Z., Scavone, C., and Kawamoto, E.M. (2015). Longevity Pathways (mTOR, SIRT, Insulin/IGF-1) as Key Modulatory Targets on Aging and Neurodegeneration. Current topics in medicinal chemistry 15, 2116-2138.10.2174/1568026615666150610125715Search in Google Scholar PubMed

8 Vezina, C., Kudelski, A., and Sehgal, S.N. (1975). Rapamycin (AY-22,989), a new antifungal antibiotic. I. Taxonomy of the producing streptomycete and isolation of the active principle. The Journal of antibiotics 28, 721-726.10.7164/antibiotics.28.721Search in Google Scholar PubMed

9 Blenis, J. (2017). TOR, the Gateway to Cellular Metabolism, Cell Growth, and Disease. Cell 171, 10-13.10.1016/j.cell.2017.08.019Search in Google Scholar PubMed

10 Ehninger, D., Neff, F., and Xie, K. (2014). Longevity, aging and rapamycin. Cellular and Molecular Life Sciences 71, 4325-4346.10.1007/s00018-014-1677-1Search in Google Scholar PubMed PubMed Central

11 Laplante, M., and Sabatini, D.M. (2009). mTOR signaling at a glance. Journal of cell science 122, 3589-3594.10.1242/jcs.051011Search in Google Scholar

12 Heitman, J., Movva, N.R., and Hall, M.N. (1991). Targets for cell cycle arrest by the immunosuppressant rapamycin in yeast. Science (New York, N.Y.) 253, 905-909.10.1126/science.1715094Search in Google Scholar

13 Heitman, J., Movva, N.R., Hiestand, P.C., and Hall, M.N. (1991). FK 506-binding protein proline rotamase is a target for the immunosuppressive agent FK 506 in Saccharomyces cerevisiae. Proceedings of the National Academy of Sciences of the United States of America 88, 1948-1952.10.1073/pnas.88.5.1948Search in Google Scholar

14 Kunz, J., Henriquez, R., Schneider, U., Deuter-Reinhard, M., Movva, N.R., and Hall, M.N. (1993). Target of rapamycin in yeast, TOR2, is an essential phosphatidylinositol kinase homolog required for G1 progression. Cell 73, 585-596.10.1016/0092-8674(93)90144-FSearch in Google Scholar

15 Helliwell, S.B., Wagner, P., Kunz, J., Deuter-Reinhard, M., Henriquez, R., and Hall, M.N. (1994). TOR1 and TOR2 are structurally and functionally similar but not identical phosphatidylinositol kinase homologues in yeast. Molecular biology of the cell 5, 105-118.10.1091/mbc.5.1.105Search in Google Scholar

16 Shertz, C.A., Bastidas, R.J., Li, W., Heitman, J., and Cardenas, M.E. (2010). Conservation, duplication, and loss of the Tor signaling pathway in the fungal kingdom. BMC genomics 11, 510.10.1186/1471-2164-11-510Search in Google Scholar

17 Brown, E.J., Albers, M.W., Shin, T.B., Ichikawa, K., Keith, C.T., Lane, W.S., and Schreiber, S.L. (1994). A mammalian protein targeted by G1-arresting rapamycin-receptor complex. Nature 369, 756-758.10.1038/369756a0Search in Google Scholar

18 Abraham, R.T. (1998). Mammalian target of rapamycin: immunosuppressive drugs uncover a novel pathway of cytokine receptor signaling. Current opinion in immunology 10, 330-336.10.1016/S0952-7915(98)80172-6Search in Google Scholar

19 Fok, W.C., Chen, Y., Bokov, A., Zhang, Y., Salmon, A.B., Diaz, V., Javors, M., Wood, W.H., 3rd, Zhang, Y., Becker, K.G., et al. (2014). Mice fed rapamycin have an increase in lifespan associated with major changes in the liver transcriptome. PloS one 9, e83988.10.1371/journal.pone.0083988Search in Google Scholar PubMed PubMed Central

20 Harrison, D.E., Strong, R., Sharp, Z.D., Nelson, J.F., Astle, C.M., Flurkey, K., Nadon, N.L., Wilkinson, J.E., Frenkel, K., Carter, C.S., et al. (2009). Rapamycin fed late in life extends lifespan in genetically heterogeneous mice. Nature 460, 392-395.10.1038/nature08221Search in Google Scholar PubMed PubMed Central

21 Johnson, S.C., and Kaeberlein, M. (2016). Rapamycin in aging and disease: maximizing efficacy while minimizing side effects. Oncotarget 7, 44876-44878.10.18632/oncotarget.10381Search in Google Scholar PubMed PubMed Central

22 Kennedy, B.K., and Lamming, D.W. (2016). The Mechanistic Target of Rapamycin: The Grand ConducTOR of Metabolism and Aging. Cell metabolism 23, 990-1003.10.1016/j.cmet.2016.05.009Search in Google Scholar PubMed PubMed Central

23 Miller, R.A., Harrison, D.E., Astle, C.M., Fernandez, E., Flurkey, K., Han, M., Javors, M.A., Li, X., Nadon, N.L., Nelson, J.F., et al. (2014). Rapamycin-mediated lifespan increase in mice is dose and sex dependent and metabolically distinct from dietary restriction. Aging Cell 13, 468-477.10.1111/acel.12194Search in Google Scholar PubMed PubMed Central

24 Neff, F., Flores-Dominguez, D., Ryan, D.P., Horsch, M., Schroder, S., Adler, T., Afonso, L.C., Aguilar-Pimentel, J.A., Becker, L., Garrett, L., et al. (2013). Rapamycin extends murine lifespan but has limited effects on aging. The Journal of clinical investigation 123, 3272-3291.10.1172/JCI67674Search in Google Scholar PubMed PubMed Central

25 Xie, J., Wang, X., and Proud, C.G. (2016). mTOR inhibitors in cancer therapy. F1000Research 5, F1000 Faculty Rev-2078.10.12688/f1000research.9207.1Search in Google Scholar PubMed PubMed Central

26 Schreiber, K.H., Ortiz, D., Academia, E.C., Anies, A.C., Liao, C.Y., and Kennedy, B.K. (2015). Rapamycin-mediated mTORC2 inhibition is determined by the relative expression of FK506-binding proteins. Aging Cell 14, 265-273.10.1111/acel.12313Search in Google Scholar PubMed PubMed Central

27 Johnson, S.C., Rabinovitch, P.S., and Kaeberlein, M. (2013). mTOR is a key modulator of ageing and age-related disease. Nature 493, 338-345.10.1038/nature11861Search in Google Scholar PubMed PubMed Central

28 Chung, H.Y., Cesari, M., Anton, S., Marzetti, E., Giovannini, S., Seo, A.Y., Carter, C., Yu, B.P., and Leeuwenburgh, C. (2009). Molecular Inflammation: Underpinnings of Aging and Age-related Diseases. Ageing research reviews 8, 18-30.10.1016/j.arr.2008.07.002Search in Google Scholar PubMed PubMed Central

29 Van Skike, C.E., and Galvan, V. (2018). A Perfect sTORm: The Role of the Mammalian Target of Rapamycin (mTOR) in Cerebrovascular Dysfunction of Alzheimer’s Disease: A Mini-Review. Gerontology 64, 205-211.10.1159/000485381Search in Google Scholar PubMed PubMed Central

30 Humpel, C. (2011). Chronic mild cerebrovascular dysfunction as a cause for Alzheimer’s disease? Experimental Gerontology 46, 225-232.10.1016/j.exger.2010.11.032Search in Google Scholar PubMed PubMed Central

31 Kim, J.-A., Jang, H.-J., Martinez-Lemus, L.A., and Sowers, J.R. (2012). Activation of mTOR/p70S6 kinase by ANG II inhibits insulin-stimulated endothelial nitric oxide synthase and vasodilation. Am J Physiol Endocrinol Metab 302, E201-208.10.1152/ajpendo.00497.2011Search in Google Scholar PubMed PubMed Central

32 Kurdi, A., De Meyer, G.R.Y., and Martinet, W. (2016). Potential therapeutic effects of mTOR inhibition in atherosclerosis. British Journal of Clinical Pharmacology 82, 1267-1279.10.1111/bcp.12820Search in Google Scholar PubMed PubMed Central

33 Kurdi, A., Martinet, W., and De Meyer, G.R.Y. (2018). mTOR Inhibition and Cardiovascular Diseases: Dyslipidemia and Atherosclerosis. Transplantation 102, S44-s46.10.1097/TP.0000000000001693Search in Google Scholar PubMed

34 Stone, G.W., Rizvi, A., Newman, W., Mastali, K., Wang, J.C., Caputo, R., Doostzadeh, J., Cao, S., Simonton, C.A., Sudhir, K., et al. (2010). Everolimus-eluting versus paclitaxel-eluting stents in coronary artery disease. The New England journal of medicine 362, 1663-1674.10.1056/NEJMoa0910496Search in Google Scholar PubMed

35 Rodriguez, A.E., Granada, J.F., Rodriguez-Alemparte, M., Vigo, C.F., Delgado, J., Fernandez-Pereira, C., Pocovi, A., Rodriguez-Granillo, A.M., Schulz, D., Raizner, A.E., et al. (2006). Oral rapamycin after coronary bare-metal stent implantation to prevent restenosis: the Prospective, Randomized Oral Rapamycin in Argentina (ORAR II) Study. Journal of the American College of Cardiology 47, 1522-1529.10.1016/j.jacc.2005.12.052Search in Google Scholar PubMed

36 Thoreen, C.C., Kang, S.A., Chang, J.W., Liu, Q., Zhang, J., Gao, Y., Reichling, L.J., Sim, T., Sabatini, D.M., and Gray, N.S. (2009). An ATP-competitive mammalian target of rapamycin inhibitor reveals rapamycin-resistant functions of mTORC1. The Journal of biological chemistry 284, 8023-8032.10.1074/jbc.M900301200Search in Google Scholar PubMed PubMed Central

37 Liu, Q., Xu, C., Kirubakaran, S., Zhang, X., Hur, W., Liu, Y., Kwiatkowski, N.P., Wang, J., Westover, K.D., Gao, P., et al. (2013). Characterization of Torin2, an ATP-competitive inhibitor of mTOR, ATM, and ATR. Cancer research 73, 2574-2586.10.1158/0008-5472.CAN-12-1702Search in Google Scholar PubMed PubMed Central

38 Leontieva, O.V., and Blagosklonny, M.V. (2016). Gerosuppression by pan-mTOR inhibitors. Aging 8, 3535-3551.10.18632/aging.101155Search in Google Scholar PubMed PubMed Central

39 Leontieva, O.V., Demidenko, Z.N., and Blagosklonny, M.V. (2015). Dual mTORC1/C2 inhibitors suppress cellular geroconversion (a senescence program). Oncotarget 6, 23238-23248.10.18632/oncotarget.4836Search in Google Scholar PubMed PubMed Central

40 Mason, J.S., Wileman, T., and Chapman, T. (2018). Lifespan extension without fertility reduction following dietary addition of the autophagy activator Torin1 in Drosophila melanogaster. 13, e0190105.10.1371/journal.pone.0190105Search in Google Scholar PubMed PubMed Central

41 Xie, J., and Proud, C.G. (2014). Signaling crosstalk between the mTOR complexes. Translation (Austin, Tex.) 2, e28174.10.4161/trla.28174Search in Google Scholar

42 Bhullar KS, H.B. (2015). Lifespan and healthspan extension by resveratrol. Biochimica et Biophysica Acta 1852, 1209-1218. 1210.1016/j.bbadis.2015.1201.1012Search in Google Scholar

43 Baur JA, S.D. (2006). Therapeutic potential of resveratrol: the in vivo evidence. Nature Reviews Drug Discovery 5, 493-506. Epub 2006 May 2026.10.1038/nrd2060Search in Google Scholar

44 Paolini M, S.A., Valgimigli L (2003). Avoidance of bioflavonoid supplements during pregnancy: a pathway to infant leukemia? Mutation Research 19, 99-101.10.1016/S0027-5107(03)00057-5Search in Google Scholar

45 Wang Y, C.H., Yu O. (2010). Metabolic engineering of resveratrol and other longevity boosting compounds. BioFactors 36, 394-400. 310.1002/biof.1126Search in Google Scholar

46 Mattivi F, R.F., Korhammer S (1995). Isolation, characterization, and evolution in red wine vinification of resveratrol monomers. Journal of Agricultural and Food Chemistry 43, 1820–1823.10.1021/jf00055a013Search in Google Scholar

47 Cao H, P.X., Li C, Zhou C, Deng F, Li T (2003). Density functional theory calculations for resveratrol. Bioorganic & Medicinal Chemistry Letters 13, 1869-1871.10.1016/S0960-894X(03)00283-XSearch in Google Scholar

48 Queiroz AN, G.B., Moraes WM Jr, Borges RS (2009). A theoretical antioxidant pharmacophore for resveratrol. European Journal of Medicinal Chemistry 44, 1644-1649. 1610.1016/j.ejmech.2008.1609.1023 Epub 2008 Sep 1630Search in Google Scholar

49 Frankel EN, W.A., Kinsella JE (1993). Inhibition of human LDL oxidation by resveratrol. Lancet 341, 1103-1104.10.1016/0140-6736(93)92472-6Search in Google Scholar

50 Tadolini B, J.C., Piu L, Franconi F, Cabrini L (2000). Resveratrol inhibition of lipid peroxidation. Free Radical Research 33, 105-114.10.1080/10715760000300661Search in Google Scholar PubMed

51 Rotondo S, R.G., Manarini S, Celardo A, Rotillo D, de Gaetano G, Evangelista V, Cerletti C (1998). Effect of trans-resveratrol, a natural polyphenolic compound, on human polymorphonuclear leukocyte function. British Journal of Pharmacology 123, 1691-1699.10.1038/sj.bjp.0701784Search in Google Scholar

52 Jang M, C.L., Udeani GO, Slowing KV, Thomas CF, Beecher CW, Fong HH, Farnsworth NR, Kinghorn AD, Mehta RG, Moon RC, Pezzuto JM (1997). Cancer chemopreventive activity of resveratrol, a natural product derived from grapes. Science 275, 218-220.10.1126/science.275.5297.218Search in Google Scholar

53 Virgili M, C.A. (2000). Partial neuroprotection of in vivo excitotoxic brain damage by chronic administration of the red wine antioxidant agent, trans-resveratrol in rats. Neuroscience Letters 281, 123-126.10.1016/S0304-3940(00)00820-XSearch in Google Scholar

54 Wilson T, K.T., Beitz DC, Lewis DS, Engen RL (1996). Resveratrol promotes atherosclerosis in hypercholesterolemic rabbits. Life Sciences 59, PL15-21.10.1016/0024-3205(96)00260-3Search in Google Scholar

55 Breinholt VM, M.A., Svendsen GW, Daneshvar B, Vinggaard AM, Poulsen M, Dragsted LO (2003). Effects of dietary antioxidants and 2-amino-3-methylimidazo[4,5-f]- quinoline (IQ) on preneoplastic lesions and on oxidative damage, hormonal status and detoxification capacity in the rat. Food and Chemical Toxicology 41, 1315-1323.10.1016/S0278-6915(03)00122-4Search in Google Scholar

56 Murias M, H.N., Erker T, Pleban K, Ecker G, Saiko P, Szekeres T, Jäger W (2004). Resveratrol analogues as selective cyclooxygenase-2 inhibitors: synthesis and structure-activity relationship. Bioorganic & Medicinal Chemistry 12, 5571-5578.10.1016/j.bmc.2004.08.008Search in Google Scholar

57 Yu C, S.Y., Kosmeder JW, Pezzuto JM, van Breemen RB (2003). Liquid chromatography/tandem mass spectrometric determination of inhibition of human cytochrome P450 isozymes by resveratrol and resveratrol-3-sulfate. Rapid Communications in Mass Spectrometry 17, 307-313.10.1002/rcm.918Search in Google Scholar

58 Gliemann L, N.M., Hellsten Y (2016). Effects of exercise training and resveratrol on vascular health in aging. Free Radical Biology & Medicine 98, 165–176. 110.1016/j.freeradbiomed.2016.1003.1037 Epub 2016 Apr 1013Search in Google Scholar

59 Haworth RS, A.M. (2001). Inhibition of protein kinase D by resveratrol. Biochemical Pharmacology 62, 1647-1651.10.1016/S0006-2952(01)00807-3Search in Google Scholar

60 Armour SM, B.J., Hsieh SN, Land-Bracha A, Thomas SM, Sinclair DA (2009). Inhibition of mammalian S6 kinase by resveratrol suppresses autophagy. Aging (Albany NY) 1, 515–528. 510.18632/aging.100056Epub 102009 Jun 100053Search in Google Scholar

61 Denu, J. (2005). The Sir 2 family of protein deacetylases. 9, 431-440.Search in Google Scholar

62 Li YR, L.S., Lin CC (2018). Effect of resveratrol and pterostilbene on aging and longevity. BioFactors 44, 69-82. 10.1002/biof.1400 Epub 2017 Dec 1006Search in Google Scholar

63 Lagouge M, A.C., Gerhart-Hines Z, Meziane H, Lerin C, Daussin F, Messadeq N, Milne J, Lambert P, Elliott P, Geny B, Laakso M, Puigserver P, Auwerx J (2006). Resveratrol improves mitochondrial function and protects against metabolic disease by activating SIRT1 and PGC-1alpha. Cell 127, 1109-1122. Epub 2006 Nov 1116.10.1016/j.cell.2006.11.013Search in Google Scholar

64 Ono, M. (2008). Molecular links between tumor angiogenesis and inflammation: inflammatory stimuli of macrophages and cancer cells as targets for therapeutic strategy. Cancer Science 99, 1501-1506. 1510.1111/j.1349-7006.2008.00853.xSearch in Google Scholar

65 Kundu JK, S.Y., Surh YJ (2006). Resveratrol modulates phorbol ester-induced pro-inflammatory signal transduction pathways in mouse skin in vivo: NF-kappaB and AP-1 as prime targets. Biochemical Pharmacology 72, 1506-1515. Epub 2006 Sep 1526.10.1016/j.bcp.2006.08.005Search in Google Scholar

66 Borradaile NM, P.J. (2009). NAD(+), sirtuins, and cardiovascular disease. Current Pharmaceutical Design 15, 110-117.10.2174/138161209787185742Search in Google Scholar

67 Wang Y, Y.O. (2012). Synthetic scaffolds increased resveratrol biosynthesis in engineered yeast cells. Journal of Biotechnology 157, 258-260. 210.1016/j.jbiotec.2011.1011.1003 Epub 2011 Nov 1019Search in Google Scholar

68 Sinclair DA, G.L. (1997). Extrachromosomal rDNA circles--a cause of aging in yeast. Cell 91, 1033-1042.10.1016/S0092-8674(00)80493-6Search in Google Scholar

69 Kaeberlein M, M.M., Guarente L (1999). The SIR2/3/4 complex and SIR2 alone promote longevity in Saccharomyces cerevisiae by two different mechanisms. Genes & Development 13, 2570-2580.10.1101/gad.13.19.2570Search in Google Scholar PubMed PubMed Central

70 Howitz KT, B.K., Cohen HY, Lamming DW, Lavu S, Wood JG, Zipkin RE, Chung P, Kisielewski A, Zhang LL, Scherer B, Sinclair DA (2003). Small molecule activators of sirtuins extend Saccharomyces cerevisiae lifespan. Nature 425, 191-196. Epub 2003 Aug 2024.10.1038/nature01960Search in Google Scholar PubMed

71 Guarente L, P.F. (2005). Calorie restriction--the SIR2 connection. Cell 120, 473-482.10.1016/j.cell.2005.01.029Search in Google Scholar PubMed

72 Tissenbaum HA, G.L. (2001). Increased dosage of a sir-2 gene extends lifespan in Caenorhabditis elegans. Nature 410, 227-230.10.1038/35065638Search in Google Scholar PubMed

73 Rogina B, H.S. (2004). Sir2 mediates longevity in the fly through a pathway related to calorie restriction. Proceedings of the National Academy of Sciences of the USA 101, 15998-16003. Epub 12004 Nov 15991.10.1073/pnas.0404184101Search in Google Scholar PubMed PubMed Central

74 Valenzano DR, T.E., Genade T, Cattaneo A, Domenici L, Cellerino A (2006). Resveratrol prolongs lifespan and retards the onset of age-related markers in a short-lived vertebrate. Current Biology 16, 296-300.10.1016/j.cub.2005.12.038Search in Google Scholar PubMed

75 Yu X, L.G. (2012). Effects of resveratrol on longevity, cognitive ability and aging-related histological markers in the annual fish Nothobranchius guentheri. Experimental Gerontology. Dec 47, 940-949. 910.1016/j.exger.2012.1008.1009 Epub 2012 Aug 1013Search in Google Scholar

76 Liu S, Z.Z., Ji S1, Liu T1, Hou Y1, Li S1, Li G (2018). Resveratrol reduces senescence-associated secretory phenotype by SIRT1/NF-κB pathway in gut of the annual fish Nothobranchius guentheri. Fish & Shellfish Immunology 80, 473-479. 410.1016/j.fsi.2018.1006.1027 Epub 2018 Jun 1013Search in Google Scholar

77 Jimenez-Gomez Y, M.J., Pearson KJ, Martin-Montalvo A, Palacios HH, Sossong AM, Ward TM, Younts CM, Lewis K, Allard JS, Longo DL, Belman JP, Malagon MM, Navas P, Sanghvi M, Moaddel R, Tilmont EM, Herbert RL, Morrell CH, E.J., Baur JA, Ferrucci L, Bogan JS, and Bernier M, d.C.R. (2013). Resveratrol improves adipose insulin signaling and reduces the inflammatory response in adipose tissue of rhesus monkeys on high-fat, high-sugar diet. Cell Metabolism 18, 533-545. 510.1016/j.cmet.2013.1009.1004Search in Google Scholar

78 Burnett, C., Valentini, S., Cabreiro, F., Goss, M., Somogyvari, M., Piper, M.D., Hoddinott, M., Sutphin, G.L., Leko, V., McElwee, J.J., et al. (2011). Absence of effects of Sir2 overexpression on lifespan in C. elegans and Drosophila. Nature 477, 482-485.10.1038/nature10296Search in Google Scholar PubMed PubMed Central

79 Price NL, G.A., Ling AJ, Duarte FV, Martin-Montalvo A, North BJ, Agarwal B, Ye L, Ramadori G, Teodoro JS, Hubbard BP, Varela AT, Davis JG, Varamini B, Hafner A, Moaddel R, Rolo AP, Coppari R, Palmeira CM, de Cabo R, Baur JA, Sinclair DA (2012). SIRT1 is required for AMPK activation and the beneficial effects of resveratrol on mitochondrial function. Cell Metabolism 15, 675-690. 610.1016/j.cmet.2012.1004.1003Search in Google Scholar

80 Hubbard BP, G.A., Dai H, Li J, Case AW, Considine T, Riera TV, Lee JE, E SY, Lamming DW, Pentelute BL, Schuman ER, Stevens LA, Ling AJ, Armour SM, Michan S, Zhao H, Jiang Y, Sweitzer SM, Blum CA, Disch JS, Ng PY, Howitz KT, Rolo AP, Hamuro Y, Moss J, and Perni RB, E.J., Vlasuk GP, Sinclair DA (2013). Evidence for a common mechanism of SIRT1 regulation by allosteric activators. Science 339, 1216-1219. 1210.1126/science.1231097Search in Google Scholar

81 Lakshminarasimhan M, R.D., Schutkowski M, Steegborn C (2013). Sirt1 activation by resveratrol is substrate sequence-selective. Aging (Albany NY) 5, 151-154.10.18632/aging.100542Search in Google Scholar PubMed PubMed Central

82 Kim D, N.M., Dobbin MM, Fischer A, Sananbenesi F, Rodgers JT, Delalle I, Baur JA, Sui G, Armour SM, Puigserver P, Sinclair DA, Tsai LH (2007). SIRT1 deacetylase protects against neurodegeneration in models for Alzheimer’s disease and amyotrophic lateral sclerosis. The EMBO Journal 26, 3169-3179. Epub 2007 Jun 3121.10.1038/sj.emboj.7601758Search in Google Scholar PubMed PubMed Central

83 Liu T, L.P., Marshall GM (2009). The critical role of the class III histone deacetylase SIRT1 in cancer. Cancer Research 69, 1702-1705. 1710.1158/0008-5472.CAN-1708-3365 Epub 2009 Feb 1724Search in Google Scholar

84 Milne JC, L.P., Schenk S, Carney DP, Smith JJ, Gagne DJ, Jin L, Boss O, Perni RB, Vu CB, Bemis JE, Xie R, Disch JS, Ng PY, Nunes JJ, Lynch AV, Yang H, Galonek H, Israelian K, Choy W, Iffland A, Lavu S, Medvedik O, Sinclair DA, Olefsky JM, and Jirousek MR, E.P., Westphal CH (2007). Small molecule activators of SIRT1 as therapeutics for the treatment of type 2 diabetes. Nature 450, 712-716.10.1038/nature06261Search in Google Scholar PubMed PubMed Central

85 Howitz KT, S.D. (2008). Xenohormesis: sensing the chemical cues of other species. Cell 133, 387-391.310.1016/j.cell.2008.1004.1019Search in Google Scholar

86 Pallauf K, R.G., Rupp PM, Chin D, Wolf IM (2016). Resveratrol and Lifespan in Model Organisms. Current Medicinal Chemistry 23, 4639-4680.10.2174/0929867323666161024151233Search in Google Scholar PubMed

87 Zadik, Z., Chalew, S.A., McCarter, R.J., Jr., Meistas, M., and Kowarski, A.A. (1985). The influence of age on the 24-hour integrated concentration of growth hormone in normal individuals. The Journal of clinical endocrinology and metabolism 60, 513-516.10.1210/jcem-60-3-513Search in Google Scholar PubMed

88 Junnila, R.K., List, E.O., Berryman, D.E., Murrey, J.W., and Kopchick, J.J. (2013). The GH/IGF-1 axis in ageing and longevity. Nature reviews. Endocrinology 9, 366-376.10.1038/nrendo.2013.67Search in Google Scholar PubMed PubMed Central

89 Liu, H., Bravata, D.M., Olkin, I., Nayak, S., Roberts, B., Garber, A.M., and Hoffman, A.R. (2007). Systematic review: the safety and efficacy of growth hormone in the healthy elderly. Annals of internal medicine 146, 104-115.10.7326/0003-4819-146-2-200701160-00005Search in Google Scholar PubMed

90 Rudman, D., Feller, A.G., Nagraj, H.S., Gergans, G.A., Lalitha, P.Y., Goldberg, A.F., Schlenker, R.A., Cohn, L., Rudman, I.W., and Mattson, D.E. (1990). Effects of human growth hormone in men over 60 years old. The New England journal of medicine 323, 1-6.10.1056/NEJM199007053230101Search in Google Scholar PubMed

91 Vestergaard, P., Jorgensen, J.O., Olesen, J.L., Bosnjak, E., Holm, L., Frystyk, J., Langberg, H., Kjaer, M., and Hansen, M. (2012). Local administration of growth hormone stimulates tendon collagen synthesis in elderly men. Journal of applied physiology (Bethesda, Md. : 1985) 113, 1432-1438.10.1152/japplphysiol.00816.2012Search in Google Scholar PubMed

92 Chanson, P., and Salenave, S. (2008). Acromegaly. Orphanet journal of rare diseases 3, 17.10.1016/B978-0-444-59602-4.00014-9Search in Google Scholar PubMed

93 Ayuk, J., and Sheppard, M.C. (2008). Does acromegaly enhance mortality? Reviews in endocrine & metabolic disorders 9, 33-39.10.1007/s11154-007-9067-8Search in Google Scholar PubMed

94 Neggers, S.J., Franck, S.E., de Rooij, F.W., Dallenga, A.H., Poublon, R.M., Feelders, R.A., Janssen, J.A., Buchfelder, M., Hofland, L.J., Jorgensen, J.O., et al. (2014). Long-term efficacy and safety of pegvisomant in combination with long-acting somatostatin analogs in acromegaly. The Journal of clinical endocrinology and metabolism 99, 3644-3652.10.1210/jc.2014-2032Search in Google Scholar PubMed

95 Freda, P.U., Gordon, M.B., Kelepouris, N., Jonsson, P., Koltowska-Haggstrom, M., and van der Lely, A.J. (2015). Long-term treatment with pegvisomant as monotherapy in patients with acromegaly: experience from ACROSTUDY. Endocrine practice : official journal of the American College of Endocrinology and the American Association of Clinical Endocrinologists 21, 264-274.10.4158/EP14330.ORSearch in Google Scholar PubMed PubMed Central

96 Hansen, M., and Kennedy, B.K. (2016). Does Longer Lifespan Mean Longer Healthspan? Trends in cell biology 26, 565-568.10.1016/j.tcb.2016.05.002Search in Google Scholar PubMed PubMed Central

97 Mao, K., Quipildor, G.F., Tabrizian, T., Novaj, A., Guan, F., Walters, R.O., Delahaye, F., Hubbard, G.B., Ikeno, Y., Ejima, K., et al. (2018). Late-life targeting of the IGF-1 receptor improves healthspan and lifespan in female mice. Nature communications 9, 2394.10.1038/s41467-018-04805-5Search in Google Scholar PubMed PubMed Central

98 Milman, S., and Barzilai, N. (2015). Dissecting the Mechanisms Underlying Unusually Successful Human Health Span and Life Span. Cold Spring Harbor perspectives in medicine 6, a025098.10.1101/cshperspect.a025098Search in Google Scholar PubMed PubMed Central

99 Custodero C, M.R., Lee SA, Chen Z, Wu S, Manini TM, Hincapie Echeverri J, Sabbà C, Beavers DP, Cauley JA, Espeland MA, Fielding RA, Kritchevsky SB, Liu CK, McDermott MM, Miller ME, Tracy RP, Newman AB, Ambrosius WT, Pahor M, Anton SD (2018). Evidence-based nutritional and pharmacological interventions targeting chronic low-grade inflammation in middle-age and older adults: A systematic review and meta-analysis. Ageing Research Review 46, 42-59. 10.1016/j.arr.2018.1005.1004 Epub ahead of printSearch in Google Scholar

100 Rena G, H.D., Pearson ER (2017). The mechanisms of action of metformin. Diabetologia 60, 1577-1585. doi:1510.1007/s00125-00017-04342-z. Epub 02017 Aug 00123.10.1007/s00125-017-4342-zSearch in Google Scholar PubMed PubMed Central

101 Bailey CJ, D.C. (2004). Metformin: its botanical background. Practical Diabetes International 21, 115–117.10.1002/pdi.606Search in Google Scholar

102 Hundal RS, K.M., Dufour S, Laurent D, Lebon V, Chandramouli V, Inzucchi SE, Schumann WC, Petersen KF, Landau BR, Shulman GI (2000). Mechanism by which metformin reduces glucose production in type 2 diabetes. Diabetes 49, 2063-2069.10.2337/diabetes.49.12.2063Search in Google Scholar PubMed PubMed Central

103 Takashima M1, O.W., Hayashi K, Inoue H, Kinoshita S, Okamoto Y, Sakaue H, Wataoka Y, Emi A, Senga Y, Matsuki Y, Watanabe E, Hiramatsu R, Kasuga M (2010). Role of KLF15 in regulation of hepatic gluconeogenesis and metformin action. Diabetes 59, 1608-1615. 1610.2337/db1609-1679 Epub 2010 Apr 1614Search in Google Scholar

104 Garcia, D., and Shaw, R.J. (2017). AMPK: Mechanisms of Cellular Energy Sensing and Restoration of Metabolic Balance. Molecular cell 66, 789-800.10.1016/j.molcel.2017.05.032Search in Google Scholar PubMed PubMed Central

105 Lage, R., Dieguez, C., Vidal-Puig, A., and Lopez, M. (2008). AMPK: a metabolic gauge regulating whole-body energy homeostasis. Trends in molecular medicine 14, 539-549.10.1016/j.molmed.2008.09.007Search in Google Scholar PubMed

106 Steinberg, G.R., and Kemp, B.E. (2009). AMPK in Health and Disease. Physiological reviews 89, 1025-1078.10.1152/physrev.00011.2008Search in Google Scholar PubMed

107 Hardie, D.G. (2014). AMPK--sensing energy while talking to other signaling pathways. Cell metabolism 20, 939-952.10.1016/j.cmet.2014.09.013Search in Google Scholar PubMed PubMed Central

108 Kurumbail, R.G., and Calabrese, M.F. (2016). Structure and Regulation of AMPK. Exs 107, 3-22.10.1007/978-3-319-43589-3_1Search in Google Scholar PubMed

109 Salminen, A., Kaarniranta, K., and Kauppinen, A. (2016). Age-related changes in AMPK activation: Role for AMPK phosphatases and inhibitory phosphorylation by upstream signaling pathways. Ageing research reviews 28, 15-26.10.1016/j.arr.2016.04.003Search in Google Scholar PubMed

110 Calabrese, M.F., Rajamohan, F., Harris, M.S., Caspers, N.L., Magyar, R., Withka, J.M., Wang, H., Borzilleri, K.A., Sahasrabudhe, P.V., Hoth, L.R., et al. (2014). Structural basis for AMPK activation: natural and synthetic ligands regulate kinase activity from opposite poles by different molecular mechanisms. Structure (London, England : 1993) 22, 1161-1172.10.1016/j.str.2014.06.009Search in Google Scholar PubMed

111 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. (2011). Structure of mammalian AMPK and its regulation by ADP. Nature 472, 230-233.10.1038/nature09932Search in Google Scholar PubMed PubMed Central

112 Mihaylova, M.M., and Shaw, R.J. (2011). The AMPK signalling pathway coordinates cell growth, autophagy and metabolism. Nature cell biology 13, 1016-1023.10.1038/ncb2329Search in Google Scholar PubMed PubMed Central

113 Dong, Y., Zhang, M., Liang, B., Xie, Z., Zhao, Z., Asfa, S., Choi, H.C., and Zou, M.H. (2010). Reduction of AMP-activated protein kinase alpha2 increases endoplasmic reticulum stress and atherosclerosis in vivo. Circulation 121, 792-803.10.1161/CIRCULATIONAHA.109.900928Search in Google Scholar PubMed PubMed Central

114 Li, X.N., Song, J., Zhang, L., LeMaire, S.A., Hou, X., Zhang, C., Coselli, J.S., Chen, L., Wang, X.L., Zhang, Y., et al. (2009). Activation of the AMPK-FOXO3 pathway reduces fatty acid-induced increase in intracellular reactive oxygen species by upregulating thioredoxin. Diabetes 58, 2246-2257.10.2337/db08-1512Search in Google Scholar PubMed PubMed Central

115 Salminen, A., Hyttinen, J.M., and Kaarniranta, K. (2011). AMP-activated protein kinase inhibits NF-kappaB signaling and inflammation: impact on healthspan and lifespan. Journal of molecular medicine (Berlin, Germany) 89, 667-676.10.1007/s00109-011-0748-0Search in Google Scholar PubMed PubMed Central

116 Curtis, R., O’Connor, G., and DiStefano, P.S. (2006). Aging networks in Caenorhabditis elegans: AMP-activated protein kinase (aak-2) links multiple aging and metabolism pathways. Aging Cell 5, 119-126.10.1111/j.1474-9726.2006.00205.xSearch in Google Scholar PubMed

117 Funakoshi, M., Tsuda, M., Muramatsu, K., Hatsuda, H., Morishita, S., and Aigaki, T. (2011). A gain-of-function screen identifies wdb and lkb1 as lifespan-extending genes in Drosophila. Biochemical and biophysical research communications 405, 667-672.10.1016/j.bbrc.2011.01.090Search in Google Scholar PubMed

118 Viollet, B., Andreelli, F., Jorgensen, S.B., Perrin, C., Flamez, D., Mu, J., Wojtaszewski, J.F., Schuit, F.C., Birnbaum, M., Richter, E., et al. (2003). Physiological role of AMP-activated protein kinase (AMPK): insights from knockout mouse models. Biochemical Society transactions 31, 216-219.10.1042/bst0310216Search in Google Scholar PubMed

119 Richter, E.A., and Ruderman, N.B. (2009). AMPK and the biochemistry of exercise: implications for human health and disease. The Biochemical journal 418, 261-275.10.1042/BJ20082055Search in Google Scholar PubMed PubMed Central

120 Dyck, J.R., and Lopaschuk, G.D. (2006). AMPK alterations in cardiac physiology and pathology: enemy or ally? The Journal of physiology 574, 95-112.10.1113/jphysiol.2006.109389Search in Google Scholar PubMed PubMed Central

121 McCullough, L.D., Zeng, Z., Li, H., Landree, L.E., McFadden, J., and Ronnett, G.V. (2005). Pharmacological inhibition of AMP-activated protein kinase provides neuroprotection in stroke. The Journal of biological chemistry 280, 20493-20502.10.1074/jbc.M409985200Search in Google Scholar PubMed

122 Reznick, R.M., Zong, H., Li, J., Morino, K., Moore, I.K., Yu, H.J., Liu, Z.X., Dong, J., Mustard, K.J., Hawley, S.A., et al. (2007). Aging-associated reductions in AMP-activated protein kinase activity and mitochondrial biogenesis. Cell metabolism 5, 151156.10.1016/j.cmet.2007.01.008Search in Google Scholar PubMed PubMed Central

123 Qiang, W., Weiqiang, K., Qing, Z., Pengju, Z., and Yi, L. (2007). Aging impairs insulin-stimulated glucose uptake in rat skeletal muscle via suppressing AMPKalpha. Experimental & molecular medicine 39, 535-543.10.1038/emm.2007.59Search in Google Scholar PubMed

124 Turdi, S., Fan, X., Li, J., Zhao, J., Huff, A.F., Du, M., and Ren, J. (2010). AMP-activated protein kinase deficiency exacerbates aging-induced myocardial contractile dysfunction. Aging Cell 9, 592-606.10.1111/j.1474-9726.2010.00586.xSearch in Google Scholar PubMed PubMed Central

125 Zhou G, M.R., Li Y, Chen Y, Shen X, Fenyk-Melody J, Wu M, Ventre J, Doebber T, Fujii N, Musi N, Hirshman MF, Goodyear LJ, Moller DE (2001). Role of AMP-activated protein kinase in mechanism of metformin action. The Journal of Clinical Investigation 108, 1167-1174.10.1172/JCI13505Search in Google Scholar PubMed PubMed Central

126 Owen MR, D.E., Halestrap AP (2000). Evidence that metformin exerts its anti-diabetic effects through inhibition of complex 1 of the mitochondrial respiratory chain. Biochemical Journal 348 Pt 3 607-614.10.1042/bj3480607Search in Google Scholar

127 Xiao B, S.M., Underwood E, Heath R, Mayer FV, Carmena D, Jing C, Walker PA, Eccleston JF, Haire LF, Saiu P, Howell SA, Aasland R, Martin SR, Carling D, Gamblin SJ (2011). Structure of mammalian AMPK and its regulation by ADP. Nature 472, 230-233. 210.1038/nature09932 Epub 02011 Mar 09913Search in Google Scholar

128 Koo SH, F.L., Qi L, Zhang X, Screaton RA, Jeffries S, Hedrick S, Xu W, Boussouar F, Brindle P, Takemori H, Montminy M (2005). The CREB coactivator TORC2 is a key regulator of fasting glucose metabolism. Nature 437, 1109-1111. Epub 2005 Sep 1107.10.1038/nature03967Search in Google Scholar PubMed

129 He L, S.A., Djedjos S, Miller R, Sun X, Hussain MA, Radovick S, Wondisford FE (2009). Metformin and insulin suppress hepatic gluconeogenesis through phosphorylation of CREB binding protein. Cell 137, 635-646. 610.1016/j.cell.2009.1003.1016Search in Google Scholar

130 Foretz M, H.S., Leclerc J, Zarrinpashneh E, Soty M, Mithieux G, Sakamoto K, Andreelli F, Viollet B (2010). Metformin inhibits hepatic gluconeogenesis in mice independently of the LKB1/AMPK pathway via a decrease in hepatic energy state. The Journal of Clinical Investigation 120, 2355-2369. 2310.1172/JCI40671 Epub 42010 Jun 40623Search in Google Scholar

131 Saisho, Y. (2015). Metformin and inflammation: its potential beyond glucose-lowering effect. Drug Targets 15, 196–205.10.2174/1871530315666150316124019Search in Google Scholar PubMed

132 Barzilai N, C.J., Kritchevsky SB, Espeland MA (2016). Metformin as a Tool to Target Aging. Cell Metabolism 23, 1060-1065. 1010.1016/j.cmet.2016.1005.1011Search in Google Scholar

133 Cuyàs E, V.S., Llorach-Pares L, Fernández-Arroyo S, Luciano-Mateo F, Cabré N, Stursa J, Werner L, Martin-Castillo B, Viollet B, Neuzil J, Joven J, Nonell-Canals A, Sanchez-Martinez M, Menendez JA (2018). Metformin directly targets the H3K27me3 demethylase KDM6A/UTX. Aging Cell, e12772. 12710.11111/acel.12772 Epub ahead of printSearch in Google Scholar

134 Finley, J. (2018). Cellular stress and AMPK activation as a common mechanism of action linking the effects of metformin and diverse compounds that alleviate accelerated aging defects in Hutchinson-Gilford progeria syndrome. Medical Hypotheses 118, 151-162. 110.1016/j.mehy.2018.1006.1029 Epub 2018 Jun 1028Search in Google Scholar

135 Kazi RS, B.R., Deshmukh AB, Patil GV, Jagadeeshaprasad MG, Kulkarni MJ (2017). Glycation inhibitors extend yeast chronological lifespan by reducing advanced glycation end products and by back regulation of proteins involved in mitochondrial respiration. Journal of Proteomics 156, 104-112. 110.1016/j.jprot.2017.1001.1015 Epub 2017 Jan 1027Search in Google Scholar

136 Cabreiro F, A.C., Leung KY, Vergara-Irigaray N, Cochemé HM, Noori T, Weinkove D, Schuster E, Greene ND, Gems D (2013). Metformin retards aging in C. elegans by altering microbial folate and methionine metabolism. Cell 153, 228-239. 210.1016/j.cell.2013.1002.1035Search in Google Scholar

137 Shin NR, L.J., Lee HY, Kim MS, Whon TW, Lee MS, Bae JW (2014). An increase in the Akkermansia spp. population induced by metformin treatment improves glucose homeostasis in diet-induced obese mice. Gut 63, 727-735. doi:710.1136/gutjnl-2012-303839.Epub 302013 Jun 303826.Search in Google Scholar

138 Na HJ, P.J., Jeon HJ, Park JS, Chung HY, Yoo MA (2018). Deficiency of Atg6 impairs beneficial effect of metformin on intestinal stem cell aging in Drosophila. Biochemical and Biophysical Research Communications 498, 18-24. 10.1016/j.bbrc.2018.1002.1191 Epub 2018 Feb 1027Search in Google Scholar

139 Vasamsetti SB, K.S., Kanugula AK, Thatipalli AR, Kumar JM, Kotamraju S (2015). Metformin inhibits monocyte-tomacrophage differentiation via AMPK-mediated inhibition of STAT3 activation: potential role in atherosclerosis. Diabetes 64, 2028-2041. 2010.2337/db2014-1225 Epub 2014 Dec 2031Search in Google Scholar

140 Cameron AR, M.V., Levin D, Mohan M, Forteath C, Beall C, McNeilly AD, Balfour DJ, Savinko T, Wong AK, Viollet B, Sakamoto K, Fagerholm SC, Foretz M, Lang CC, Rena G (2016). Anti-Inflammatory Effects of Metformin Irrespective of Diabetes Status. Circulation Research 119, 652-665. 610.1161/CIRCRESAHA.1116.308445 Epub 302016 Jul 308414Search in Google Scholar

141 Howell JJ, H.K., Turner M, Talbott G, Kolar MJ, Ross DS, Hoxhaj G, Saghatelian A, Shaw RJ, Manning BD (2017). Metformin Inhibits Hepatic mTORC1 Signaling via Dose-Dependent Mechanisms Involving AMPK and the TSC Complex. Cell Metabolism 25, 463-471. 410.1016/j.cmet.2016.1012.1009 Epub 2017 Jan 1012Search in Google Scholar

142 Dubbelhuis PF, M.A. (2002). Hepatic amino acid-dependent signaling is under the control of AMP-dependent protein kinase. 521, 39-42.10.1016/S0014-5793(02)02815-6Search in Google Scholar

143 Zhou W, K.A., Heijnen CJ (2016). Metformin Prevents Cisplatin-Induced Cognitive Impairment and Brain Damage in Mice. PLoS One 11, e0151890. 0151810.0151371/journal.pone.0151890.eCollection0152016Search in Google Scholar

144 Tanokashira D, K.E., Fukuokaya W, Kawabe K, Kashiwada M, Takeuchi H, Nakazato M, Taguchi A (2018). Metformin treatment ameliorates diabetes-associated decline in hippocampal neurogenesis and memory via phosphorylation of insulin receptor substrate 1. FEBS Open Bio 8, 1104-1118. 1110.1002/2211-5463.12436.eCollection12018JulSearch in Google Scholar

145 Thun MJ, J.E., Patrono C (2012). The role of aspirin in cancer prevention. Nature Reviews Clinical Oncology 9, 259-267. 210.1038/nrclinonc.2011.1199Search in Google Scholar

146 Hawley SA, F.M., Ross FA, Schertzer JD, Chevtzoff C, Walker KJ, Peggie MW, Zibrova D, Green KA, Mustard KJ, Kemp BE, Sakamoto K, Steinberg GR, Hardie DG (2012). The ancient drug salicylate directly activates AMP-activated protein kinase. Science 336, 918-922. 910.1126/science.1215327 Epub 1212012 Apr 1215319Search in Google Scholar

147 Castoldi F, P.F., Maiuri MC, Kroemer G (2018). Aspirin induces autophagy via inhibition of the acetyltransferase EP300. Oncotarget 9, 24574-24575. 24510.18632/oncotarget.25364 Epub 22018 May 24515Search in Google Scholar

148 Vane, J. (1971). Inhibition of prostaglandin synthesis as a mechanism of action for aspirin-like drugs. Nature: New biology 231, 232-235.10.1038/newbio231232a0Search in Google Scholar

149 Petrescu C, T.I. (1997). Uncoupling effects of diclofenac and aspirin in the perfused liver and isolated hepatic mitochondria of rat. Biochim Biophys Acta 1318, 385-394.10.1016/S0005-2728(96)00109-0Search in Google Scholar

150 Ai G, D.R., Kumar DR, Alfonso LF, Marimuthu S, Bhat GJ. (2016). Aspirin inhibits glucose6phosphate dehydrogenase activity in HCT 116 cells through acetylation: Identification of aspirin-acetylated sites. Molecular Medicine Reports 14, 1726-1732. 1710.3892/mmr.2016.5449 Epub 2016 Jun 1727Search in Google Scholar

151 Yuan M, K.N., Lee J, Hansen L, Li ZW, Karin M, Shoelson SE (2001). Reversal of obesity-and diet-induced insulin resistance with salicylates or targeted disruption of Ikkbeta. Science 293, 1673-1677.10.1126/science.1061620Search in Google Scholar PubMed

152 Madeo F, P.F., Eisenberg T, Kroemer G (2014). Caloric restriction mimetics: towards a molecular definition. Nature Reviews Drug Discovery 13, 727-740. 710.1038/nrd4391 Epub 2014 Sep 1012Search in Google Scholar

153 Pietrocola F, C.F., Maiuri MC, Kroemer G (2018). Aspirin-another caloric-restriction mimetic. Autophagy, 1-2. 10.1080/15548627.15542018.11454810 Epub ahead of printSearch in Google Scholar

154 Huang XB, M.X., Wan QL, He XM, Wu GS, Luo HR (2017). Aspirin increases metabolism through germline signalling to extend the lifespan of Caenorhabditis elegans. PLoS One 12, e0184027. 0184010.0181371/journal.pone.0184027.eCollection0182017Search in Google Scholar

155 Ayyadevara S, B.P., Dandapat A, Hu C, Khaidakov M, Mitra S, Shmookler Reis RJ, Mehta JL (2013). Aspirin inhibits oxidant stress, reduces age-associated functional declines, and extends lifespan of Caenorhabditis elegans. Antioxidant & Redox Signaling 18, 481-490. 410.1089/ars.2011.4151 Epub 2012 Sep 1087Search in Google Scholar

156 Tootle TL, S.A. (2008). Drosophila Pxt: a cyclooxygenase-like facilitator of follicle maturation. 135, 839-847. 810.1242/dev.017590 Epub 012008 Jan 017523Search in Google Scholar

157 Danilov A, S.M., Shevchenko O, Zemskaya N, Zhavoronkov A, Moskalev A (2015). Influence of non-steroidal anti-inflammatory drugs on Drosophila melanogaster longevity. Oncotarget 6, 19428-19444.10.18632/oncotarget.5118Search in Google Scholar PubMed PubMed Central

158 Strong R, M.R., Astle CM, Floyd RA, Flurkey K, Hensley KL, Javors MA, Leeuwenburgh C, Nelson JF, Ongini E, Nadon NL, Warner HR, Harrison DE (2008). Nordihydroguaiaretic acid and aspirin increase lifespan of genetically heterogeneous male mice. Aging Cell 7, 641-650. 610.1111/j.1474-9726.2008.00414.xSearch in Google Scholar

159 Bardia A, E.J., Vierkant RA, Limburg PJ, Anderson K, Wang AH, Olson JE, Vachon CM, Cerhan JR (2007). Association of aspirin and nonaspirin nonsteroidal anti-inflammatory drugs with cancer incidence and mortality.10.1093/jnci/djk200Search in Google Scholar PubMed

160 Mills EJ, W.P., Alberton M, Kanters S, Lanas A, Lester R (2012). Low-dose aspirin and cancer mortality: a meta-analysis of randomized trials. American Journal of Medicine 125, 560-567. 510.1016/j.amjmed.2012.1001.1017 Epub 2012 Apr 1017Search in Google Scholar

161 Madreiter-Sokolowski CT, S.A., Waldeck-Weiermair M, Malli R, Graier WF (2018). Targeting Mitochondria to Counteract Age-Related Cellular Dysfunction. Genes (Basel) 9, pii: E165. 110.3390/genes9030165Search in Google Scholar

162 Gazzerro, P., Proto, M.C., Gangemi, G., Malfitano, A.M., Ciaglia, E., Pisanti, S., Santoro, A., Laezza, C., and Bifulco, M. (2012). Pharmacological actions of statins: a critical appraisal in the management of cancer. Pharmacological reviews 64, 102-146.10.1124/pr.111.004994Search in Google Scholar PubMed

163 Basso, N., Paglia, N., Stella, I., de Cavanagh, E.M., Ferder, L., del Rosario Lores Arnaiz, M., and Inserra, F. (2005). Protective effect of the inhibition of the renin-angiotensin system on aging. Regulatory peptides 128, 247-252.10.1016/j.regpep.2004.12.027Search in Google Scholar PubMed

164 Inserra, F., Basso, N., Ferder, M., Userpater, M., Stella, I., Paglia, N., Inserra, P., Tenembaum, D., and Ferder, L. (2009). Changes seen in the aging kidney and the effect of blocking the renin-angiotensin system. Therapeutic advances in cardiovascular disease 3, 341-346.10.1177/1753944709339195Search in Google Scholar PubMed

165 Linz, W., Heitsch, H., Scholkens, B.A., and Wiemer, G. (2000). Long-term angiotensin II type 1 receptor blockade with fonsartan doubles lifespan of hypertensive rats. Hypertension (Dallas, Tex. : 1979) 35, 908-913.10.1161/01.HYP.35.4.908Search in Google Scholar PubMed

166 Linz, W., Jessen, T., Becker, R.H., Scholkens, B.A., and Wiemer, G. (1997). Long-term ACE inhibition doubles lifespan of hypertensive rats. Circulation 96, 3164-3172.10.1161/01.CIR.96.9.3164Search in Google Scholar

167 Huang, C.C., Chan, W.L., Chen, Y.C., Chen, T.J., Lin, S.J., Chen, J.W., and Leu, H.B. (2011). Angiotensin II receptor blockers and risk of cancer in patients with systemic hypertension. The American journal of cardiology 107, 1028-1033.10.1016/j.amjcard.2010.11.026Search in Google Scholar PubMed

168 Slack, C., Alic, N., Foley, A., Cabecinha, M., Hoddinott, M.P., and Partridge, L. (2015). The Ras-Erk-ETS-Signaling Pathway Is a Drug Target for Longevity. Cell 162, 72-83.10.1016/j.cell.2015.06.023Search in Google Scholar PubMed PubMed Central

169 Filer, D., Thompson, M.A., Takhaveev, V., Dobson, A.J., Kotronaki, I., Green, J.W.M., Heinemann, M., Tullet, J.M.A., and Alic, N. (2017). RNA polymerase III limits longevity downstream of TORC1. Nature 552, 263-267.10.1038/nature25007Search in Google Scholar PubMed PubMed Central

170 Chiang, W.C., Wei, Y., Kuo, Y.C., Wei, S., Zhou, A., Zou, Z., Yehl, J., Ranaghan, M.J., Skepner, A., Bittker, J.A., et al. (2018). High-Throughput Screens To Identify Autophagy Inducers That Function by Disrupting Beclin 1/Bcl-2 Binding. ACS chemical biology 13, 2247-2260.10.1021/acschembio.8b00421Search in Google Scholar PubMed PubMed Central

171 Li, Y., McGreal, S., Zhao, J., Huang, R., Zhou, Y., Zhong, H., Xia, M., and Ding, W.X. (2016). A cell-based quantitative high-throughput image screening identified novel autophagy modulators. Pharmacological research 110, 35-49.10.1016/j.phrs.2016.05.004Search in Google Scholar PubMed PubMed Central

172 Catana, C.S., Atanasov, A.G., and Berindan-Neagoe, I. (2018). Natural products with anti-aging potential: Affected targets and molecular mechanisms. Biotechnology advances 36, 1649-1656.10.1016/j.biotechadv.2018.03.012Search in Google Scholar PubMed

173 Pasyukova, E.G., and Vaiserman, A.M. (2017). HDAC inhibitors: A new promising drug class in anti-aging research. Mechanisms of ageing and development 166, 6-15.10.1016/j.mad.2017.08.008Search in Google Scholar PubMed

174 Chueh, A.C., Tse, J.W., Togel, L., and Mariadason, J.M. (2015). Mechanisms of Histone Deacetylase Inhibitor-Regulated Gene Expression in Cancer Cells. Antioxidants & redox signaling 23, 66-84.10.1089/ars.2014.5863Search in Google Scholar PubMed PubMed Central

175 Mottamal, M., Zheng, S., Huang, T.L., and Wang, G. (2015). Histone deacetylase inhibitors in clinical studies as templates for new anticancer agents. Molecules (Basel, Switzerland) 20, 3898-3941.10.3390/molecules20033898Search in Google Scholar PubMed PubMed Central

176 Suresh, P.S., Devaraj, V.C., Srinivas, N.R., and Mullangi, R. (2017). Review of bioanalytical assays for the quantitation of various HDAC inhibitors such as vorinostat, belinostat, panobinostat, romidepsin and chidamine. Biomedical chromatography : BMC 31.10.1002/bmc.3807Search in Google Scholar PubMed

177 de Magalhaes, J.P., and Passos, J.F. (2018). Stress, cell senescence and organismal ageing. Mechanisms of ageing and development 170, 2-9.10.1016/j.mad.2017.07.001Search in Google Scholar PubMed

178 Bernardes de Jesus, B., Schneeberger, K., Vera, E., Tejera, A., Harley, C.B., and Blasco, M.A. (2011). The telomerase activator TA-65 elongates short telomeres and increases health span of adult/old mice without increasing cancer incidence. Aging Cell 10, 604-621.10.1111/j.1474-9726.2011.00700.xSearch in Google Scholar PubMed PubMed Central

179 Harley, C.B., Liu, W., Blasco, M., Vera, E., Andrews, W.H., Briggs, L.A., and Raffaele, J.M. (2011). A natural product telomerase activator as part of a health maintenance program. Rejuvenation research 14, 45-56.10.1089/rej.2010.1085Search in Google Scholar PubMed PubMed Central

180 Zhu, Y., Tchkonia, T., Pirtskhalava, T., Gower, A.C., Ding, H., Giorgadze, N., Palmer, A.K., Ikeno, Y., Hubbard, G.B., Lenburg, M., et al. (2015). The Achilles’ heel of senescent cells: from transcriptome to senolytic drugs. Aging Cell 14, 644-658.10.1111/acel.12344Search in Google Scholar PubMed PubMed Central

181 Chang, J., Wang, Y., Shao, L., Laberge, R.M., Demaria, M., Campisi, J., Janakiraman, K., Sharpless, N.E., Ding, S., Feng, W., et al. (2016). Clearance of senescent cells by ABT263 rejuvenates aged hematopoietic stem cells in mice. Nature medicine 22, 78-83.10.1038/nm.4010Search in Google Scholar PubMed PubMed Central

182 Myrianthopoulos, V. (2018). The emerging field of senotherapeutic drugs. Future medicinal chemistry 10, 2369-2372.10.4155/fmc-2018-0234Search in Google Scholar PubMed

183 Wang, Y., Chang, J., Liu, X., Zhang, X., Zhang, S., Zhou, D., and Zheng, G. (2016). Discovery of piperlongumine as a potential novel lead for the development of senolytic agents. Aging 8, 2915-2926.10.18632/aging.101100Search in Google Scholar PubMed PubMed Central

184 Zhu, Y., Doornebal, E.J., Pirtskhalava, T., Giorgadze, N., Wentworth, M., Fuhrmann-Stroissnigg, H., Niedernhofer, L.J., Robbins, P.D., Tchkonia, T., and Kirkland, J.L. (2017). New agents that target senescent cells: the flavone, fisetin, and the BCL-XL inhibitors, A1331852 and A1155463. Aging 9, 955-963.10.18632/aging.101202Search in Google Scholar PubMed PubMed Central

185 Baar, M.P., Brandt, R.M.C., Putavet, D.A., Klein, J.D.D., Derks, K.W.J., Bourgeois, B.R.M., Stryeck, S., Rijksen, Y., van Willigenburg, H., Feijtel, D.A., et al. (2017). Targeted Apoptosis of Senescent Cells Restores Tissue Homeostasis in Response to Chemotoxicity and Aging. Cell 169, 132-147 e116.10.1016/j.cell.2017.02.031Search in Google Scholar PubMed PubMed Central

186 Birch, J., and Passos, J.F. (2017). Targeting the SASP to combat ageing: Mitochondria as possible intracellular allies? BioEssays : news and reviews in molecular, cellular and developmental biology 39.10.1002/bies.201600235Search in Google Scholar PubMed

Received: 2018-10-15
Accepted: 2018-12-04
Published Online: 2018-12-31

© 2018 O. Hillson et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 License.

Downloaded on 25.4.2024 from https://www.degruyter.com/document/doi/10.1515/bmc-2018-0018/html
Scroll to top button