Next Article in Journal
Pre-Treatment Hemoglobin Concentration and Absolute Monocyte Count as Independent Prognostic Factors for Survival in Localized or Locally Advanced Prostate Cancer Patients Undergoing Radiotherapy
Next Article in Special Issue
Lactobacillus Strains for Vegetable Juice Fermentation—Quality and Health Aspects
Previous Article in Journal
A Candidate Therapeutic Monoclonal Antibody Inhibits Both HRSV and HMPV Replication in Mice
Previous Article in Special Issue
Mineral-Enriched Postbiotics: A New Perspective for Microbial Therapy to Prevent and Treat Gut Dysbiosis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Therapeutic Antiaging Strategies

by
Shailendra Kumar Mishra
1,
Vyshnavy Balendra
2,
Josephine Esposto
3,
Ahmad A. Obaid
4,
Ricardo B. Maccioni
5,
Niraj Kumar Jha
6,7,8,
George Perry
9,
Mahmoud Moustafa
10,11,
Mohammed Al-Shehri
10,
Mahendra P. Singh
12,
Anmar Anwar Khan
13,
Emanuel Vamanu
14,* and
Sandeep Kumar Singh
1,*
1
Indian Scientific Education and Technology Foundation, Lucknow 226002, India
2
Saint James School of Medicine, Park Ridge, IL 60068, USA
3
Department of Environmental and Life Sciences, Trent University, Peterborough, ON K9L 0G2, Canada
4
Laboratory Medicine Department, Faculty of Applied Medical Sciences, Umm Al-Qura University, Makkah 21955, Saudi Arabia
5
Laboratory of Neurosciences and Functional Medicine, International Center for Biomedicine, University of Chile, Santiago 7750000, Chile
6
Department of Biotechnology, School of Engineering & Technology (SET), Sharda University, Greater Noida 201310, India
7
Department of Biotechnology, School of Applied & Life Sciences, Uttranchal University, Dehradun 248007, India
8
Department of Biotechnology, Engineering and Food Technology, Chandigarh University, Mohali 140413, India
9
Department of Biology, The University of Texas at San Antonio (UTSA), San Antonio, TX 78249, USA
10
Department of Biology, College of Science, King Khalid University, Abha 9004, Saudi Arabia
11
Department of Botany and Microbiology, Faculty of Science, South Valley University, Qena 83523, Egypt
12
School of Bioengineering and Biosciences, Lovely Professional University, Phagwara 144411, India
13
Laboratory Medicine Department, College of Applied Medical Sciences, Umm Al-Qura University, Makkah 78249, Saudi Arabia
14
Faculty of Biotechnology, University of Agricultural Sciences and Veterinary Medicine, 011464 Bucharest, Romania
*
Authors to whom correspondence should be addressed.
Biomedicines 2022, 10(10), 2515; https://doi.org/10.3390/biomedicines10102515
Submission received: 13 August 2022 / Revised: 21 September 2022 / Accepted: 24 September 2022 / Published: 8 October 2022

Abstract

:
Aging constitutes progressive physiological changes in an organism. These changes alter the normal biological functions, such as the ability to manage metabolic stress, and eventually lead to cellular senescence. The process itself is characterized by nine hallmarks: genomic instability, telomere attrition, epigenetic alterations, loss of proteostasis, deregulated nutrient sensing, mitochondrial dysfunction, cellular senescence, stem cell exhaustion, and altered intercellular communication. These hallmarks are risk factors for pathologies, such as cardiovascular diseases, neurodegenerative diseases, and cancer. Emerging evidence has been focused on examining the genetic pathways and biological processes in organisms surrounding these nine hallmarks. From here, the therapeutic approaches can be addressed in hopes of slowing the progression of aging. In this review, data have been collected on the hallmarks and their relative contributions to aging and supplemented with in vitro and in vivo antiaging research experiments. It is the intention of this article to highlight the most important antiaging strategies that researchers have proposed, including preventive measures, systemic therapeutic agents, and invasive procedures, that will promote healthy aging and increase human life expectancy with decreased side effects.

1. Introduction

Aging is depicted by a gradual functional decline causing the progressive deterioration that occurs heterogeneously across multiple organs, leading to organ dysfunction in mammals. It is a major risk factor for common chronic diseases, such as heart disease, cancer, diabetes, and Alzheimer’s diseases [1]. In recent years, aging research has been geared toward advancing upon knowledge established on the nine hallmarks of aging at the cellular and molecular levels [2]. Several hallmarks of aging are as follows: (i) genomic instability. high frequency of mutations within the genome of the cell; (ii) telomere attrition, i.e., protective caps going to end with each subsequent cell division; (iii) epigenetic alterations, i.e., DNA methylation, histone modification, and chromatin remodeling that change in gene expression; (iv) loss of proteostasis, i.e., folding, chaperoning, and maintenance of protein function collapses; (v) deregulation of nutrient sensing, i.e., reduction in the function of IGF-1, mTOR, sitrulins, and AMPK that regulate the metabolism; (vi) mitochondrial dysfunction, i.e., reduced efficiency of oxidative phosphorylation and reduction in the production of ATP; (vii) cellular senescence, i.e., ended the activation of p53 and the cyclin-dependent kinase (CDK) inhibitor p16 due to a variety of mechanisms, such as telomere shortening, forms of genotoxic stress, mitogens, and inflammatory cytokines; (viii) stem cell exhaustion, i.e., reduction in stem cell activity and altered intercellular communication, and (ix) the burden of senescent cells (Figure 1).
Recently, scientific achievements have been focused on producing effective antiaging therapeutics that have dramatically improved human life expectancy. Many studies on animal models looking at genetics and dietary and pharmacological interventions have shown an enhanced lifespan [3,4,5]. Other studies have examined antiaging strategies, such as enhancement of autophagy, elimination of senescent cells, transfusion of young blood, intermittent fasting, stem cell therapy, physical exercise, adult neurogenesis boost, and antioxidant and herbal intakes, which have emerged in recent years [6]. In humans, evidence has been accumulated on the styles of life in increasing the lifespan [7].
In this review, we have attempted to collect various in vitro and in vivo research experiments for antiaging approaches based on the major hallmarks of aging. Here, we describe antiaging strategies such as telomere reactivation, epigenetic drugs, activation of chaperons and proteolytic systems, and dietary restriction. Mitophagy and the clearance of senescent cells, etc. will be beneficial in understanding the unique approaches for successful aging and to extend healthy lifespans.

2. Anti-Inflammatory Drugs Used as an Antiaging Approach

Chronic inflammation is one of the major contributors to age-associated diseases and aging and disrupts the normal functioning of tissues [1,8,9]. Increased activity of proinflammatory pathways accompanies inflammation with age [10,11]. Serum concentrations of proinflammatory cytokines (IL-1, IL-2, IL-6, IL-8, IL-12, IL-15, IL-17, IL-18, IL-22, IL-23, tumor necrosis factor-α, and interferon-γ) are significantly increased in normal process aging compared with younger individuals in a normal stage [12,13,14]. A chronic proinflammatory status is a pervasive feature of aging, and increased systemic inflammation is closely associated with aging and age-related diseases [15,16]. The term “inflammaging” is used to describe aging induced by chronic and persistent inflammation. Anti-inflammatory agents block certain substances in the body that cause inflammation, and various studies have shown that anti-inflammatory agents are linked to antiaging [17]. The most important drivers of age-dependent inflammation are derived at the cellular and molecular levels. In a cell, the proinflammatory senescence-associated secretory phenotype (SASP) is associated with cellular senescence that is triggered by agents such as radiation and viruses and by continuous exposure to cellular debris [18] and cellular senescence [19]. In a molecule, ROS (reactive oxygen species) and other agents can trigger inflammatory DNA damage responses that affect DNA and telomeres [20] and activate the inflammasomes and NF-kB pathway [21]. Inhibition of the inflammatory processes by genetic and pharmacological intervention is considered an effective and verified antiaging strategy [2]. Nonsteroidal anti-inflammatory drugs (NSAIDs) not only prevent certain age-associated features but also increase the lifespan in various model organisms, such as yeasts [22], nematodes [23], mice [24,25], and flies [22]; however, their effectiveness for neurodegenerative disorders (Alzheimer’s disease and Huntington’s disease [26]) is not clear, and there is a search for anti-inflammatory bioactive compounds [27,28,29,30,31]. Anti-inflammatory drugs might be considered to have great potential for extending the lifespan. In some studies, spermidine (polyamines) and their action on the expression of pro- and anti-inflammatory cytokines can directly reduce inflammation and indirectly alter inflammation and cell growth by the action of autophagy [32]. Spermidine has been reported to slow down aging due to its antiaging effects [33]. Aspirin, a potent anti-inflammatory and antioxidant compound, may affect oxidant production and cytokine responses and block glycoxidation reactions that protect against oxidative stress, as well as extend the lifespan of Caenorhabditis elegans and mice [34,35,36]. Ibuprofen (NSAID) has been shown to reduce the risk of age-related pathologies and increase the lifespan of Saccharomyces cerevisiae, Caenorhabditis elegans, and Drosophila melanogaster [37,38]. A novel NSAID, M2000, could modify oxidative stress pathways by lowering the expression levels of the SOD2, GST, iNOS, and MPO genes and reduce the risk of inflammatory diseases through its immunosuppressive effects, with no adverse side effects on the enzymatic and nonenzymatic determinants [39]. This can be recommended as an antiaging drug. MAAs (mycosporine-like amino acids), such as M2G (mycosporine-2-glycine), exhibit antioxidant, anti-inflammatory, anti-protein-glycation, and collagenase inhibition activities and show the ability to protect DNA against UV damage [40,41]. Many nutraceuticals (apigenin, quercetin, kaempferol, naringenin, catechins, epigallocatechin, genistein, cyanidin, resveratrol, etc.) and functional foods possess antioxidant activity that might play an important role in delaying aging and be effective in various human neurodegenerative diseases [30,42,43,44,45].

3. Antioxidant Activity

Phytochemicals such as phenolic acids and flavonoids have antioxidant activity, which acts by scavenging free radicals and increasing the levels of antioxidant enzymes in plasma [46]. The function of a primary antioxidant enzyme is to protect organisms from the damaging effects of superoxide radicals, which are quickened by their dismutation into hydrogen peroxide and oxygen [47]. Several studies have confirmed that quercetin is a strong antioxidant that accumulates in nematodes and displays reactive oxygen species (ROS) scavenging activity and has been demonstrated to have a positive effect on longevity and stress resistance in various animal models [48,49,50,51]. Many studies have demonstrated that NSAIDs have antioxidant activity that is mediated by free radical scavenging and antioxidant enzyme activation [52]. The antioxidant activity of NSAIDs has been witnessed in membranes, cells, and at the organismal level [52,53,54].

4. Telomere Reactivation

Telomeres are conserved microsatellite repeats TTAGGG that protect the ends of chromosomes from DNA breakage and prevent DNA end-joining, recombination, and DNA repair [55]. DNA polymerases are incapable of fully replicating the linear chromosomes owing to end replication in somatic cells, and telomeres become gradually shortened after each cell division [56]. This shortening of telomeres is usually fulfilled by the telomerase enzyme, but most somatic cells and adult stem cells do not express enough telomerase to compensate for the telomere length that leads to entering ‘replicative senescence’, which might be followed by cell death [57,58,59]. Telomere shortening occurs during normal aging and is an important biomarker of aging and longevity that is influenced by several factors, such as genetics, epigenetics, and environments [60,61,62,63]. It is also associated with many age-related diseases, such as osteoarthritis, atherosclerosis, coronary heart disease, and atrial fibrillation [64,65,66]. Several studies have reported that aging can be inhibited by the overexpression of telomerase; however, it can enhance tumorigenesis [67,68,69]. A telomerase activator, telomerase expression activator, and telomerase gene therapy have been developed as telomerase-based antiaging strategies in recent years. TA-65 is an extract of a Chinese plant (Astragalus membranaceus), a telomerase activator that can restore telomere length without cancer occurrence and improve age-related indicators, including glucose tolerance, bone health, and skin quality [70]. Additionally, some studies found that TERT transcription activator and sex hormones are directly involved in activating telomerase, which rescued telomere shortening and enhanced the lifespan [71,72]. Evidence has suggested that the reactivation of telomerase expression by using a gene therapy approach is the best example of the lifespan extension of mice and delay aging without cancer occurrence [73]. A recent study has found that Metadichol, a telomerase activator, is used to overcome organ failure by enriching cells with telomerase and is a safer alternative [74]. Another study explained that natural compounds such as 08AGTLF (Centella asiatica), Nutrient 4 (Astragalus), TA-65 (Astragalus membranaceus), OA (oleanolic acid), and MA (maslinic acid); and Nutrients 1, 2, and 3 have telomerase activation, and among all, 08AGTLF has the greatest potential to activate telomerase [75]. The impact of telomeric length on humans has been evidenced by the fact that the expression of telomerase in normal cells may extend a healthy lifespan; however, inhibition of telomerase in cancer cells may be a viable target for anticancer therapeutics [76].

5. Antiaging Approaches Using Epigenetic Drugs

The effects of chromatin on aging are probably complex and bidirectional. Chromatin remodeling appears to counter aging and age-associated diseases and extend organismal lifespan [77]. Chromatin is intensely altered during aging owing to the decreased level of histone proteins and adequate changes in histone modification that were found in recent studies in budding yeast and human fibroblast cells [78]. Elevating histone expression, reducing H4 K16 acetylation, reducing H3 N-terminal acetylation, inactivating the HDAC Rpd3, and inactivating the H3K4 methylase might be capable of extending lifespan or reverting the aged phenotype to a more youthful state of chromatin [79]. These epigenetic factors are influenced by diet, lifestyle and exogenous stress, which raises the possibility of enhancing age-related cellular dysfunction [80,81]. Several studies revealed age-associated changes in DNA methylation patterns leading to a global reduction in DNA methylation [82] (Figure 2); site-specific hypermethylation, specifically at CpG islands and polycomb target sites [83]; site-specific hypomethylation specifically at gene-poor regions, tissue-specific promoters, and polycomb protein regions; and hypermethylation in different tissues but hypomethylation, to be more tissue specific [84,85]. These changes accumulate gradually; such changes are indicative of the aging process and strongly associated with epigenetic changes such as replicative senescence [86]. Diet, lifestyle, environmental interventions, and inhibitors of epigenetic enzymes have proven to be effective in promoting longevity, as seen in various experiments [87]. Natural substances such as spermidine and resveratrol have been found to lead to deacetylation of chromatin, indicating that it has the potential to extend the lifespan in humans [88]. Many studies report that HDAC (histone deacetylase) inhibitors show evidence as an antiaging strategy [89]. HDAC inhibitors have the potential to reverse the aging process that allows healthy aging [90].

6. Activation of Chaperons and the Proteolytic System against Aging

Loss of proteostasis is a common feature of aging that leads to protein aggregation, unfolding, oxidative damage, posttranslational modification, and an altered rate of protein turnover and, ultimately, to cellular dysfunction [91,92,93,94,95]. Two proteolytic systems—the ubiquitin–proteasome system (UPS) and autophagy–lysosome system (ALS)—and chaperones play a major role in maintaining proteostasis [96]. The alteration or deterioration of these pathways impairs normal cell functioning and cell physiology, causing aging [2]. Many studies have found proteostasis changes with age owing to reduced activity of heat-shocked protein chaperones [97,98]. To compensate for this decline, increasing the chaperone protein level has been shown to beneficially impact longevity in worms and flies [99,100,101,102]. A study reported that the aggregation of Hsp104, a chaperone, has been associated with aging and has increased the lifespan in Saccharomyces cerevisiae [103,104,105]. The UPS and ALS systems are the main proteolytic systems that influence the cellular fate and aging process [106,107]. The availability of the chaperone is extremely compromised in aged cells in which the proteostasis collapses by decreased G1-cyclin function that causes an irreversible arrest in G1, configuring a molecular pathway claiming proteostasis deterioration leads to cell senescence [108]. Promoting proteasomal activity via overexpression of the proteasomal β5 subunit either in Caenorhabditis elegans [109] or in human fibroblast [110] and the overexpression of Rpn11 in Drosophila melanogaster [111] increases the lifespan and stress resistance. The compound 18α-glycyrrhetinic acid and loss of IGF (insulin-like growth factor) signaling due to mutated daf-2 induce proteasomal activation and extend the lifespan of Caenorhabditis elegans [112]. Spermidine, metformin, rapamycin, and resveratrol are pharmaceutical approaches well-known to activate the autophagy system [113]. In a recent study, it was found that minocycline, JZL184, monorden, and paxilline directly targeted the 18S rRNA/ribosome, FAAH-4, Hsp90, and the SLO-1 BK channel, significantly increasing the lifespan of Caenorhabditis elegans [114]. The proteostatic system governs the synthesis and conformation of target proteins, and the ubiquitin—proteasome system and autophagy act as the main scavengers of misfolded or excessive proteins. The main cause of Alzheimer’s disease is the accumulation of misfolded proteins as Aβ plaques and tau aggregates owing to dysregulation of proteostasis, which contributes to the accumulation of proteotoxins in Alzheimer’s disease [115]. It has been reported that a decrease in the efficiency of the autophagy and ubiquitin—proteasome systems might lead to aging and neurodegenerative diseases such as AD, PD, and ALS [116]. The prion diseases in mammals are related to altered versions of PrPc (cellular), which is a key component of the infectious agent responsible for transmission, and the disease-associated version of PrPc can be partially resistant to the protease–digestion system, designated PrPsc (scrapie) [117]. Various cellular components, predominantly chaperones such as Hsp104, Hsp40s, HSP42, and HSP70s, can lead to the curing of yeast prions by their deficiency or overproduction. These studies have revealed the requirements for prion propagation and conditions that affect prion stability, which have led to the discovery of anti-prion systems. Btn2p, a component of the yeast anti-prion system, has the aggregate-sequestering abilities; it can cure an artificial and natural yeast prion, and works on a variety of non-prion aggregates as well [118]. This system might be advantageous for neurodegenerative diseases that result from aggregates of proteo-toxins.

7. Mitophagy Activators as an Antiaging Approach

Mitochondrial dysfunction is one of the hallmarks of aging [2] that is associated with aging and is also involved in the development of many neurodegenerative diseases [119]. Mitochondrial dysfunction is caused by a defect in mitophagy that has been associated with aging and other diseases such as metabolic disorders, cancer, senescence, inflammation, and genomic instability [120]. Mitochondrial health plays an important role in the aging process that can assist in the therapeutic approach toward longevity [121]. Mitophagy could turn to a pro-death pathway due to excessive superoxide stress that leads to accelerated aging [122]. Parkin is a key protein in mitophagy [123] and is overexpressed, leading to a decreased lifespan in Drosophila [124,125]. Mitochondrial health depends on inflammation, because many inflammatory pathologies have been associated with mitochondrial defects [126]. Studies report that many natural products have the potential to act as antiaging strategies [127]. Many inducers, such as sirtuin-activating compounds (STACs), NADþ precursors (NMN and NR), and resveratrol, have been shown to modulate mitophagy and repair mitochondrial functions [128,129,130] (Figure 3).
Resveratrol (3,5,4-trihydroxystilbene) is a natural polyphenol that can induce autophagy and mitophagy by different mechanisms, such as activation of AMPK and SIRT1; induction of P38 MAPK; suppression of mTOR and p70S6K; and upregulation of eNOS, GABARAP, LC3B, and ATG3 genes, which is a beneficial compound for extending the lifespan [131,132,133,134,135,136,137,138].
Some natural compounds, such as spermidine and urolithin, preserve mitochondrial function and extend longevity through mitophagy induction, which has been demonstrated in several model organisms, such as yeast, nematodes, flies, and mice [113,139,140,141]. Moreover, tomatidine is a natural compound that stimulates the elimination of defective mitochondria, leading to increased longevity and enhanced muscular function in nematodes and mice [142]. Some studies reveal that antibiotics severely affect mitochondrial homeostasis [143]. Actinomycin and doxycycline interrupt energy metabolism and facilitate mitophagy in mammalian cells [144,145]. The mitochondria generates most cellular ROS, but mtDNA is comparatively unprotected from ROS damage due to a lack of histone proteins and a lack of robust mtDNA repair mechanisms. A decline in mitochondrial energy metabolism and accumulation of mtDNA mutations in tissue cells are important contributors to human aging, and some type of mtDNA mutations are leading to the development of cancer [146]. A study showed that deficiency and/or dysfunction in many of the nuclear-encoded mtDNA replicative proteins, such as Pol γ, PolG2, Twinkle, TFAM, MGME1, and RNase H1, are responsible for the development of age-related diseases and/or phenotypes both in vitro and in vivo [147].

8. Inhibition of mTOR and Insulin/IGF-1 Signaling (IIS) as an Antiaging Approach

mTOR promotes cell growth by either promoting protein synthesis or inhibiting autophagy activity. It is involved in senescence-associated phenotypes. The mTOR pathway acts as a nutrient sensor through regulation of mitochondrial biogenesis [148], regulation of mitophagy [149], promotion of the secretory phenotype of senescent cells [150], and inhibition of stem cell senescence [151,152]. Many studies have demonstrated that a reduction of mTOR signaling is associated with extension of the lifespan in Caenorhabditis elegans [153], Drosophila melanogaster [154,155], Saccharomyces cerevisiae [156,157], and Mus musculus [158]. Thus, it is the most observed target for the study of pharmacological treatments. Pharmacological inhibition of mTOR by rapamycin confirms that it can be observed to slow down aging in yeast (S. cerevisiae) [159], Drosophila melanogaster [154,160,161,162,163,164], worm (C. elegans) [165], and mice (M. musculus) [166,167,168,169,170,171,172,173,174,175,176,177,178,179,180,181,182,183,184,185,186,187]. It has also been seen to delay age-related diseases in different species including humans [188,189,190,191,192,193,194,195,196,197,198,199,200]. mTORC1 has two substrates, S6 kinase and 4E-BP1, and both are linked to longevity through the reduction of S6 kinase, which promotes longevity [201]. Another nutrient-sensing pathway is the insulin/insulin-like growth factor 1 (IGF-1) signaling (IIS) network, which is a major determinant of longevity important to regulate the lifespan in various species, such as C. elegans [202], D. melanogaster [203,204], and M. musculus [205,206]. It has been shown that suppression of the IIS/mTOR pathways is involved in prolonging the lifespan but also delays the onset of age-related pathologies in numerous organisms, such as yeast, fruit flies, nematodes, mice, rats, and primates [207,208,209]. In recent years, studies have reported that celecoxib (a nonsteroidal anti-inflammatory drug), recombinant BTI (rBTI-mimicking restriction), and the ethyl acetate fraction of Ribes fasciculatum (ERF) downregulate the insulin/IGF-1 signaling cascade and extend the lifespan in Caenorhabditis elegans [23,210]. Recently, a study demonstrated that hypo-taurine, an antioxidant, induces the lifespan in Caenorhabditis elegans by regulating the insulin/IGF-1 signaling (IIS) pathway [211,212].

9. Activation of AMPK and Sirtuin Signaling as an Antiaging Approach

5′adenosine monophosphate-activated protein kinase (AMPK) is a master regulator of energy metabolism in the cell that plays a critical role in regulating health and longevity [213]. AMPK is a serine/threonine protein kinase that is a highly conserved sensor that increases the levels of AMP and ADP [214,215]. Sirtuins (SIRT1 to SIRT7), also known as histone deacetylases (HDACs), delay cellular senescence and promote longevity in various species through increased expression [216,217,218,219,220,221]. Many FDA-approved drugs, such as biguanides, thiazolidinediones, glucagon-like peptide-1 receptor agonists, salicylates and resveratrol, and 5-aminoimidazole-4-carboxamide riboside (AICAR), have AMPK-activating properties [222,223]. Resveratrol (3,5,40-trihydroxystilbene) is a natural activator that increases the production of synthetic SIRT1 activators, which are more potent, soluble and bioavailable [136,224]. Some synthetic activators, such as SRT1720 (imidazothiazoles), thiazolopyridine, benzimidazole, bridged ureas, cilostazol, paeonol, statins, hydrogen sulfide, persimmon, and SRT2104 also extend the lifespan in mice and protect cells against age-related changes [225,226,227,228,229,230,231,232]. Moreover, natural compounds such as quercetin, proanthocyanidins, fisetin, catechins, kaempferol, and butein have been reported to have antiaging properties [233]. Metformin activates AAK-2/AMPK and induces autophagy through AMPK, which is pro-longevity in nematodes, Drosophila, rats, and mice [234,235,236,237,238,239,240,241]. Oligonol is an antioxidant polyphenolic compound with an anti-inflammatory property that activates the autophagy pathway and phosphorylation of AMPK in C. elegans [242].

10. Clearance of Senescent Cells

Cellular senescence is a central component of aging that leads to cell cycle arrest in a damaged cell by preventing the cell from promulgating further damaged tissues [243,244,245]. Senescent cells accumulate with age and promote aging and age-associated pathologies [246,247,248,249]. To extend the lifespan, senescent cells need to be decreased or suppressed through the senolytics approach. There are many senolytics that have been reported to induce apoptosis of senescent cells and reduce the expression of senescence markers, including a combination of dasatinib and quercetin, BCL2 family inhibitors, SCAPs, FOXO4, and piperlongumine [250,251,252,253,254,255,256,257]. Moreover, HSP90 has been identified as a novel senolytics that can induce apoptosis of senescent cells to improve the health span [258]. There are several senomorphic drugs that suppress markers of senescence or their secretory phenotype, including inhibitors of IkB kinase (IKK) and nuclear factor (NF), inhibitors of free radical scavengers and the Janus kinase (JAK) pathway. Rapamycin acts as a senomorphic agent to reduce the SASP, and ABT263 inhibits the antiapoptotic proteins BCL-2 and BCL-XL [259,260,261]. In recent years, two well-known antibiotics, azithromycin and roxithromycin, had senolytic activity to target senescent cells [262]. Curcumin is an ayurvedic formulation that has antisenescence properties and can modulate cellular senescence by activating the sirtuins and AMPK pathways [263].

11. Stem Cell-Based Therapies

Stem cell exhaustion (decline in stem cell number and function) is one of the causes of aging that is observed in all tissues and organs and is maintained by adult stem cells through repair and regeneration during life [264,265]. There is a loss of adult stem cells such as hematopoietic stem cells (HPSCs), which respond to stress and differentiation, and HPSCs are decreased by the hyperactivation of the mechanistic target of rapamycin complex 1 and the disruption of 5′ adenosine monophosphate-activated protein kinase (AMPK) [266,267]. It has been reported that the accumulation of DNA damage and increased levels of p16INK4a have been closely linked to a decline in stem cell populations with aging [268]. Age-associated phenotypes could be restored by the induction of stem cell rejuvenation [269,270,271]. Reprogramming aged hematopoietic stem cells (HPSCs) to a pluripotent state and inducing pluripotent stem cells (iPSCs) from aged human fibroblasts into NSCs are alternate options [272,273]. Metformin is an FDA-approved drug that can induce stem cell rejuvenation in adult stem cell therapy [274]. Studies have proposed that using AMPK activators such as 5-aminoimidazole-4-carboxamide ribonucleotide, A769662, metformin, and oxidized nicotinamide adenine dinucleotide (NAD+) can be used in stem cell-based transplantation therapies with better results [275]. A study revealed that resveratrol-induced SIRT1 activation promotes the self-renewal of human embryonic stem cells (hESCs) [276]. A recent study reported that the administration of caloric restriction (CR) could enhance stem cell proliferation by regulating niche cells and reducing the major energy metabolic pathways that prevent stem cell aging [277].

12. Role of the Microbiome in Aging and Potential Antiaging Therapeutics

The evidence highlighted that microbiome composition may affect the rate of aging [278,279], although no evidence has been found that microbiota composition harshly changes the chronological threshold or age; rather, these changes proceed gradually with time [280]. Between the host and intestinal microbiota, the rate of age-related deterioration is strongly influenced by particular factors such as age-associated alterations in lifestyle, nutrition, frailty, and inflammation [281]. A study reported that the microbial composition is highly similar in young adults and seventy-year-old people but significantly differs in centenarians [280]. This difference was associated with the enrichment in opportunistic proinflammatory bacteria known as pathobionts, which are generally present in adult gut ecosystems in low numbers. The microbiota may favorably affect the host health and aging processes using prebiotics and probiotics that have been shown to be efficient in preventing particular pathological conditions, such as the suppression of inappropriate chronic inflammation in elderly populations. It includes a decrease in the synthesis of proinflammatory cytokines such as interleukins (IL-6, IL-8, IL-10, and TNF) and, thus, an increase in the levels of activated lymphocytes and natural killer cells and phagocytic activity [282]. Some other effects of microbiota, including the regulation of host fat deposition and metabolism [283], prevention of insulin resistance [284], degradation of nondigestible carbohydrates, enhancement of antioxidant activity, production of vitamin B and conjugated linoleic acids [285], and improved maintenance of mucosal barrier integrity and immune homeostasis [286]. Several research findings suggest that direct modulation of the gut microbiome might be beneficial to be applied in treating age-related disorders [287] and considered to be a novel potential therapeutic for healthy aging [288]. Some studies suggest that the intake of functional foods such as prebiotics, probiotics, or synbiotics might be an effective strategy to counter natural aging through modifying the gut microbiota of the elderly population [289].

13. Role of Noncoding RNAs in Aging and Their Potential Therapeutics

Long noncoding RNAs (lncRNAs) can affect key cellular processes, such as proliferation, differentiation, quiescence, senescence, the cellular response to stress and immune agents, and other cellular functions related to the biology of aging. LncRNAs have the ability to modulate gene expression patterns at the transcriptional, posttranscriptional, and posttranslational levels [290,291,292]. The changes in the subsets of expressed proteins are responsible for the aging traits. LncRNAs can modulate protein expression patterns by controlling gene transcription, mRNA stability, and protein abundance; thus, they can modulate key molecular events underlying the aging process, such as the control of telomere length, epigenetic gene expression, proteostasis, stem cell function, intercellular communication, cell proliferation, and cellular senescence [293].

14. Conclusions

This review has attempted to explain the various antiaging strategies used based on the major hallmarks of aging. Data have been collected from scientific evidence, such as studies conducted on organisms and species targeting aging. These approaches would not only postpone chronic diseases but also prevent many age-related disorders and extend healthy lifespans. Aging and age-related diseases are due to abnormalities in normal biological pathways, such as metabolism, inflammation, growth, and protein synthesis, which also alter the rate of aging. Interestingly, modifications in the normal epigenetic pathway appear to be a major cause of chronic disorders. This review presents various protective interventions as antiaging strategies that are valuable to extend the human healthy lifespan. These include telomere reactivation, epigenetic drugs, activation of chaperons and the proteolytic system, caloric restriction, mitophagy activation, clearance of senescent cells, antioxidant and anti-inflammatory drugs, inhibition of mTOR and insulin/IGF-1 signaling (IIS), and stem cell-based therapy. Targeting senescent cells and inflammation plays a prominent role in aging, such as employing anti-inflammatory bioactive molecules in recent studies or, in some cases, nonsteroidal anti-inflammatory drugs (NSAIDs), senolytics, and SASP suppressors. To achieve adequate telomere length through telomerase activators, HDAC (histone deacetylase) inhibitors may be considered as antiaging agents. To rescue mitochondrial function and promote proteasomal activity, SIRT1 and AMPK activators and mTOR suppressors may be an effective way to combat aging. Reprogramming technology such as iPSCs and stem cell therapy, are emerging pieces of evidence for supplementing stem cells to rejuvenate tissues and organ functions. Thus, exploring the aging process at the molecular level is still a challenge, but the possibility of developing novel therapeutics by using nutraceuticals, molecular medicine, and pharmacogenomics approaches may be options for successful aging (Figure 4).

Expert Opinion

Various scientific achievements have been focused on producing effective antiaging therapeutics that have dramatically improved human life expectancy. Many studies on animal models looking at genetics and dietary and pharmacological intervention have shown an enhanced lifespan. Thus, in this article, we have highlighted various antiaging strategies that have targeted approach-based hallmarks of aging. Researchers have proposed various strategies, such as preventive measurements, systemic therapeutic agents, and invasive procedures, that will promote healthy aging and increase human life expectancy with decreased side effects.

Author Contributions

Conceptualization, S.K.M., R.B.M., M.P.S., E.V., S.K.S., and M.A.-S.; writing—original draft preparation, S.K.M., V.B., J.E., G.P., E.V., and S.K.S.; and writing—review and editing, S.K.M., V.B., J.E., A.A.O., N.K.J., R.B.M., M.M., M.A.-S., A.A.K., and S.K.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was performed with resources from the Non-Government/Nonprofit Organization Indian Scientific Education and Technology Foundation, Lucknow, India, and Project ID19I-10301 from ANID, Chile. Authors from King Khalid University are highly acknowledge funding support from the Deanship of Scientific Research at King Khalid University, Saudi Arabia, Grant No. (R.G.P.. 2/192/43).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rojo, L.E.; Fernández, J.A.; Maccioni, A.A.; Jimenez, J.M.; Maccioni, R.B. Neuroin flammation: Implications for the Pathogenesis and Molecular Diagnosis of Alzheimer’s Disease. Arch. Med. Res. 2008, 39, 1–16. [Google Scholar] [CrossRef]
  2. López-Otín, C.; Blasco, M.A.; Partridge, L.; Serrano, M.; Kroemer, G. The Hallmarks of Aging. Cell 2013, 153, 1194–1217. [Google Scholar] [CrossRef] [Green Version]
  3. de Magalhães, J.P. Programmatic features of aging originating in development: Aging mechanisms beyond molecular damage? FASEB J. 2012, 26, 4821–4826. [Google Scholar] [CrossRef] [PubMed]
  4. Kenyon, C.J. The genetics of ageing. Nature 2010, 464, 504–512. [Google Scholar] [CrossRef]
  5. Tacutu, R.; Craig, T.; Budovsky, A.; Wuttke, D.; Lehmann, G.; Taranukha, D.; Costa, J.; Fraifeld, V.E.; de Magalhaes, J.P. Human Ageing Genomic Resources: Integrated Databases and Tools for the Biology and Genetics Of Ageing. Nucleic Acids Res. 2012, 41, D1027–D1033. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Shetty, A.K.; Kodali, M.; Upadhya, R.; Madhu, L.N. Emerging Anti-Aging Strategies—Scientific Basis and Efficacy. Aging Dis. 2018, 9, 1165–1184. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Helfand, S.L.; de Cabo, R. Evidence that overnight fasting could extend healthy lifespan. Nature 2021, 598, 265–266. [Google Scholar] [CrossRef]
  8. Franceschi, C.; Campisi, J. Chronic Inflammation (Inflammaging) and Its Potential Contribution to Age-Associated Diseases. J. Gerontol. Ser. A 2014, 69, S4–S9. [Google Scholar] [CrossRef] [PubMed]
  9. Howcroft, T.K.; Campisi, J.; Louis, G.B.; Smith, M.T.; Wise, B.; Wyss-Coray, T.; Augustine, A.D.; McElhaney, J.E.; Kohanski, R.; Sierra, F. The role of inflammation in age-related disease. Aging 2013, 5, 84–93. [Google Scholar] [CrossRef] [Green Version]
  10. De Magalhães, J.P.; Curado, J.; Church, G. Meta-analysis of age-related gene expression profiles identifies common signatures of aging. Bioinformatics 2009, 25, 875–881. [Google Scholar] [CrossRef] [PubMed]
  11. Kriete, A.; Mayo, K.L.; Yalamanchili, N.; Beggs, W.; Bender, P.; Kari, C.; Rodeck, U. Cell autonomous expression of inflammatory genes in biologically aged fibroblasts associated with elevated NF-kappaB activity. Immun. Ageing 2008, 5, 5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Minciullo, P.L.; Catalano, A.; Mandraffino, G.; Casciaro, M.; Crucitti, A.; Maltese, G.; Morabito, N.; Lasco, A.; Gangemi, S.; Basile, G. Inflammaging and Anti-Inflammaging: The Role of Cytokines in Extreme Longevity. Arch. Immunol. Ther. Exp. 2016, 64, 111–126. [Google Scholar] [CrossRef] [PubMed]
  13. Ventura, M.T.; Casciaro, M.; Gangemi, S.; Buquicchio, R. Immunosenescence in aging: Between immune cells depletion and cytokines up-regulation. Clin. Mol. Allergy 2017, 15, 21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Rea, I.M.; Gibson, D.S.; McGilligan, V.; McNerlan, S.E.; Alexander, H.D.; Ross, O.A. Age and Age-Related Diseases: Role of Inflammation Triggers and Cytokines. Front. Immunol. 2018, 9, 586. [Google Scholar] [CrossRef] [PubMed]
  15. Sanada, F.; Taniyama, Y.; Muratsu, J.; Otsu, R.; Shimizu, H.; Rakugi, H.; Morishita, R. Source of Chronic Inflammation in Aging. Front. Cardiovasc. Med. 2018, 5, 12. [Google Scholar] [CrossRef] [Green Version]
  16. Chung, H.Y.; Ki, W.C.; Lee, E.K.; Chung, K.W.; Chung, S.; Lee, B.; Seo, A.Y.; Chung, J.H.; Jung, Y.S.; Im, E.; et al. Redefining Chronic Inflammation in Aging and Age-Related Diseases: Proposal of the Senoinflammation Concept. Aging Dis. 2019, 10, 367–382. [Google Scholar] [CrossRef] [Green Version]
  17. Neves, J.M.; Sousa-Victor, P. Regulation of inflammation as an anti-aging intervention. FEBS J. 2020, 287, 43–52. [Google Scholar] [CrossRef] [Green Version]
  18. Coppé, J.-P.; Desprez, P.-Y.; Krtolica, A.; Campisi, J. The Senescence-Associated Secretory Phenotype: The Dark Side of Tumor Suppression. Annu. Rev. Pathol. Mech. Dis. 2010, 5, 99–118. [Google Scholar] [CrossRef] [Green Version]
  19. Jurk, D.; Wilson, C.; Passos, J.F.; Oakley, F.; Correia-Melo, C.; Greaves, L.; Saretzki, G.; Fox, C.; Lawless, C.; Anderson, R.; et al. Chronic inflammation induces telomere dysfunction and accelerates ageing in mice. Nat. Commun. 2014, 5, 4172. [Google Scholar] [CrossRef] [Green Version]
  20. Vitale, G.; Salvioli, S.; Franceschi, C. Oxidative stress and the ageing endocrine system. Nat. Rev. Endocrinol. 2013, 9, 228–240. [Google Scholar] [CrossRef]
  21. Youm, Y.-H.; Grant, R.W.; McCabe, L.R.; Albarado, D.C.; Nguyen, K.Y.; Ravussin, A.; Pistell, P.; Newman, S.; Carter, R.; Laque, A.; et al. Canonical Nlrp3 Inflammasome Links Systemic Low-Grade Inflammation to Functional Decline in Aging. Cell Metab. 2013, 18, 519–532. [Google Scholar] [CrossRef] [Green Version]
  22. He, C.; Tsuchiyama, S.K.; Nguyen, Q.T.; Plyusnina, E.N.; Terrill, S.R.; Sahibzada, S.; Patel, B.; Faulkner, A.R.; Shaposhnikov, M.; Tian, R.; et al. Enhanced Longevity by Ibuprofen, Conserved in Multiple Species, Occurs in Yeast through Inhibition of Tryptophan Import. PLoS Genet. 2014, 10, e1004860. [Google Scholar] [CrossRef] [Green Version]
  23. Ching, T.-T.; Chiang, W.-C.; Chen, C.-S.; Hsu, A.-L. Celecoxib extends C. elegans lifespan via inhibition of insulin-like signaling but not cyclooxygenase-2 activity. Aging Cell 2011, 10, 506–519. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Strong, R.; Miller, R.A.; Astle, C.M.; Floyd, R.A.; Flurkey, K.; Hensley, K.L.; Javors, M.A.; Leeuwenburgh, C.; Nelson, J.F.; Ongini, E.; et al. Nordihydroguaiaretic acid and aspirin increase lifespan of genetically heterogeneous male mice. Aging Cell 2008, 7, 641–650. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Lee, M.E.; Kim, S.R.; Lee, S.; Jung, Y.-J.; Choi, S.S.; Kim, W.J.; Han, J.A. Cyclooxygenase-2 inhibitors modulate skin aging in a catalytic activity-independent manner. Exp. Mol. Med. 2012, 44, 536–544. [Google Scholar] [CrossRef] [PubMed]
  26. Kalonia, H.; Kumar, P.; Kumar, A. Licofelone attenuates quinolinic acid induced Huntington like symptoms: Possible behavioral, biochemical and cellular alterations. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2011, 35, 607–615. [Google Scholar] [CrossRef] [PubMed]
  27. Asanuma, M.; Miyazaki, I.; Diaz-Corrales, F.J.; Ogawa, N. Quinone formation as dopaminergic neuron-specific oxidative stress in the pathogenesis of sporadic Parkinson’s disease and neurotoxin-induced parkinsonism. Acta Medica Okayama 2004, 58, 221–233. [Google Scholar] [PubMed]
  28. Black, P.H. Stress and the inflammatory response: A review of neurogenic inflammation. Brain Behav. Immun. 2002, 16, 622–653. [Google Scholar] [CrossRef]
  29. Choi, S.H.; Aid, S.; Caracciolo, L.; Sakura Minami, S.; Niikura, T.; Matsuoka, Y.; Turner, R.S.; Mattson, M.P.; Bosetti, F. Cyclooxygenase-1 inhibition reduces amyloid pathology and improves memory deficits in a mouse model of Alzheimer’s disease. J. Neurochem. 2013, 124, 59–68. [Google Scholar] [CrossRef] [Green Version]
  30. Calfio, C.; Gonzalez, A.; Singh, S.K.; Rojo, L.E.; Maccioni, R.B. The Emerging Role of Nutraceuticals and Phytochemicals in the Prevention and Treatment of Alzheimer’s Disease. J. Alzheimer’s Dis. 2020, 77, 33–51. [Google Scholar] [CrossRef]
  31. Veurink, G.; Perry, G.; Singh, S.K. Role of antioxidants and a nutrient rich diet in Alzheimer’s disease. Open Biol. 2020, 10, 200084. [Google Scholar] [CrossRef]
  32. Minois, N. Molecular Basis of the ‘Anti-Aging’ Effect of Spermidine and Other Natural Polyamines—A Mini-Review. Gerontology 2014, 60, 319–326. [Google Scholar] [CrossRef]
  33. Madeo, F.; Carmona-Gutierrez, D.; Kepp, O.; Kroemer, G. Spermidine delays aging in humans. Aging 2018, 10, 2209–2211. [Google Scholar] [CrossRef]
  34. Phillips, T.; Leeuwenburgh, C. Lifelong Aspirin Supplementation as a Means to Extending Life Span. Rejuvenation Res. 2004, 7, 243–252. [Google Scholar] [CrossRef]
  35. Ayyadevara, S.; Bharill, P.; Dandapat, A.; Hu, C.; Khaidakov, M.; Mitra, S.; Shmookler Reis, R.J.; Mehta, J.L. Aspirin inhibits oxidant stress, reduces age-associated functional declines, and extends lifespan of Caenorhabditis elegans. Antioxid. Redox Signal. 2013, 18, 481–490. [Google Scholar] [CrossRef]
  36. Nadon, N.L.; Strong, R.; Miller, R.A.; Harrison, D.E. NIA Interventions Testing Program: Investigating Putative Aging Intervention Agents in a Genetically Heterogeneous Mouse Model. eBioMedicine 2017, 21, 3–4. [Google Scholar] [CrossRef] [Green Version]
  37. Orhan, H.; Doğruer, D.; Cakir, B.; Şahin, G.; Şahin, M. The in vitro effects of new non-steroidal antiinflammatory compounds on antioxidant system of human erythrocytes. Exp. Toxicol. Pathol. 1999, 51, 397–402. [Google Scholar] [CrossRef]
  38. Danilov, A.; Shaposhnikov, M.; Shevchenko, O.; Zemskaya, N.; Zhavoronkov, A.; Moskalev, A. Influence of non-steroidal anti-inflammatory drugs on Drosophila melanogaster longevity. Oncotarget 2015, 6, 19428–19444. [Google Scholar] [CrossRef] [Green Version]
  39. Hosseini, S.; Abdollahi, M.; Azizi, G.; Fattahi, M.J.; Rastkari, N.; Zavareh, F.T.; Aghazadeh, Z.; Mirshafiey, A. Anti-aging effects of M2000 (β-D-mannuronic acid) as a novel immunosuppressive drug on the enzymatic and non-enzymatic oxidative stress parameters in an experimental model. J. Basic Clin. Physiol. Pharmacol. 2017, 28, 249–255. [Google Scholar] [CrossRef] [PubMed]
  40. Waditee-Sirisattha, R.; Kageyama, H. Protective effects of mycosporine-like amino acid-containing emulsions on UV-treated mouse ear tissue from the viewpoints of antioxidation and antiglycation. J. Photochem. Photobiol. B 2021, 223, 112296. [Google Scholar] [CrossRef]
  41. Kageyama, H.; Waditee-Sirisattha, R. Antioxidative, Anti-Inflammatory, and Anti-Aging Properties of Mycosporine-Like Amino Acids: Molecular and Cellular Mechanisms in the Protection of Skin-Aging. Mar. Drugs 2019, 17, 222. [Google Scholar] [CrossRef] [Green Version]
  42. Kumar Singh, S.; Barreto, G.E.; Aliev, G.; Echeverria, V. Ginkgo biloba as an Alternative Medicine in the Treatment of Anxiety in Dementia and other Psychiatric Disorders. Curr. Drug. Metab. 2017, 18, 112–119. [Google Scholar] [CrossRef]
  43. Singh, S.K.; Srikrishna, S.; Castellani, R.J.; Perry, G. Antioxidants in the prevention and treatment of alzheimer’s disease. In Nutritional Antioxidant Therapies: Treatments and Perspectives; Springer International Publishing: Berlin/Heidelberg, Germany, 2018; pp. 523–553. [Google Scholar] [CrossRef]
  44. Singh, S.K.; Srivastav, S.; Castellani, R.J.; Plascencia-Villa, G.; Perry, G. Neuroprotective and Antioxidant Effect of Ginkgo biloba Extract Against AD and Other Neurological Disorders. Neurotherapeutics 2019, 16, 666–674. [Google Scholar] [CrossRef]
  45. Mishra, S.; Mishra, S.K.; Singh, S.K. Ayurveda and Yoga practices: A synergistic approach for the treatment of Alzheimer’s disease. Eur. J. Biol. Res. 2020, 11, 65–74. [Google Scholar]
  46. Pietta, P.G. Flavonoids as Antioxidants. J. Nat. Prod. 2000, 63, 1035–1042. [Google Scholar] [CrossRef]
  47. Lobo, V.; Patil, A.; Phatak, A.; Chandra, N. Free radicals, antioxidants and functional foods: Impact on human health. Pharmacogn. Rev. 2010, 4, 118–126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Kampkötter, A.; Timpel, C.; Zurawski, R.F.; Ruhl, S.; Chovolou, Y.; Proksch, P.; Wätjen, W. Increase of stress resistance and lifespan of Caenorhabditis elegans by quercetin. Comp. Biochem. Physiol. Part B Biochem. Mol. Biol. 2008, 149, 314–323. [Google Scholar] [CrossRef]
  49. Sugawara, T.; Sakamoto, K. Quercetin enhances motility in aged and heat-stressed Caenorhabditis elegans nematodes by modulating both HSF-1 activity, and insulin-like and p38-MAPK signalling. PLoS ONE 2020, 15, e0238528. [Google Scholar] [CrossRef]
  50. Pietsch, K.; Saul, N.; Swain, S.C.; Menzel, R.; Steinberg, C.E.; Stürzenbaum, S.R. Meta-analysis of global transcriptomics suggests that conserved genetic pathways are responsible for quercetin and tannic acid mediated longevity in C. elegans. Front. Genet. 2012, 3, 48. [Google Scholar] [CrossRef] [Green Version]
  51. Proshkina, E.; Lashmanova, E.; Dobrovolskaya, E.; Zemskaya, N.; Kudryavtseva, A.; Shaposhnikov, M.; Moskalev, A. Geroprotective and Radioprotective Activity of Quercetin, (-)-Epicatechin, and Ibuprofen in Drosophila melanogaster. Front. Pharmacol. 2016, 7, 505. [Google Scholar] [CrossRef]
  52. Končič, M.; Rajić, Z.; Petrič, N.; Zorc, B. Antioxidant activity of NSAID hydroxamic acids. Acta Pharm. 2009, 59, 235–242. [Google Scholar] [CrossRef] [PubMed]
  53. Orhan, H.; Şahin, G. In vitro effects of NSAIDS and paracetamolon oxidative stress-related parameters of human erythrocytes. Exp. Toxicol. Pathol. 2001, 53, 133–140. [Google Scholar] [CrossRef]
  54. Peng, C.; Wang, X.; Chen, J.; Jiao, R.; Wang, L.; Li, Y.M.; Zuo, Y.; Liu, Y.; Lei, L.; Ma, K.Y.; et al. Biology of Ageing and Role of Dietary Antioxidants. BioMed Res. Int. 2014, 2014, 831841. [Google Scholar] [CrossRef] [Green Version]
  55. O’Sullivan, R.J.; Karlseder, J. Telomeres: Protecting chromosomes against genome instability. Nat. Rev. Mol. Cell Biol. 2010, 11, 171–181. [Google Scholar] [CrossRef] [Green Version]
  56. Lingner, J.; Cooper, J.P.; Cech, T.R. Telomerase and DNA end replication: No longer a lagging strand problem? Science 1995, 269, 1533–1534. [Google Scholar] [CrossRef]
  57. Greider, C.W. Telomerase is processive. Mol. Cell. Biol. 1991, 11, 4572–4580. [Google Scholar]
  58. Cesare, A.J.; Reddel, R.R. Alternative lengthening of telomeres: Models, mechanisms and implications. Nat. Rev. Genet. 2010, 11, 319–330. [Google Scholar] [CrossRef] [PubMed]
  59. Victorelli, S.; Passos, J.F. Telomeres and Cell Senescence—Size Matters Not. eBioMedicine 2017, 21, 14–20. [Google Scholar] [CrossRef] [Green Version]
  60. Dodig, S.; Čepelak, I.; Pavić, I. Hallmarks of senescence and aging. Biochem. Med. 2019, 29, 483–497. [Google Scholar] [CrossRef] [PubMed]
  61. Cassidy, A.; De Vivo, I.; Liu, Y.; Han, J.; Prescott, J.; Hunter, D.J.; Rimm, E.B. Associations between diet, lifestyle factors, and telomere length in women. Am. J. Clin. Nutr. 2010, 91, 1273–1280. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Prescott, J.; Kraft, P.; Chasman, D.I.; Savage, S.A.; Mirabello, L.; Berndt, S.I.; Weissfeld, J.L.; Han, J.; Hayes, R.B.; Chanock, S.J.; et al. Genome-Wide Association Study of Relative Telomere Length. PLoS ONE 2011, 6, e19635. [Google Scholar] [CrossRef]
  63. Song, S.; Johnson, F.B. Epigenetic Mechanisms Impacting Aging: A Focus on Histone Levels and Telomeres. Genes 2018, 9, 201. [Google Scholar] [CrossRef] [Green Version]
  64. Kuszel, L.; Trzeciak, T.; Richter, M.; Czarny-Ratajczak, M. Osteoarthritis and telomere shortening. J. Appl. Genet. 2015, 56, 169–176. [Google Scholar] [CrossRef] [Green Version]
  65. Carlquist, J.F.; Knight, S.; Cawthon, R.M.; Le, V.; Bunch, T.J.; Horne, B.D.; Rollo, J.S.; Huntinghouse, J.A.; Muhlestein, J.B.; Anderson, J.L. Shortened telomere length is associated with paroxysmal atrial fibrillation among cardiovascular patients enrolled in the Intermountain Heart Collaborative Study. Hear. Rhythm 2016, 13, 21–27. [Google Scholar] [CrossRef]
  66. Hunt, S.C.; Kimura, M.; Hopkins, P.N.; Carr, J.J.; Heiss, G.; Province, M.A.; Aviv, A. Leukocyte Telomere Length and Coronary Artery Calcium. Am. J. Cardiol. 2015, 116, 214–218. [Google Scholar] [CrossRef] [Green Version]
  67. Blasco, M.A. Telomere length, stem cells and aging. Nat. Chem. Biol. 2007, 3, 640–649. [Google Scholar] [CrossRef]
  68. Pereira, B.; Ferreira, M.G. Sowing the seeds of cancer: Telomeres and age-associated tumorigenesis. Curr. Opin. Oncol. 2013, 25, 93–98. [Google Scholar] [CrossRef]
  69. Wang, S.; Madu, C.O.; Lu, Y. Telomere and Its Role in Diseases. Oncomedicine 2019, 4, 1–9. [Google Scholar] [CrossRef] [Green Version]
  70. Bernades de Jesus, B.; Schneeberger, K.; Vera, E.; Tejera, A.; Harley, C.B.; Blasco, M.A. The telomerase activator TA-65 elongates short telomeres and increases health span of adult/old mice without increasing cancer incidence. Aging Cell 2011, 10, 604–621. [Google Scholar] [CrossRef] [Green Version]
  71. Bär, C.; Huber, N.; Beier, F.; Blasco, M.A. Therapeutic effect of androgen therapy in a mouse model of aplastic anemia produced by short telomeres. Haematologica 2015, 100, 1267–1274. [Google Scholar] [CrossRef] [Green Version]
  72. Calado, R.T.; Yewdell, W.T.; Wilkerson, K.L.; Regal, J.A.; Kajigaya, S.; Stratakis, C.A.; Young, N.S. Sex hormones, acting on the TERT gene, increase telomerase activity in human primary hematopoietic cells. Blood J. Am. Soc. Hematol. 2009, 114, 2236–2243. [Google Scholar] [CrossRef]
  73. de Jesus, B.B.; Vera, E.; Schneeberger, K.; Tejera, A.M.; Ayuso, E.; Bosch, F.; Blasco, M.A. Telomerase gene therapy in adult and old mice delays aging and increases longevity without increasing cancer. EMBO Mol. Med. 2012, 4, 691–704. [Google Scholar] [CrossRef]
  74. Raghavan, P.R. Metadichol®: A Novel Nanolipid Formulation That Inhibits SARS-CoV-2 and a Multitude of Pathological Viruses In Vitro. BioMed Res. Int. 2020. [Google Scholar] [CrossRef]
  75. Tsoukalas, D.; Fragkiadaki, P.; Docea, A.O.; Alegakis, A.K.; Sarandi, E.; Thanasoula, M.; Spandidos, D.A.; Tsatsakis, A.; Razgonova, M.P.; Calina, D. Discovery of potent telomerase activators: Unfolding new therapeutic and anti-aging perspectives. Mol. Med. Rep. 2019, 20, 3701–3708. [Google Scholar] [CrossRef] [Green Version]
  76. Shay, J.W. Role of Telomeres and Telomerase in Aging and Cancer. Cancer Discov. 2016, 6, 584–593. [Google Scholar] [CrossRef] [Green Version]
  77. Sedivy, J.M.; Banumathy, G.; Adams, P.D. Aging by epigenetics—A consequence of chromatin damage? Exp. Cell. Res. 2008, 314, 1909–1917. [Google Scholar] [CrossRef] [Green Version]
  78. Feser, J.; Tyler, J. Chromatin structure as a mediator of aging. FEBS Lett. 2011, 585, 2041–2048. [Google Scholar] [CrossRef] [Green Version]
  79. O’Sullivan, R.J.; Karlseder, J. The great unravelling: Chromatin as a modulator of the aging process. Trends Biochem. Sci. 2012, 37, 466–476. [Google Scholar] [CrossRef] [Green Version]
  80. Imai, S.-I.; Armstrong, C.M.; Kaeberlein, M.; Guarente, L. Transcriptional silencing and longevity protein Sir2 is an NAD-dependent histone deacetylase. Nature 2000, 403, 795–800. [Google Scholar] [CrossRef]
  81. Longo, V.D. Linking sirtuins, IGF-I signaling, and starvation. Exp. Gerontol. 2009, 44, 70–74. [Google Scholar] [CrossRef]
  82. Zhang, L.; Xie, W.J.; Liu, S.; Meng, L.; Gu, C.; Gao, Y.Q. DNA Methylation Landscape Reflects the Spatial Organization of Chromatin in Different Cells. Biophys. J. 2017, 113, 1395–1404. [Google Scholar] [CrossRef]
  83. McClay, J.L.; Aberg, K.A.; Clark, S.L.; Nerella, S.; Kumar, G.; Xie, L.Y.; Hudson, A.D.; Harada, A.; Hultman, C.M.; Magnusson, P.K.; et al. A methylome-wide study of aging using massively parallel sequencing of the methyl-CpG-enriched genomic fraction from blood in over 700 subjects. Hum. Mol. Genet. 2014, 23, 1175–1185. [Google Scholar] [CrossRef] [Green Version]
  84. Teschendorff, A.E.; Menon, U.; Gentry-Maharaj, A.; Ramus, S.J.; Weisenberger, D.J.; Shen, H.; Campan, M.; Noushmehr, H.; Bell, C.G.; Maxwell, A.P.; et al. Age-dependent DNA methylation of genes that are suppressed in stem cells is a hallmark of cancer. Genome Res. 2010, 20, 440–446. [Google Scholar] [CrossRef] [Green Version]
  85. Horvath, S. DNA methylation age of human tissues and cell types. Genome Biol. 2013, 14, R115. [Google Scholar] [CrossRef] [Green Version]
  86. Pal, S.; Tyler, J.K. Epigenetics and aging. Sci. Adv. 2016, 2. [Google Scholar] [CrossRef] [Green Version]
  87. Jylhava, J. Determinants of longevity: Genetics, biomarkers and therapeutic approaches. Current Pharmaceutical Design 2014, 20, 6058–6070. [Google Scholar] [CrossRef]
  88. Morselli, E.; Mariño, G.; Bennetzen, M.V.; Eisenberg, T.; Megalou, E.; Schroeder, S.; Cabrera, S.; Bénit, P.; Rustin, P.; Criollo, A.; et al. Spermidine and resveratrol induce autophagy by distinct pathways converging on the acetylproteome. J. Cell Biol. 2011, 192, 615–629. [Google Scholar] [CrossRef] [Green Version]
  89. Pasyukova, E.G.; Vaiserman, A.M. HDAC inhibitors: A new promising drug class in anti-aging research. Mech. Ageing Dev. 2017, 166, 6–15. [Google Scholar] [CrossRef]
  90. McIntyre, R.L.; Daniels, E.G.; Molenaars, M.; Houtkooper, R.H.; Janssens, G.E. From molecular promise to preclinical results: HDAC inhibitors in the race for healthy aging drugs. EMBO Mol. Med. 2019, 11, e9854. [Google Scholar] [CrossRef]
  91. Hetz, C.; Chevet, E.; Oakes, S.A. Proteostasis control by the unfolded protein response. Nat. Cell Biol. 2015, 17, 829–838. [Google Scholar] [CrossRef] [Green Version]
  92. Cannizzo, E.S.; Clement, C.C.; Morozova, K.; Valdor, R.; Kaushik, S.; Almeida, L.N.; Follo, C.; Sahu, R.; Cuervo, A.M.; Macian, F.; et al. Age-Related Oxidative Stress Compromises Endosomal Proteostasis. Cell Rep. 2012, 2, 136–149. [Google Scholar] [CrossRef]
  93. Brehm, A.; Krüger, E. Dysfunction in protein clearance by the proteasome: Impact on autoinflammatory diseases. Semin. Immunopathol. 2015, 37, 323–333. [Google Scholar] [CrossRef]
  94. Dai, D.-F.; Karunadharma, P.P.; Chiao, Y.A.; Basisty, N.; Crispin, D.; Hsieh, E.J.; Chen, T.; Gu, H.; Djukovic, D.; Raftery, D.; et al. Altered proteome turnover and remodeling by short-term caloric restriction or rapamycin rejuvenate the aging heart. Aging Cell 2014, 13, 529–539. [Google Scholar] [CrossRef]
  95. Gong, Z.; Tasset, I. Humanin enhances the cellular response to stress by activation of chaperone-mediated autophagy. Oncotarget 2018, 9, 10832–10833. [Google Scholar] [CrossRef]
  96. Amm, I.; Sommer, T.; Wolf, D.H. Protein quality control and elimination of protein waste: The role of the ubiquitin–proteasome system. Biochim. Biophys. Acta BBA—Mol. Cell Res. 2013, 1843, 182–196. [Google Scholar] [CrossRef] [Green Version]
  97. Koga, H.; Kaushik, S.; Cuervo, A.M. Protein homeostasis and aging: The importance of exquisite quality control. Ageing Res. Rev. 2011, 10, 205–215. [Google Scholar] [CrossRef] [Green Version]
  98. Calderwood, S.K.; Murshid, A.; Prince, T. The Shock of Aging: Molecular Chaperones and the Heat Shock Response in Longevity and Aging—A Mini-Review. Gerontology 2009, 55, 550–558. [Google Scholar] [CrossRef] [Green Version]
  99. Morrow, G.; Samson, M.; Michaud, S.; Tanguay, R.M. Overexpression of the small mitochondrial Hsp22 extends Drosophila life span and increases resistance to oxidative stress. FASEB J. 2004, 18, 598–599. [Google Scholar] [CrossRef] [Green Version]
  100. Walker, G.A.; Lithgow, G.J. Lifespan extension in C. elegans by a molecular chaperone dependent upon insulin-like signals. Aging Cell 2003, 2, 131–139. [Google Scholar] [CrossRef] [Green Version]
  101. Chiang, W.C.; Ching, T.T.; Lee, H.C.; Mousigian, C.; Hsu, A.L. HSF-1 Regulators DDL-1/2 Link Insulin-like Signaling to Heat-Shock Responses and Modulation of Longevity. Cell 2012, 148, 322–334. [Google Scholar] [CrossRef] [Green Version]
  102. Hsu, A.-L.; Murphy, C.T.; Kenyon, C. Regulation of Aging and Age-Related Disease by DAF-16 and Heat-Shock Factor. Science 2003, 300, 1142–1145. [Google Scholar] [CrossRef]
  103. Ünal, E.; Kinde, B.; Amon, A. Gametogenesis Eliminates Age-Induced Cellular Damage and Resets Life Span in Yeast. Science 2011, 332, 1554–1557. [Google Scholar] [CrossRef] [Green Version]
  104. Erjavec, N.; Larsson, L.; Grantham, J.; Nyström, T. Accelerated aging and failure to segregate damaged proteins in Sir2 mutants can be suppressed by overproducing the protein aggregation-remodeling factor Hsp104p. Genes Dev. 2007, 21, 2410–2421. [Google Scholar] [CrossRef] [Green Version]
  105. Kaeberlein, M.; Kirkland, K.T.; Fields, S.; Kennedy, B.K. Genes determining yeast replicative life span in a long-lived genetic background. Mech. Ageing Dev. 2005, 126, 491–504. [Google Scholar] [CrossRef]
  106. Andersson, V.; Hanzén, S.; Liu, B.; Molin, M.; Nyström, T. Enhancing protein disaggregation restores proteasome activity in aged cells. Aging 2013, 5, 802–812. [Google Scholar] [CrossRef] [Green Version]
  107. Press, M.; Jung, T.; König, J.; Grune, T.; Höhn, A. Protein aggregates and proteostasis in aging: Amylin and β-cell function. Mech. Ageing Dev. 2019, 177, 46–54. [Google Scholar] [CrossRef]
  108. Moreno, D.; Jenkins, K.; Morlot, S.; Charvin, G.; Csikasz-Nagy, A.; Aldea, M. Proteostasis collapse, a hallmark of aging, hinders the chaperone-Start network and arrests cells in G1. eLife 2019, 8, 48240. [Google Scholar] [CrossRef] [PubMed]
  109. Chondrogianni, N.; Georgila, K.; Kourtis, N.; Tavernarakis, N.; Gonos, E.S. 20S proteasome activation promotes life span extension and resistance to proteotoxicity in Caenorhabditis elegans. FASEB J. 2015, 29, 611–622. [Google Scholar] [CrossRef] [Green Version]
  110. Chondrogianni, N.; Tzavelas, C.; Pemberton, A.J.; Nezis, I.P.; Rivett, A.J.; Gonos, E.S. Overexpression of Proteasome β5 Assembled Subunit Increases the Amount of Proteasome and Confers Ameliorated Response to Oxidative Stress and Higher Survival Rates*[boxs]. J. Biol. Chem. 2005, 280, 11840–11850. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Tonoki, A.; Kuranaga, E.; Tomioka, T.; Hamazaki, J.; Murata, S.; Tanaka, K.; Miura, M. Genetic Evidence Linking Age-Dependent Attenuation of the 26S Proteasome with the Aging Process. Mol. Cell. Biol. 2009, 29, 1095–1106. [Google Scholar] [CrossRef] [Green Version]
  112. Papaevgeniou, N.; Sakellari, M.; Jha, S.; Tavernarakis, N.; Holmberg, C.I.; Gonos, E.S.; Chondrogianni, N. 18α-Glycyrrhetinic acid proteasome activator decelerates aging and Alzheimer’s disease progression in Caenorhabditis elegans and neuronal cultures. Antioxid. Redox Signal. 2016, 25, 855–869. [Google Scholar] [CrossRef]
  113. Eisenberg, T.; Knauer, H.; Schauer, A.; Büttner, S.; Ruckenstuhl, C.; Carmona-Gutierrez, D.; Ring, J.; Schroeder, S.; Magnes, C.; Antonacci, L.; et al. Induction of autophagy by spermidine promotes longevity. Nat. Cell Biol. 2009, 11, 1305–1314. [Google Scholar] [CrossRef]
  114. Kim, E.J.E.; Lee, S.-J.V. Recent progresses on anti-aging compounds and their targets in Caenorhabditis elegans. Transl. Med. Aging 2019, 3, 121–124. [Google Scholar] [CrossRef]
  115. Cheng, J.; North, B.J.; Zhang, T.; Dai, X.; Tao, K.; Guo, J.; Wei, W. The emerging roles of protein homeostasis-governing pathways in Alzheimer’s disease. Aging Cell 2018, 17, e12801. [Google Scholar] [CrossRef] [Green Version]
  116. Höhn, A.; Tramutola, A.; Cascella, R. Proteostasis Failure in Neurodegenerative Diseases: Focus on Oxidative Stress. Oxidative Med. Cell. Longev. 2020, 2020, 5497046. [Google Scholar] [CrossRef] [Green Version]
  117. Sigurdson, C.J.; Bartz, J.C.; Glatzel, M. Cellular and Molecular Mechanisms of Prion Disease. Annu. Rev. Pathol. Mech. Dis. 2019, 14, 497–516. [Google Scholar] [CrossRef]
  118. Wickner, R.B.; Bezsonov, E.E.; Son, M.; Ducatez, M.; DeWilde, M.; Edskes, H.K. Anti-Prion Systems in Yeast and Inositol Polyphosphates. Biochemistry 2018, 57, 1285–1292. [Google Scholar] [CrossRef]
  119. Lin, M.T.; Beal, M.F. Mitochondrial dysfunction and oxidative stress in neurodegenerative diseases. Nature 2006, 443, 787–795. [Google Scholar] [CrossRef]
  120. Fang, E.F.; Scheibye-Knudsen, M.; Chua, K.F.; Mattson, M.P.; Croteau, D.L.; Bohr, V.A. Nuclear DNA damage signalling to mitochondria in ageing. Nat. Rev. Mol. Cell Biol. 2016, 17, 308–321. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Babbar, M.; Basu, S.; Yang, B.; Croteau, D.L.; Bohr, V.A. Mitophagy and DNA damage signaling in human aging. Mech. Ageing Dev. 2020, 186, 111207. [Google Scholar] [CrossRef]
  122. Knuppertz, L.; Warnsmann, V.; Hamann, A.; Grimm, C.; Osiewacz, H.D. Stress-dependent opposing roles for mitophagy in aging of the ascomycete Podospora anserina. Autophagy 2017, 13, 1037–1052. [Google Scholar] [CrossRef] [PubMed]
  123. Jin, S.M.; Youle, R.J. PINK1- and Parkin-mediated mitophagy at a glance. J. Cell Sci. 2012, 125, 795–799. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Greene, J.C.; Whitworth, A.J.; Kuo, I.; Andrews, L.A.; Feany, M.B.; Pallanck, L.J. Mitochondrial pathology and apoptotic muscle degeneration in Drosophila parkin mutants. Proc. Natl. Acad. Sci. USA 2003, 100, 4078–4083. [Google Scholar] [CrossRef] [Green Version]
  125. Rana, A.; Rera, M.; Walker, D.W. Parkin overexpression during aging reduces proteotoxicity, alters mitochondrial dynamics, and extends lifespan. Proc. Natl. Acad. Sci. USA 2013, 110, 8638–8643. [Google Scholar] [CrossRef] [Green Version]
  126. López-Armada, M.J.; Riveiro-Naveira, R.R.; Vaamonde-García, C.; Valcárcel-Ares, M.N. Mitochondrial dysfunction and the inflammatory response. Mitochondrion 2013, 13, 106–118. [Google Scholar] [CrossRef]
  127. Hashemzaei, M.; Heravi, R.E.; Rezaee, R.; Roohbakhsh, A.; Karimi, G. Regulation of autophagy by some natural products as a potential therapeutic strategy for cardiovascular disorders. Eur. J. Pharmacol. 2017, 802, 44–51. [Google Scholar] [CrossRef]
  128. Yoshino, J.; Mills, K.F.; Yoon, M.J.; Imai, S.-I. Nicotinamide Mononucleotide, a Key NAD+ Intermediate, Treats the Pathophysiology of Diet- and Age-Induced Diabetes in Mice. Cell Metab. 2011, 14, 528–536. [Google Scholar] [CrossRef] [Green Version]
  129. Mouchiroud, L.; Houtkooper, R.H.; Moullan, N.; Katsyuba, E.; Ryu, D.; Cantó, C.; Mottis, A.; Jo, Y.S.; Viswanathan, M.; Schoonjans, K.; et al. The NAD+/sirtuin pathway modulates longevity through activation of mitochondrial UPR and FOXO signaling. Cell 2013, 154, 430–441. [Google Scholar] [CrossRef] [Green Version]
  130. Bonkowski, M.S.; Sinclair, D.A. Slowing ageing by design: The rise of NAD+ and sirtuin-activating compounds. Nat. Rev. Mol. Cell Biol. 2016, 17, 679–690. [Google Scholar] [CrossRef]
  131. Feige, J.N.; Lagouge, M.; Canto, C.; Strehle, A.; Houten, S.M.; Milne, J.C.; Lambert, P.D.; Mataki, C.; Elliott, P.J.; Auwerx, J. Specific SIRT1 Activation Mimics Low Energy Levels and Protects against Diet-Induced Metabolic Disorders by Enhancing Fat Oxidation. Cell Metab. 2008, 8, 347–358. [Google Scholar] [CrossRef]
  132. Canto, C.; Auwerx, J. PGC-1α, SIRT1 and AMPK, an energy sensing network that controls energy expenditure. Curr. Opin. Lipidol. 2009, 20, 98–105. [Google Scholar] [CrossRef] [PubMed]
  133. Park, D.; Jeong, H.; Lee, M.N.; Koh, A.; Kwon, O.; Yang, Y.R.; Noh, J.; Suh, P.-G.; Park, H.; Ryu, S.H. Resveratrol induces autophagy by directly inhibiting mTOR through ATP competition. Sci. Rep. 2016, 6, 21772. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Kuno, A.; Hosoda, R.; Sebori, R.; Hayashi, T.; Sakuragi, H.; Tanabe, M.; Horio, Y. Resveratrol Ameliorates Mitophagy Disturbance and Improves Cardiac Pathophysiology of Dystrophin-deficient mdx Mice. Sci. Rep. 2018, 8, 1–12. [Google Scholar] [CrossRef] [Green Version]
  135. Cao, S.; Shen, Z.; Wang, C.; Zhang, Q.; Hong, Q.; He, Y.; Hu, C. Resveratrol improves intestinal barrier function, alleviates mitochondrial dysfunction and induces mitophagy in diquat challenged piglets 1. Food Funct. 2019, 10, 344–354. [Google Scholar] [CrossRef]
  136. Pearson, K.J.; Baur, J.A.; Lewis, K.N.; Peshkin, L.; Price, N.L.; Labinskyy, N.; Swindell, W.R.; Kamara, D.; Minor, R.K.; Perez, E.; et al. Resveratrol Delays Age-Related Deterioration and Mimics Transcriptional Aspects of Dietary Restriction without Extending Life Span. Cell Metab. 2008, 8, 157–168. [Google Scholar] [CrossRef] [Green Version]
  137. Zarse, K.; Schmeisser, S.; Birringer, M.; Falk, E.; Schmoll, D.; Ristow, M. Differential Effects of Resveratrol and SRT1720 on Lifespan of Adult Caenorhabditis elegans. Horm. Metab. Res. 2010, 42, 837–839. [Google Scholar] [CrossRef]
  138. Poulsen, M.M.; Vestergaard, P.F.; Clasen, B.F.; Radko, Y.; Christensen, L.P.; Stødkilde-Jørgensen, H.; Møller, N.; Jessen, N.; Pedersen, S.B.; Jørgensen, J.O. High-dose resveratrol supplementation in obese men: An investigator-initiated, randomized, placebo-controlled clinical trial of substrate metabolism, insulin sensitivity, and body composition. Diabetes 2013, 62, 1186–1195. [Google Scholar] [CrossRef] [Green Version]
  139. Eisenberg, T.; Abdellatif, M.; Schroeder, S.; Primessnig, U.; Stekovic, S.; Pendl, T.; Harger, A.; Schipke, J.; Zimmermann, A.; Schmidt, A.; et al. Cardioprotection and lifespan extension by the natural polyamine spermidine. Nat. Med. 2016, 22, 1428–1438. [Google Scholar] [CrossRef]
  140. Ryu, D.; Mouchiroud, L.; Andreux, P.A.; Katsyuba, E.; Moullan, N.; Nicolet-Dit-Félix, A.A.; Williams, E.G.; Jha, P.; Lo Sasso, G.; Huzard, D.; et al. Urolithin A induces mitophagy and prolongs lifespan in C. elegans and increases muscle function in rodents. Nat. Med. 2016, 22, 879–888. [Google Scholar] [CrossRef]
  141. Andreux, P.A.; Blanco-Bose, W.; Ryu, D.; Burdet, F.; Ibberson, M.; Aebischer, P.; Auwerx, J.; Singh, A.; Rinsch, C. The mitophagy activator urolithin A is safe and induces a molecular signature of improved mitochondrial and cellular health in humans. Nat. Metab. 2019, 1, 595–603. [Google Scholar] [CrossRef]
  142. Fang, E.F.; Waltz, T.B.; Kassahun, H.; Lu, Q.; Kerr, J.S.; Morevati, M.; Fivenson, E.M.; Wollman, B.N.; Marosi, K.; Wilson, M.A.; et al. Tomatidine enhances lifespan and healthspan in C. elegans through mitophagy induction via the SKN-1/Nrf2 pathway. Sci. Rep. 2017, 7, 46208. [Google Scholar] [CrossRef] [PubMed]
  143. Richter, U.; Lahtinen, T.; Marttinen, P.; Myöhänen, M.; Greco, D.; Cannino, G.; Jacobs, H.T.; Lietzén, N.; Nyman, T.A.; Battersby, B.J. A Mitochondrial Ribosomal and RNA Decay Pathway Blocks Cell Proliferation. Curr. Biol. 2013, 23, 535–541. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Sun, N.; Yun, J.; Liu, J.; Malide, D.; Liu, C.; Rovira, I.I.; Holmström, K.M.; Fergusson, M.M.; Yoo, Y.H.; Combs, C.A.; et al. Measuring In Vivo Mitophagy. Mol. Cell 2015, 60, 685–696. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Xing, Y.; Liqi, Z.; Jian, L.; Qinghua, Y.; Qian, Y. Doxycycline Induces Mitophagy and Suppresses Production of Interferon-β in IPEC-J2 Cells. Front. Cell. Infect. Microbiol. 2017, 7, 21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Kujoth, G.C.; Bradshaw, P.C.; Haroon, S.; Prolla, T.A. The Role of Mitochondrial DNA Mutations in Mammalian Aging. PLoS Genet. 2007, 3, e24. [Google Scholar] [CrossRef] [Green Version]
  147. DeBalsi, K.L.; Hoff, K.E.; Copeland, W.C. Role of the mitochondrial DNA replication machinery in mitochondrial DNA mutagenesis, aging and age-related diseases. Ageing Res. Rev. 2017, 33, 89–104. [Google Scholar] [CrossRef] [Green Version]
  148. Morita, M.; Gravel, S.-P.; Chénard, V.; Sikström, K.; Zheng, L.; Alain, T.; Gandin, V.; Avizonis, D.; Arguello, M.; Zakaria, C.; et al. mTORC1 Controls Mitochondrial Activity and Biogenesis through 4E-BP-Dependent Translational Regulation. Cell Metab. 2013, 18, 698–711. [Google Scholar] [CrossRef] [Green Version]
  149. Bartolomé, A.; García-Aguilar, A.; Asahara, S.-I.; Kido, Y.; Guillén, C.; Pajvani, U.B.; Benito, M. MTORC1 Regulates both General Autophagy and Mitophagy Induction after Oxidative Phosphorylation Uncoupling. Mol. Cell. Biol. 2017, 37, e00441-17. [Google Scholar] [CrossRef] [Green Version]
  150. Narita, M.; Young, A.R.J.; Arakawa, S.; Samarajiwa, S.A.; Nakashima, T.; Yoshida, S.; Hong, S.; Berry, L.S.; Reichelt, S.; Ferreira, M.; et al. Spatial Coupling of mTOR and Autophagy Augments Secretory Phenotypes. Science 2011, 332, 966–970. [Google Scholar] [CrossRef] [Green Version]
  151. Iglesias-Bartolome, R.; Patel, V.; Cotrim, A.; Leelahavanichkul, K.; Molinolo, A.A.; Mitchell, J.B.; Gutkind, J.S. mTOR Inhibition Prevents Epithelial Stem Cell Senescence and Protects from Radiation-Induced Mucositis. Cell Stem Cell 2012, 11, 401–414. [Google Scholar] [CrossRef] [Green Version]
  152. Kolesnichenko, M.; Hong, L.; Liao, R.; Vogt, P.K.; Sun, P. Attenuation of TORC1 signaling delays replicative and oncogenic RAS-induced senescence. Cell Cycle 2012, 11, 2391–2401. [Google Scholar] [CrossRef] [PubMed]
  153. Vellai, T.; Takacs-Vellai, K.; Zhang, Y.; Kovacs, A.L.; Orosz, L.; Müller, F. Influence of TOR kinase on lifespan in C. elegans. Nature 2003, 426, 620. [Google Scholar] [CrossRef] [PubMed]
  154. Kapahi, P.; Zid, B.M.; Harper, T.; Koslover, D.; Sapin, V.; Benzer, S. Regulation of Lifespan in Drosophila by Modulation of Genes in the TOR Signaling Pathway. Curr. Biol. 2004, 14, 885–890. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Bai, H.; Post, S.; Kang, P.; Tatar, M. Drosophila Longevity Assurance Conferred by Reduced Insulin Receptor Substrate Chico Partially Requires d4eBP. PLoS ONE 2015, 10, e0134415. [Google Scholar] [CrossRef] [Green Version]
  156. Kaeberlein, M.; Powers, R.W., III; Steffen, K.K.; Westman, E.A.; Hu, D.; Dang, N.; Kerr, E.O.; Kirkland, K.T.; Fields, S.; Kennedy, B.K. Regulation of yeast replicative life span by TOR and Sch9 in response to nutrients. Science 2005, 310, 1193–1196. [Google Scholar] [CrossRef] [Green Version]
  157. Deprez, M.-A.; Eskes, E.; Winderickx, J.; Wilms, T. The TORC1-Sch9 pathway as a crucial mediator of chronological lifespan in the yeast Saccharomyces cerevisiae. FEMS Yeast Res. 2018, 18, 48. [Google Scholar] [CrossRef]
  158. Wu, J.J.; Liu, J.; Chen, E.B.; Wang, J.J.; Cao, L.; Narayan, N.; Fergusson, M.M.; Rovira, I.I.; Allen, M.; Springer, D.A.; et al. Increased Mammalian Lifespan and a Segmental and Tissue-Specific Slowing of Aging after Genetic Reduction of mTOR Expression. Cell Rep. 2013, 4, 913–920. [Google Scholar] [CrossRef] [Green Version]
  159. Powers, R.W.; Kaeberlein, M.; Caldwell, S.D.; Kennedy, B.K.; Fields, S. Extension of chronological life span in yeast by decreased TOR pathway signaling. Genes Dev. 2006, 20, 174–184. [Google Scholar] [CrossRef] [Green Version]
  160. Bjedov, I.; Toivonen, J.M.; Kerr, F.; Slack, C.; Jacobson, J.; Foley, A.; Partridge, L. Mechanisms of Life Span Extension by Rapamycin in the Fruit Fly Drosophila melanogaster. Cell Metab. 2010, 11, 35–46. [Google Scholar] [CrossRef] [Green Version]
  161. Scialo, F.; Sriram, A.; Naudi, A.; Ayala, V.; Jové, M.; Pamplona, R.; Sanz, A. Target of rapamycin activation predicts lifespan in fruit flies. Cell Cycle 2015, 14, 2949–2958. [Google Scholar] [CrossRef] [Green Version]
  162. Moskalev, A.; Shaposhnikov, M. Pharmacological Inhibition of Phosphoinositide 3 and TOR Kinases Improves Survival of Drosophila melanogaster. Rejuvenation Res. 2010, 13, 246–247. [Google Scholar] [CrossRef]
  163. Danilov, A.; Shaposhnikov, M.; Plyusnina, E.; Kogan, V.; Fedichev, P.; Moskalev, A. Selective anticancer agents suppress aging in Drosophila. Oncotarget 2013, 4, 1507–1526. [Google Scholar] [CrossRef] [Green Version]
  164. Fan, X.; Liang, Q.; Lian, T.; Wu, Q.; Gaur, U.; Li, D.; Yang, D.; Mao, X.; Jin, Z.; Li, Y.; et al. Rapamycin preserves gut homeostasis during Drosophila aging. Oncotarget 2015, 6, 35274–35283. [Google Scholar] [CrossRef] [Green Version]
  165. Robida-Stubbs, S.; Glover-Cutter, K.; Lamming, D.W.; Mizunuma, M.; Narasimhan, S.D.; Neumann-Haefelin, E.; Sabatini, D.M.; Blackwell, T.K. TOR signaling and rapamycin influence longevity by regulating SKN-1/Nrf and DAF-16/FoxO. Cell Metab. 2012, 15, 713–724. [Google Scholar] [CrossRef] [Green Version]
  166. 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. Rapamycin fed late in life extends lifespan in genetically heterogeneous mice. Nature 2009, 460, 392–395. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Anisimov, V.N.; Zabezhinski, M.A.; Popovich, I.G.; Piskunova, T.S.; Semenchenko, A.V.; Tyndyk, M.L.; Yurova, M.; Antoch, M.P.; Blagosklonny, M.V. Rapamycin Extends Maximal Lifespan in Cancer-Prone Mice. Am. J. Pathol. 2010, 176, 2092–2097. [Google Scholar] [CrossRef]
  168. Anisimov, V.N.; Zabezhinski, M.A.; Popovich, I.G.; Piskunova, T.S.; Semenchenko, A.V.; Tyndyk, M.L.; Yurova, M.; Rosenfeld, S.V.; Blagosklonny, M.V. Rapamycin increases lifespan and inhibits spontaneous tumorigenesis in inbred female mice. Cell Cycle 2011, 10, 4230–4236. [Google Scholar] [CrossRef] [PubMed]
  169. Wilkinson, J.E.; Burmeister, L.; Brooks, S.V.; Chan, C.-C.; Friedline, S.; Harrison, D.E.; Hejtmancik, J.F.; Nadon, N.; Strong, R.; Wood, L.K.; et al. Rapamycin slows aging in mice. Aging Cell 2012, 11, 675–682. [Google Scholar] [CrossRef] [Green Version]
  170. Comas, M.; Toshkov, I.; Kuropatwinski, K.K.; Chernova, O.B.; Polinsky, A.; Blagosklonny, M.V.; Gudkov, A.V.; Antoch, M.P. New nanoformulation of rapamycin Rapatar extends lifespan in homozygous p53−/− mice by delaying carcinogenesis. Aging 2012, 4, 715–722. [Google Scholar] [CrossRef] [PubMed]
  171. Komarova, E.A.; Antoch, M.P.; Novototskaya, L.R.; Chernova, O.B.; Paszkiewicz, G.; Leontieva, O.V.; Blagosklonny, M.V.; Gudkov, A.V. Rapamycin extends lifespan and delays tumorigenesis in heterozygous p53+/− mice. Aging 2012, 4, 709. [Google Scholar] [CrossRef]
  172. Ramos, F.J.; Chen, S.C.; Garelick, M.G.; Dai, D.F.; Liao, C.Y.; Schreiber, K.H.; MacKay, V.L.; An, E.H.; Strong, R.; Ladiges, W.C.; et al. Rapamycin reverses elevated mTORC1 signaling in lamin A/C–deficient mice, rescues cardiac and skeletal muscle function, and extends survival. Sci. Transl. Med. 2012, 4, 144ra103. [Google Scholar] [CrossRef]
  173. Livi, C.B.; Hardman, R.L.; Christy, B.A.; Dodds, S.G.; Jones, D.; Williams, C.; Strong, R.; Bokov, A.; Javors, M.A.; Ikeno, Y.; et al. Rapamycin extends life span of Rb1+/− mice by inhibiting neuroendocrine tumors. Aging 2013, 5, 100. [Google Scholar] [CrossRef] [Green Version]
  174. 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. Rapamycin-mediated lifespan increase in mice is dose and sex dependent and metabolically distinct from dietary restriction. Aging Cell 2014, 13, 468–477. [Google Scholar] [CrossRef] [PubMed]
  175. Popovich, I.G.; Anisimov, V.N.; Zabezhinski, M.A.; Semenchenko, A.V.; Tyndyk, M.L.; Yurova, M.N.; Blagosklonny, M.V. Lifespan extension and cancer prevention in HER-2/neu transgenic mice treated with low intermittent doses of rapamycin. Cancer Biol. Ther. 2014, 15, 586–592. [Google Scholar] [CrossRef] [Green Version]
  176. Zhang, Y.; Bokov, A.; Gelfond, J.; Soto, V.; Ikeno, Y.; Hubbard, G.; Diaz, V.; Sloane, L.; Maslin, K.; Treaster, S.; et al. Rapamycin extends life and health in C57BL/6 mice. J. Gerontol. Ser. A Biomed. Sci. Med. Sci. 2014, 69, 119–130. [Google Scholar] [CrossRef] [Green Version]
  177. Fok, W.C.Y.; Chen, Y.; Bokov, A.; Zhang, Y.; Salmon, A.; Diaz, V.; Javors, M.; Wood, W.H.; Zhang, Y.; Becker, K.; et al. Mice Fed Rapamycin Have an Increase in Lifespan Associated with Major Changes in the Liver Transcriptome. PLoS ONE 2014, 9, e83988. [Google Scholar] [CrossRef]
  178. Leontieva, O.V.; Paszkiewicz, G.M.; Blagosklonny, M.V. Weekly administration of rapamycin improves survival and biomarkers in obese male mice on high-fat diet. Aging Cell 2014, 13, 616–622. [Google Scholar] [CrossRef]
  179. Hasty, P.; Livi, C.B.; Dodds, S.G.; Jones, D.; Strong, R.; Javors, M.; Fischer, K.E.; Sloane, L.; Murthy, K.; Hubbard, G.; et al. eRapa Restores a Normal Life Span in a FAP Mouse ModelFAP Mice Can Live a Long Healthy Life. Cancer Prev. Res. 2014, 7, 169–178. [Google Scholar] [CrossRef] [Green Version]
  180. Fischer, K.E.; Gelfond, J.A.L.; Soto, V.Y.; Han, C.; Someya, S.; Richardson, A.; Austad, S.N. Health Effects of Long-Term Rapamycin Treatment: The Impact on Mouse Health of Enteric Rapamycin Treatment from Four Months of Age throughout Life. PLoS ONE 2015, 10, e0126644. [Google Scholar] [CrossRef]
  181. Ye, L.; Widlund, A.L.; Sims, C.A.; Lamming, D.W.; Guan, Y.; Davis, J.G.; Harrison, D.E.; Vang, O.; Baur, J.A.; Sabatini, D. Rapamycin doses sufficient to extend lifespan do not compromise muscle mitochondrial content or endurance. Aging 2013, 5, 7. [Google Scholar] [CrossRef] [Green Version]
  182. Johnson, S.C.; Yanos, M.E.; Kayser, E.-B.; Quintana, A.; Sangesland, M.; Castanza, A.; Uhde, L.; Hui, J.; Wall, V.Z.; Gagnidze, A.; et al. mTOR Inhibition Alleviates Mitochondrial Disease in a Mouse Model of Leigh Syndrome. Science 2013, 342, 1524–1528. [Google Scholar] [CrossRef] [PubMed]
  183. Hurez, V.; Dao, V.; Liu, A.; Pandeswara, S.; Gelfond, J.; Sun, L.; Bergman, M.; Orihuela, C.J.; Galvan, V.; Padrón, Á.; et al. Chronic mTOR inhibition in mice with rapamycin alters T, B, myeloid, and innate lymphoid cells and gut flora and prolongs life of immune-deficient mice. Aging Cell 2015, 14, 945–956. [Google Scholar] [CrossRef]
  184. Verlingue, L.; Dugourd, A.; Stoll, G.; Barillot, E.; Calzone, L.; Londoño-Vallejo, A. A comprehensive approach to the molecular determinants of lifespan using a Boolean model of geroconversion. Aging Cell 2016, 15, 1018–1026. [Google Scholar] [CrossRef]
  185. Johnson, S.C.; Yanos, M.E.; Bitto, A.; Castanza, A.; Gagnidze, A.; Gonzalez, B.; Gupta, K.; Hui, J.; Jarvie, C.; Johnson, B.M.; et al. Dose-dependent effects of mTOR inhibition on weight and mitochondrial disease in mice. Front. Genet. 2015, 6, 247. [Google Scholar] [CrossRef] [Green Version]
  186. Xue, Q.-L.; Yang, H.; Li, H.-F.; Abadir, P.M.; Burks, T.N.; Koch, L.G.; Britton, S.L.; Carlson, J.; Chen, L.; Walston, J.D.; et al. Rapamycin increases grip strength and attenuates age-related decline in maximal running distance in old low capacity runner rats. Aging 2016, 8, 769–776. [Google Scholar] [CrossRef] [Green Version]
  187. Bitto, A.; Ito, T.K.; Pineda, V.V.; Letexier, N.J.; Huang, H.; Sutlief, E.; Tung, H.; Vizzini, N.; Chen, B.; Smith, K.; et al. Transient rapamycin treatment can increase lifespan and healthspan in middle-aged mice. eLife 2016, 5, e16351. [Google Scholar] [CrossRef]
  188. Fang, Y.; Bartke, A. Prolonged Rapamycin treatment led to beneficial metabolic switch. Aging 2013, 5, 328–329. [Google Scholar] [CrossRef] [Green Version]
  189. Blagosklonny, M.V. Rapamycin extends life- and health span because it slows aging. Aging 2013, 5, 592–598. [Google Scholar] [CrossRef] [Green Version]
  190. Stanfel, M.N.; Shamieh, L.S.; Kaeberlein, M.; Kennedy, B.K. The TOR pathway comes of age. Biochim. Biophys. Acta BBA—Gen. Subj. 2009, 1790, 1067–1074. [Google Scholar] [CrossRef] [Green Version]
  191. Campistol, J.M.; Eris, J.; Oberbauer, R.; Friend, P.; Hutchison, B.; Morales, J.M.; Claesson, K.; Stallone, G.; Russ, G.; Rostaing, L.; et al. Sirolimus Therapy after Early Cyclosporine Withdrawal Reduces the Risk for Cancer in Adult Renal Transplantation. J. Am. Soc. Nephrol. 2006, 17, 581–589. [Google Scholar] [CrossRef] [Green Version]
  192. Kauffman, H.M.; Cherikh, W.S.; Cheng, Y.; Hanto, D.W.; Kahan, B.D. Maintenance Immunosuppression with Target-of-Rapamycin Inhibitors is Associated with a Reduced Incidence of De Novo Malignancies. Transplantation 2005, 80, 883–889. [Google Scholar] [CrossRef]
  193. Euvrard, S.; Morelon, E.; Rostaing, L.; Goffin, E.; Brocard, A.; Tromme, I.; Broeders, N.; del Marmol, V.; Chatelet, V.; Dompmartin, A.; et al. Sirolimus and Secondary Skin-Cancer Prevention in Kidney Transplantation. N. Engl. J. Med. 2012, 367, 329–339. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Pedro, J.M.B.-S.; Senovilla, L. Immunostimulatory activity of lifespan-extending agents. Aging 2013, 5, 793–801. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Mannick, J.B.; Del Giudice, G.; Lattanzi, M.; Valiante, N.M.; Praestgaard, J.; Huang, B.; Lonetto, M.A.; Maecker, H.T.; Kovarik, J.; Carson, S.; et al. mTOR inhibition improves immune function in the elderly. Sci. Transl. Med. 2014, 6, 268ra179. [Google Scholar] [CrossRef]
  196. Kennedy, B.K.; Pennypacker, J.K. Aging interventions get human. Oncotarget 2015, 6, 590–591. [Google Scholar] [CrossRef]
  197. Blagosklonny, M.V. Rejuvenating immunity: “Anti-aging drug today” eight years later. Oncotarget 2015, 6, 19405. [Google Scholar] [CrossRef] [Green Version]
  198. Ross, C.; Salmon, A.; Strong, R.; Fernandez, E.; Javors, M.; Richardson, A.; Tardif, S. Metabolic consequences of long-term rapamycin exposure on common marmoset monkeys (Callithrix jacchus). Aging 2015, 7, 964–973. [Google Scholar] [CrossRef] [Green Version]
  199. Zaseck, L.W.; Miller, R.A.; Brooks, S.V. Rapamycin Attenuates Age-associated Changes in Tibialis Anterior Tendon Viscoelastic Properties. J. Gerontol. Ser. A 2016, 71, 858–865. [Google Scholar] [CrossRef] [Green Version]
  200. Lelegren, M.; Liu, Y.; Ross, C.; Tardif, S.; Salmon, A.B. Pharmaceutical inhibition of mTOR in the common marmoset: Effect of rapamycin on regulators of proteostasis in a non-human primate. Pathobiol. Aging Age-Related Dis. 2016, 6, 31793. [Google Scholar] [CrossRef] [Green Version]
  201. Johnson, S.; Rabinovitch, P.S.; Kaeberlein, M. mTOR is a key modulator of ageing and age-related disease. Nat. 2013, 493, 338–345. [Google Scholar] [CrossRef] [Green Version]
  202. Lin, K.; Dorman, J.B.; Rodan, A.; Kenyon, C. daf-16: An HNF-3/forkhead Family Member That Can Function to Double the Life-Span of Caenorhabditis elegans. Science 1979, 278, 1319–1322. [Google Scholar] [CrossRef] [PubMed]
  203. Hwangbo, D.S.; Gersham, B.; Tu, M.-P.; Palmer, M.; Tatar, M. Drosophila dFOXO controls lifespan and regulates insulin signalling in brain and fat body. Nat. 2004, 429, 562–566. [Google Scholar] [CrossRef] [PubMed]
  204. Tatar, M.; Kopelman, A.; Epstein, D.; Tu, M.-P.; Yin, C.-M.; Garofalo, R.S. A Mutant Drosophila Insulin Receptor Homolog That Extends Life-Span and Impairs Neuroendocrine Function. Science 2001, 292, 107–110. [Google Scholar] [CrossRef] [Green Version]
  205. Selman, C.; Lingard, S.; Choudhury, A.I.; Batterham, R.L.; Claret, M.; Clements, M.; Ramadani, F.; Okkenhaug, K.; Schuster, E.; Blanc, E.; et al. Evidence for lifespan extension and delayed age–related biomarkers in insulin receptor substrate 1 null mice. FASEB J. 2008, 22, 807–818. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Selman, C.; Partridge, L.; Withers, D. Replication of Extended Lifespan Phenotype in Mice with Deletion of Insulin Receptor Substrate 1. PLoS ONE 2011, 6, e16144. [Google Scholar] [CrossRef]
  207. Anderson, R.M.; le Couteur, D.G.; de Cabo, R. Caloric Restriction Research: New Perspectives on the Biology of Aging. J. Gerontol. Ser. A 2018, 73, 1–3. [Google Scholar] [CrossRef] [Green Version]
  208. Gems, D.; Partridge, L. Genetics of Longevity in Model Organisms: Debates and Paradigm Shifts. Annu. Rev. Physiol. 2013, 75, 621–644. [Google Scholar] [CrossRef] [Green Version]
  209. Mattison, J.A.; Colman, R.J.; Beasley, T.M.; Allison, D.B.; Kemnitz, J.W.; Roth, G.S.; Ingram, D.K.; Weindruch, R.; De Cabo, R.; Anderson, R.M. Caloric restriction improves health and survival of rhesus monkeys. Nat. Commun. 2017, 8, 14063. [Google Scholar] [CrossRef] [Green Version]
  210. Li, J.; Cui, X.; Wang, Z.; Li, Y. rBTI extends Caenorhabditis elegans lifespan by mimicking calorie restriction. Exp. Gerontol. 2015, 67, 62–71. [Google Scholar] [CrossRef]
  211. Jeon, H.; Cha, D.S. Anti-aging properties of Ribes fasciculatum in Caenorhabditis elegans. Chin. J. Nat. Med. 2016, 14, 335–342. [Google Scholar] [CrossRef]
  212. Wan, Q.L.; Fu, X.; Meng, X.; Luo, Z.; Dai, W.; Yang, J.; Wang, C.; Wang, H.; Zhou, Q. Hypotaurine promotes longevity and stress tolerance via the stress response factors DAF-16/FOXO and SKN-1/NRF2 in Caenorhabditis elegans. Food Funct. 2020, 11, 347–357. [Google Scholar] [CrossRef]
  213. Steinberg, G.R.; Kemp, B.E. AMPK in Health and Disease. Physiol. Rev. 2009, 89, 1025–1078. [Google Scholar] [CrossRef]
  214. Hardie, D.G. AMP-activated protein kinase—An energy sensor that regulates all aspects of cell function. Genes Dev. 2011, 25, 1895–1908. [Google Scholar] [CrossRef] [Green Version]
  215. Mihaylova, M.M.; Shaw, R.J. The AMPK signalling pathway coordinates cell growth, autophagy and metabolism. Nat. Cell Biol. 2011, 13, 1016–1023. [Google Scholar] [CrossRef] [PubMed]
  216. Burnett, C.; Valentini, S.; Cabreiro, F.; Goss, M.; Somogyvári, M.; Piper, M.D.; Hoddinott, M.; Sutphin, G.L.; Leko, V.; McElwee, J.J.; et al. Absence of effects of Sir2 overexpression on lifespan in C. elegans and Drosophila. Nature 2011, 477, 482–485. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Kanfi, Y.; Naiman, S.; Amir, G.; Peshti, V.; Zinman, G.; Nahum, L.; Bar-Joseph, Z.; Cohen, H.Y. The sirtuin SIRT6 regulates lifespan in male mice. Nature 2012, 483, 218–221. [Google Scholar] [CrossRef] [PubMed]
  218. Kapahi, P.; Kaeberlein, M.; Hansen, M. Dietary restriction and lifespan: Lessons from invertebrate models. Ageing Res. Rev. 2017, 39, 3–14. [Google Scholar] [CrossRef]
  219. Satoh, A.; Brace, C.S.; Rensing, N.; Cliften, P.; Wozniak, D.F.; Herzog, E.D.; Yamada, K.A.; Imai, S.-I. Sirt1 Extends Life Span and Delays Aging in Mice through the Regulation of Nk2 Homeobox 1 in the DMH and LH. Cell Metab. 2013, 18, 416–430. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  220. Rogina, B.; Helfand, S.L. Sir2 mediates longevity in the fly through a pathway related to calorie restriction. Proc. Natl. Acad. Sci. USA 2004, 101, 15998–16003. [Google Scholar] [CrossRef] [Green Version]
  221. Kaeberlein, M.; McVey, M.; Guarente, L. The SIR2/3/4 complex and SIR2 alone promote longevity in Saccharomyces cerevisiae by two different mechanisms. Genes Dev. 1999, 13, 2570–2580. [Google Scholar] [CrossRef] [Green Version]
  222. Coughlan, K.A.; Valentine, R.J.; Ruderman, N.B.; Saha, A.K. AMPK activation: A therapeutic target for type 2 diabetes? Diabetes Metab. Syndr. Obes. 2014, 7, 241. [Google Scholar] [PubMed]
  223. Tissenbaum, H.A.; Guarente, L. Increased dosage of a sir-2 gene extends lifespan in Caenorhabditis elegans. Nature 2001, 410, 227–230. [Google Scholar] [CrossRef] [PubMed]
  224. Wang, N.; Luo, Z.; Jin, M.; Sheng, W.; Wang, H.-T.; Long, X.; Wu, Y.; Hu, P.; Xu, H.; Zhang, X. Exploration of age-related mitochondrial dysfunction and the anti-aging effects of resveratrol in zebrafish retina. Aging 2019, 11, 3117–3137. [Google Scholar] [CrossRef] [PubMed]
  225. Hubbard, B.P.; Sinclair, D.A. Small molecule SIRT1 activators for the treatment of aging and age-related diseases. Trends Pharmacol. Sci. 2014, 35, 146–154. [Google Scholar] [CrossRef] [Green Version]
  226. Minor, R.K.; Baur, J.A.; Gomes, A.P.; Ward, T.M.; Csiszar, A.; Mercken, E.M.; Abdelmohsen, K.; Shin, Y.-K.; Canto, C.; Scheibye-Knudsen, M.; et al. SRT1720 improves survival and healthspan of obese mice. Sci. Rep. 2011, 1, 70. [Google Scholar] [CrossRef] [Green Version]
  227. Milne, J.C.; Lambert, P.D.; Schenk, S.; Carney, D.P.; Smith, J.J.; Gagne, D.J.; Jin, L.; Boss, O.; Perni, R.B.; Vu, C.B.; et al. Small molecule activators of SIRT1 as therapeutics for the treatment of type 2 diabetes. Nature 2007, 450, 712–716. [Google Scholar] [CrossRef] [Green Version]
  228. Ota, H.; Eto, M.; Kano, M.R.; Ogawa, S.; Iijima, K.; Akishita, M.; Ouchi, Y. Cilostazol Inhibits Oxidative Stress–Induced Premature Senescence Via Upregulation of Sirt1 in Human Endothelial Cells. Arter. Thromb. Vasc. Biol. 2008, 28, 1634–1639. [Google Scholar] [CrossRef] [Green Version]
  229. Jamal, J.; Mustafa, M.R.; Wong, P.-F. Paeonol protects against premature senescence in endothelial cells by modulating Sirtuin 1 pathway. J. Ethnopharmacol. 2014, 154, 428–436. [Google Scholar] [CrossRef]
  230. Ota, H.; Eto, M.; Kano, M.R.; Kahyo, T.; Setou, M.; Ogawa, S.; Iijima, K.; Akishita, M.; Ouchi, Y. Induction of Endothelial Nitric Oxide Synthase, SIRT1, and Catalase by Statins Inhibits Endothelial Senescence Through the Akt Pathway. Arter. Thromb. Vasc. Biol. 2010, 30, 2205–2211. [Google Scholar] [CrossRef] [Green Version]
  231. Zheng, M.; Qiao, W.; Cui, J.; Liu, L.; Liu, H.; Wang, Z.; Yan, C. Hydrogen sulfide delays nicotinamide-induced premature senescence via upregulation of SIRT1 in human umbilical vein endothelial cells. Mol. Cell. Biochem. 2014, 393, 59–67. [Google Scholar] [CrossRef]
  232. Lee, Y.A.; Cho, E.J.; Yokozawa, T. Protective Effect of Persimmon (Diospyros kaki) Peel Proanthocyanidin against Oxidative Damage under H2O2-Induced Cellular Senescence. Biol. Pharm. Bull. 2008, 31, 1265–1269. [Google Scholar] [CrossRef]
  233. Jayasena, T.; Poljak, A.; Smythe, G.; Braidy, N.; Muench, G.; Sachdev, P. The role of polyphenols in the modulation of sirtuins and other pathways involved in Alzheimer’s disease. Ageing Res. Rev. 2013, 12, 867–883. [Google Scholar] [CrossRef] [PubMed]
  234. Apfeld, J.; O’Connor, G.; McDonagh, T.; DiStefano, P.S.; Curtis, R. The AMP-activated protein kinase AAK-2 links energy levels and insulin-like signals to lifespan in C. elegans. Genes Dev. 2004, 18, 3004–3009. [Google Scholar] [CrossRef] [Green Version]
  235. Onken, B.; Driscoll, M. Metformin Induces a Dietary Restriction–Like State and the Oxidative Stress Response to Extend C. elegans Healthspan via AMPK, LKB1, and SKN-1. PLoS ONE 2010, 5, e8758. [Google Scholar] [CrossRef] [PubMed]
  236. Anisimov, V.N. Metformin for aging and cancer prevention. Aging 2010, 2, 760–774. [Google Scholar] [CrossRef] [Green Version]
  237. Funakoshi, M.; Tsuda, M.; Muramatsu, K.; Hatsuda, H.; Morishita, S.; Aigaki, T. A gain-of-function screen identifies wdb and lkb1 as lifespan-extending genes in Drosophila. Biochem. Biophys. Res. Commun. 2011, 405, 667–672. [Google Scholar] [CrossRef] [PubMed]
  238. Martin-Montalvo, A.; Mercken, E.M.; Mitchell, S.J.; Palacios, H.H.; Mote, P.L.; Scheibye-Knudsen, M.; Gomes, A.P.; Ward, T.M.; Minor, R.K.; Blouin, M.-J.; et al. Metformin improves healthspan and lifespan in mice. Nat. Commun. 2013, 4, 2192. [Google Scholar] [CrossRef] [Green Version]
  239. Cabreiro, F.; Au, C.; Leung, K.-Y.; Vergara-Irigaray, N.; Cochemé, H.M.; Noori, T.; Weinkove, D.; Schuster, E.; Greene, N.D.; Gems, D. Metformin Retards Aging in C. elegans by Altering Microbial Folate and Methionine Metabolism. Cell 2013, 153, 228–239. [Google Scholar] [CrossRef] [Green Version]
  240. Piskovatska, V.; Stefanyshyn, N.; Storey, K.B.; Vaiserman, A.M.; Lushchak, O. Metformin as a geroprotector: Experimental and clinical evidence. Biogerontology 2019, 20, 33–48. [Google Scholar] [CrossRef] [PubMed]
  241. Curtis, R.; O’Connor, G.; DiStefano, P.S. Aging networks in Caenorhabditis elegans: AMP-activated protein kinase (aak-2) links multiple aging and metabolism pathways. Aging Cell 2006, 5, 119–126. [Google Scholar] [CrossRef]
  242. Park, S.-K.; Seong, R.-K.; Kim, J.-A.; Son, S.-J.; Kim, Y.; Yokozawa, T.; Shin, O.S. Oligonol promotes anti-aging pathways via modulation of SIRT1-AMPK-Autophagy Pathway. Nutr. Res. Pract. 2016, 10, 3–10. [Google Scholar] [CrossRef]
  243. Campisi, J. Aging, Cellular Senescence, and Cancer. Annu. Rev. Physiol. 2013, 75, 685. [Google Scholar] [CrossRef] [Green Version]
  244. van Deursen, J.M. The role of senescent cells in ageing. Nature 2014, 509, 439–446. [Google Scholar] [CrossRef] [Green Version]
  245. Ogrodnik, M.; Miwa, S.; Tchkonia, T.; Tiniakos, D.; Wilson, C.L.; Lahat, A.; Day, C.P.; Burt, A.; Palmer, A.; Anstee, Q.M.; et al. Cellular senescence drives age-dependent hepatic steatosis. Nat. Commun. 2017, 8, 15691. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  246. Campisi, J.; Robert, L. Cell Senescence: Role in Aging and Age-Related Diseases. Aging Facts Theor. 2014, 39, 45–61. [Google Scholar] [CrossRef] [Green Version]
  247. Muñoz-Espín, D.; Serrano, M. Cellular senescence: From physiology to pathology. Nat. Rev. Mol. Cell Biol. 2014, 15, 482–496. [Google Scholar] [CrossRef]
  248. Farr, J.N.; Xu, M.; Weivoda, M.M.; Monroe, D.G.; Fraser, D.G.; Onken, J.L.; Negley, B.A.; Sfeir, J.G.; Ogrodnik, M.B.; Hachfeld, C.M.; et al. Targeting cellular senescence prevents age-related bone loss in mice. Nat. Med. 2017, 23, 1072–1079. [Google Scholar] [CrossRef] [Green Version]
  249. Childs, B.G.; Baker, D.J.; Wijshake, T.; Conover, C.A.; Campisi, J.; Van Deursen, J.M. Senescent intimal foam cells are deleterious at all stages of atherosclerosis. Science 2016, 354, 472–477. [Google Scholar] [CrossRef] [Green Version]
  250. Baker, D.J.; Childs, B.G.; Durik, M.; Wijers, M.E.; Sieben, C.J.; Zhong, J.; Saltness, R.A.; Jeganathan, K.B.; Verzosa, G.C.; Pezeshki, A.; et al. Naturally occurring p16Ink4a-positive cells shorten healthy lifespan. Nature 2016, 530, 184–189. [Google Scholar] [CrossRef] [Green Version]
  251. Baker, D.J.; Wijshake, T.; Tchkonia, T.; Lebrasseur, N.K.; Childs, B.G.; Van De Sluis, B.; Kirkland, J.L.; Van Deursen, J.M. Clearance of p16Ink4a-positive senescent cells delays ageing-associated disorders. Nature 2011, 479, 232–236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  252. Sagiv, A.; Bar-Shai, A.; Levi, N.; Hatzav, M.; Zada, L.; Ovadya, Y.; Roitman, L.; Manella, G.; Regev, O.; Majewska, J.; et al. p53 in Bronchial Club Cells Facilitates Chronic Lung Inflammation by Promoting Senescence. Cell Rep. 2018, 22, 3468–3479. [Google Scholar] [CrossRef]
  253. Childs, B.G.; Gluscevic, M.; Baker, D.J.; Laberge, R.-M.; Marquess, D.; Dananberg, J.; Van Deursen, J.M. Senescent cells: An emerging target for diseases of ageing. Nat. Rev. Drug Discov. 2017, 16, 718–735. [Google Scholar] [CrossRef] [Green Version]
  254. von Kobbe, C. Targeting senescent cells: Approaches, opportunities, challenges. Aging 2019, 11, 12844. [Google Scholar] [CrossRef]
  255. Zhu, Y.I.; Tchkonia, T.; Pirtskhalava, T.; Gower, A.C.; Ding, H.; Giorgadze, N.; Palmer, A.K.; Ikeno, Y.; Hubbard, G.B.; Lenburg, M.; et al. The Achilles’ heel of senescent cells: From transcriptome to senolytic drugs. Aging Cell 2015, 14, 644–658. [Google Scholar] [CrossRef] [PubMed]
  256. Zhu, Y.; Doornebal, E.J.; Pirtskhalava, T.; Giorgadze, N.; Wentworth, M.; Fuhrmann-Stroissnigg, H.; Niedernhofer, L.J.; Robbins, P.D.; Tchkonia, T.; Kirkland, J.L. New agents that target senescent cells: The flavone, fisetin, and the BCL-XL inhibitors, A1331852 and A1155463. Aging 2017, 9, 955–963. [Google Scholar] [CrossRef] [Green Version]
  257. 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. Targeted Apoptosis of Senescent Cells Restores Tissue Homeostasis in Response to Chemotoxicity and Aging. Cell 2017, 169, 132–147.e16. [Google Scholar] [CrossRef] [Green Version]
  258. Chang, J.; Wang, Y.; Shao, L.; Laberge, R.-M.; DeMaria, M.; Campisi, J.; Janakiraman, K.; Sharpless, N.E.; Ding, S.; Feng, W.; et al. Clearance of senescent cells by ABT263 rejuvenates aged hematopoietic stem cells in mice. Nat. Med. 2017, 22, 78–83. [Google Scholar] [CrossRef] [Green Version]
  259. Chung, C.L.; Lawrence, I.; Hoffman, M.; Elgindi, D.; Nadhan, K.; Potnis, M.; Jin, A.; Sershon, C.; Binnebose, R.; Lorenzini, A.; et al. Topical rapamycin reduces markers of senescence and aging in human skin: An exploratory, prospective, randomized trial. GeroScience 2019, 41, 861–869. [Google Scholar] [CrossRef] [Green Version]
  260. Xu, M.; Tchkonia, T.; Ding, H.; Ogrodnik, M.; Lubbers, E.R.; Pirtskhalava, T.; White, T.A.; Johnson, K.O.; Stout, M.B.; Mezera, V.; et al. JAK inhibition alleviates the cellular senescence-associated secretory phenotype and frailty in old age. Proc. Natl. Acad. Sci. USA 2015, 112, E6301–E6310. [Google Scholar] [CrossRef] [Green Version]
  261. Tilstra, J.S.; Robinson, A.R.; Wang, J.; Gregg, S.Q.; Clauson, C.L.; Reay, D.P.; Nasto, L.A.; St Croix, C.M.; Usas, A.; Vo, N.; et al. NF-κB inhibition delays DNA damage–induced senescence and aging in mice. J. Clin. Investig. 2012, 122, 2601–2612. [Google Scholar] [CrossRef] [Green Version]
  262. Ozsvari, B.; Nuttall, J.R.; Sotgia, F.; Lisanti, M.P. Azithromycin and Roxithromycin define a new family of “senolytic” drugs that target senescent human fibroblasts. Aging 2018, 10, 3294–3307. [Google Scholar] [CrossRef] [PubMed]
  263. Bielak-Zmijewska, A.; Grabowska, W.; Ciolko, A.; Bojko, A.; Mosieniak, G.; Bijoch, Ł.; Sikora, E. The Role of Curcumin in the Modulation of Ageing. Int. J. Mol. Sci. 2019, 20, 1239. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Conboy, I.M.; Rando, T.A. Heterochronic parabiosis for the study of the effects of aging on stem cells and their niches. Cell Cycle 2012, 11, 2260–2267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  265. Gruber, R.; Koch, H.; Doll, B.A.; Tegtmeier, F.; Einhorn, T.A.; Hollinger, J.O. Fracture healing in the elderly patient. Exp. Gerontol. 2006, 41, 1080–1093. [Google Scholar] [CrossRef] [PubMed]
  266. Goodell, M.A.; Rando, T.A. Stem cells and healthy aging. Science 2015, 350, 1199–1204. [Google Scholar] [CrossRef] [PubMed]
  267. Rando, T.A. Stem cells, ageing and the quest for immortality. Nature 2006, 441, 1080–1086. [Google Scholar] [CrossRef]
  268. Molofsky, A.V.; Slutsky, S.G.; Joseph, N.M.; He, S.; Pardal, R.; Krishnamurthy, J.; Sharpless, N.; Morrison, S.J. Increasing p16INK4a expression decreases forebrain progenitors and neurogenesis during ageing. Nature 2006, 443, 448–452. [Google Scholar] [CrossRef] [Green Version]
  269. Lavasani, M.; Robinson, A.R.; Lu, A.; Song, M.; Feduska, J.M.; Ahani, B.; Tilstra, J.S.; Feldman, C.H.; Robbins, P.D.; Niedernhofer, L.J.; et al. Muscle-derived stem/progenitor cell dysfunction limits healthspan and lifespan in a murine progeria model. Nat. Commun. 2012, 3, 608. [Google Scholar] [CrossRef] [Green Version]
  270. Rando, T.A.; Chang, H.Y. Aging, Rejuvenation, and Epigenetic Reprogramming: Resetting the Aging Clock. Cell 2012, 148, 46–57. [Google Scholar] [CrossRef] [Green Version]
  271. Rossi, D.J.; Bryder, D.; Seita, J.; Nussenzweig, A.; Hoeijmakers, J.; Weissman, I.L. Deficiencies in DNA damage repair limit the function of haematopoietic stem cells with age. Nature 2007, 447, 725–729. [Google Scholar] [CrossRef]
  272. Mertens, J.; Paquola, A.C.; Ku, M.; Hatch, E.; Böhnke, L.; Ladjevardi, S.; McGrath, S.; Campbell, B.; Lee, H.; Herdy, J.R.; et al. Directly Reprogrammed Human Neurons Retain Aging-Associated Transcriptomic Signatures and Reveal Age-Related Nucleocytoplasmic Defects. Cell Stem Cell 2015, 17, 705–718. [Google Scholar] [CrossRef] [PubMed]
  273. Wahlestedt, M.; Norddahl, G.L.; Sten, G.; Ugale, A.; Frisk, M.-A.M.; Mattsson, R.; Deierborg, T.; Sigvardsson, M.; Bryder, D. An epigenetic component of hematopoietic stem cell aging amenable to reprogramming into a young state. Blood 2013, 121, 4257–4264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  274. Fatt, M.; Hsu, K.; He, L.; Wondisford, F.; Miller, F.D.; Kaplan, D.R.; Wang, J. Metformin Acts on Two Different Molecular Pathways to Enhance Adult Neural Precursor Proliferation/Self-Renewal and Differentiation. Stem Cell Rep. 2015, 5, 988–995. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  275. Khorraminejad-Shirazi, M.; Farahmandnia, M.; Kardeh, B.; Estedlal, A.; Kardeh, S.; Monabati, A. Aging and stem cell therapy: AMPK as an applicable pharmacological target for rejuvenation of aged stem cells and achieving higher efficacy in stem cell therapy. Hematol. Stem Cell Ther. 2017, 11, 189–194. [Google Scholar] [CrossRef]
  276. Safaeinejad, Z.; Nabiuni, M.; Peymani, M.; Ghaedi, K.; Nasr-Esfahani, M.-H.; Baharvand, H. Resveratrol promotes human embryonic stem cells self-renewal by targeting SIRT1-ERK signaling pathway. Eur. J. Cell Biol. 2017, 96, 665–672. [Google Scholar] [CrossRef]
  277. Candela, M.; Biagi, E.; Brigidi, P.; O’Toole, P.W.; De Vos, W.M. Maintenance of a healthy trajectory of the intestinal microbiome during aging: A dietary approach. Mech. Ageing Dev. 2014, 136-137, 70–75. [Google Scholar] [CrossRef]
  278. Saraswati, S.; Sitaraman, R. Aging and the human gut microbiota—From correlation to causality. Front. Microbiol. 2014, 5, 764. [Google Scholar]
  279. O’Toole, P.W.; Jeffery, I.B. Gut microbiota and aging. Science 2015, 350, 1214–1215. [Google Scholar] [CrossRef]
  280. Biagi, E.; Nylund, L.; Candela, M.; Ostan, R.; Bucci, L.; Pini, E.; Nikkïla, J.; Monti, D.; Satokari, R.; Franceschi, C.; et al. Through ageing, and beyond: Gut microbiota and inflammatory status in seniors and centenarians. PLoS ONE 2010, 5, e10667. [Google Scholar] [CrossRef]
  281. Pérez Martínez, G.; Bäuerl, C.; Collado, M.C. Understanding gut microbiota in elderly’s health will enable intervention through probiotics. Benef. Microbes 2014, 5, 235–246. [Google Scholar] [CrossRef]
  282. Lowry, C.A.; Smith, D.G.; Siebler, P.H.; Schmidt, D.; Stamper, C.E.; Hassell, J.E.; Yamashita, P.S.; Fox, J.H.; Reber, S.O.; Brenner, L.A.; et al. The Microbiota, Immunoregulation, and Mental Health: Implications for Public Health. Curr. Environ. Health Rep. 2016, 3, 270–286. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  283. Nehra, V.; Allen, J.M.; Mailing, L.J.; Kashyap, P.C.; Woods, J.A. Gut Microbiota: Modulation of Host Physiology in Obesity. Physiology 2016, 31, 327–335. [Google Scholar] [CrossRef] [PubMed]
  284. Zhang, X.; Zheng, X.; Yuan, Y. Treatment of insulin resistance: Straight from the gut. Drug Discov. Today 2016, 21, 1284–1290. [Google Scholar] [CrossRef]
  285. Rivière, A.; Selak, M.; Lantin, D.; Leroy, F.; De Vuyst, L. Bifidobacteria and Butyrate-Producing Colon Bacteria: Importance and Strategies for Their Stimulation in the Human Gut. Front. Microbiol. 2016, 7, 979. [Google Scholar] [CrossRef] [Green Version]
  286. Patel, R.; Dupont, H.L. New Approaches for Bacteriotherapy: Prebiotics, New-Generation Probiotics, and Synbiotics. Clin. Infect. Dis. 2015, 60, S108–S121. [Google Scholar] [CrossRef]
  287. Vaiserman, A.M.; Koliada, A.K.; Marotta, F. Gut microbiota: A player in aging and a target for anti-aging intervention. Ageing Res. Rev. 2017, 35, 36–45. [Google Scholar] [CrossRef]
  288. Wu, M.; Luo, Q.; Nie, R.; Yang, X.; Tang, Z.; Chen, H. Potential implications of polyphenols on aging considering oxidative stress, inflammation, autophagy, and gut microbiota. Crit. Rev. Food Sci. Nutr. 2020, 61, 2175–2193. [Google Scholar] [CrossRef] [PubMed]
  289. Du, Y.; Gao, Y.; Zeng, B.; Fan, X.; Yang, D.; Yang, M. Effects of anti-aging interventions on intestinal microbiota. Gut Microbes 2021, 13, 1994835. [Google Scholar] [CrossRef]
  290. Yoon, J.-H.; Abdelmohsen, K.; Kim, J.; Yang, X.; Martindale, J.L.; Tominaga-Yamanaka, K.; White, E.J.; Orjalo, A.V.; Rinn, J.L.; Kreft, S.G.; et al. Scaffold function of long non-coding RNA HOTAIR in protein ubiquitination. Nat. Commun. 2013, 4, 2939. [Google Scholar] [CrossRef] [Green Version]
  291. Singh, D.K.; Prasanth, K.V. Functional insights into the role of nuclear-retained long noncoding RNAs in gene expression control in mammalian cells. Chromosom. Res. 2013, 21, 695–711. [Google Scholar] [CrossRef] [Green Version]
  292. Yoon, J.-H.; Abdelmohsen, K.; Gorospe, M. Posttranscriptional Gene Regulation by Long Noncoding RNA. J. Mol. Biol. 2013, 425, 3723–3730. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  293. Grammatikakis, I.; Panda, A.C.; Abdelmohsen, K.; Gorospe, M. Long noncoding RNAs (lncRNAs) and the molecular hallmarks of aging. Aging 2014, 6, 992–1009. [Google Scholar] [CrossRef]
Figure 1. Hallmarks of aging. This scheme identifies the nine hallmarks briefly described in this review: loss of proteostasis, epigenetic alterations, telomere attrition, genomic instability, cellular senescence, mitochondrial dysfunction, deregulated nutrient sensing, altered intercellular communication, and stem cell exhaustion.
Figure 1. Hallmarks of aging. This scheme identifies the nine hallmarks briefly described in this review: loss of proteostasis, epigenetic alterations, telomere attrition, genomic instability, cellular senescence, mitochondrial dysfunction, deregulated nutrient sensing, altered intercellular communication, and stem cell exhaustion.
Biomedicines 10 02515 g001
Figure 2. Epigenetic Mechanisms via DNA Methylation and histone modification. Epigenetics can be altered via developmental mechanisms, environmental chemicals, pharmaceuticals, aging, and diet. Some dietary sources can lead to a direct production of DNA methylation, allowing for the overexpression or repression of genes, thereby increasing the aging process.
Figure 2. Epigenetic Mechanisms via DNA Methylation and histone modification. Epigenetics can be altered via developmental mechanisms, environmental chemicals, pharmaceuticals, aging, and diet. Some dietary sources can lead to a direct production of DNA methylation, allowing for the overexpression or repression of genes, thereby increasing the aging process.
Biomedicines 10 02515 g002
Figure 3. Mitophagy Modulators Against Mitochondrial Dysfunction. Several factors contribute to the aging process via the mitochondria, including mitophagy, mitochondrial fusion and fission, biogenesis, age-related diseases (i.e., AD and PD), mitochondrial dysfunction in general, and metabolic shifts. STACs, NMN, NR, and resveratrol are compounds that act against the aging process, directly targeting mitophagy and preventing the selective degradation of healthy mitochondria.
Figure 3. Mitophagy Modulators Against Mitochondrial Dysfunction. Several factors contribute to the aging process via the mitochondria, including mitophagy, mitochondrial fusion and fission, biogenesis, age-related diseases (i.e., AD and PD), mitochondrial dysfunction in general, and metabolic shifts. STACs, NMN, NR, and resveratrol are compounds that act against the aging process, directly targeting mitophagy and preventing the selective degradation of healthy mitochondria.
Biomedicines 10 02515 g003
Figure 4. Antiaging strategies to counter aging that might extend the human lifespan based on the use of chemical compounds and bioactive molecules and their use in various approaches mentioned in the figure.
Figure 4. Antiaging strategies to counter aging that might extend the human lifespan based on the use of chemical compounds and bioactive molecules and their use in various approaches mentioned in the figure.
Biomedicines 10 02515 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mishra, S.K.; Balendra, V.; Esposto, J.; Obaid, A.A.; Maccioni, R.B.; Jha, N.K.; Perry, G.; Moustafa, M.; Al-Shehri, M.; Singh, M.P.; et al. Therapeutic Antiaging Strategies. Biomedicines 2022, 10, 2515. https://doi.org/10.3390/biomedicines10102515

AMA Style

Mishra SK, Balendra V, Esposto J, Obaid AA, Maccioni RB, Jha NK, Perry G, Moustafa M, Al-Shehri M, Singh MP, et al. Therapeutic Antiaging Strategies. Biomedicines. 2022; 10(10):2515. https://doi.org/10.3390/biomedicines10102515

Chicago/Turabian Style

Mishra, Shailendra Kumar, Vyshnavy Balendra, Josephine Esposto, Ahmad A. Obaid, Ricardo B. Maccioni, Niraj Kumar Jha, George Perry, Mahmoud Moustafa, Mohammed Al-Shehri, Mahendra P. Singh, and et al. 2022. "Therapeutic Antiaging Strategies" Biomedicines 10, no. 10: 2515. https://doi.org/10.3390/biomedicines10102515

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

Article Metrics

Back to TopTop