Blog

Genetic and epigenetic regulation of human aging and longevity

Genetic and epigenetic regulation of human aging and longevity

Biochim Biophys Acta Mol Basis Dis. Author manuscript; available in PMC 2020 Jun 15.

Published in final edited form as:

Biochim Biophys Acta Mol Basis Dis. 2019 Jul 1; 1865(7): 1718–1744.

Published online 2018 Sep 1. doi: 10.1016/j.bbadis.2018.08.039

PMCID: PMC7295568

NIHMSID: NIHMS1593376

PMID: 31109447

Genetic and epigenetic regulation of human aging and longevity

Brian J. Morris,a,b,c,* Bradley J. Willcox,b,c and Timothy A. Donlonb,d

Author information Copyright and License information Disclaimer

The publisher's final edited version of this article is available at Biochim Biophys Acta Mol Basis Dis

Abstract

Here we summarize the latest data on genetic and epigenetic contributions to human aging and longevity. Whereas environmental and lifestyle factors are important at younger ages, the contribution of genetics appears more important in reaching extreme old age. Genome-wide studies have implicated ~57 gene loci in lifespan. Epigenomic changes during aging profoundly affect cellular function and stress resistance. Dysregulation of transcriptional and chromatin networks is likely a crucial component of aging. Large-scale bioinformatic analyses have revealed involvement of numerous interaction networks. As the young well-differentiated cell replicates into eventual senescence there is drift in the highly regulated chromatin marks towards an entropic middle-ground between repressed and active, such that genes that were previously inactive “leak”. There is a breakdown in chromatin connectivity such that topologically associated domains and their insulators weaken, and well-defined blocks of constitutive heterochromatin give way to generalized, senescence-associated heterochromatin, foci. Together, these phenomena contribute to aging.

Keywords: APOEFOXO3, Genome-wide association studies, Transcription, Epigenetics, Histones, Network analysis, Aging, Longevity

Go to:

1. Introduction

Heritability of human longevity has been estimated as 15–40% [1-9], although recent data from 5.3 million family trees of up to 13 million members generated from 86 million public profiles on an online genealogy database obtained an estimate of 16% [10]. The heritability of longevity may, however, depend on how extreme the survival probabilities used to define longevity are [11]. At younger ages, environmental factors such as infectious disease, rare conditions and externally inflicted trauma are the main causes of death. By old age (circa 70 years), individuals have partly escaped the most common causes of death in middle age, such as cancer and cardiovascular disease. Beyond age 70 the genetic component becomes increasingly important, influencing to a variable extent most common polygenic conditions that ramp up from middle-age onwards. In very old age (> 90 years) specific longevity genes emerge from the shadows and dominate over environmental influences in lifespan determination. Recent data from the Netherlands suggested paternal transmission of longevity is stronger than maternal transmission [12].

When considering the genetic basis of longevity, it is important to note that the genome is the “hardware” we are bom with. Our epigenome – chemical modifications to DNA and associated proteins – is the “software” influencing gene expression. Both are important. The epigenome is malleable and its composition can be influenced by environmental factors. Whereas there is virtually nothing one can do to favourably alter our genome, there is a keen interest in understanding factors, such as dietary components, that are able to modify our epigenome in order to establish a “healthy” transcriptome. There is also an increasing interest, and major investment by entrepreneurs, in developing drugs capable of affecting the epigenome in beneficial ways to slow, and even reverse, aging. Here we will review the current understanding of the genetic and epigenetic basis of aging and longevity with a focus on humans.

Go to:

2. Molecular genetic basis of longevity

2.1. Background

The modern study of the genetics of human longevity began with candidate gene studies based on major human physiological systems and diseases. The first such study focused on the human immune system and found that several human leukocyte antigen (HLA) polymorphisms were over-represented while other polymorphisms were under-represented in long-lived Okinawans (nonagenarians and centenarians) from the Okinawa Centenarian Study, versus younger controls [13]. Based on this initial discovery, the second such study utilized the same study design and focused on the human cardiovascular system. It found that apolipoprotein A, APOE, and angiotensin converting enzyme gene, ACE, polymorphisms differed between French and German centenarians versus younger individuals [14]. As findings from model organisms began to grow, and it became evident that evolutionarily conserved biological pathways existed that impact longevity, a pathway approach emerged. The next stage of human longevity studies therefore focused on the molecular genetic basis of longevity by testing polymorphisms in genes encoding proteins involved in pathways found in model organisms to affect lifespan. Such genes have roles in mitochondrial function, oxidative stress resistance, metabolism, DNA repair, control of the cell cycle, proteostasis, telomere shortening and other functions that might potentially affect the aging process [15-19]. Such case-control studies determine the genotype frequency of polymorphisms of potential candidate genes and look for alleles enriched in long-lived individuals. As is common for genetic studies of complex polygenic conditions, many of the initial candidate associations failed to replicate in other populations or racial groups [20-22]. Part of the problem is that a large number of genetic variants with small to moderate effects contribute to the longevity phenotype. Currently, the GenAge database lists over 300 human aging-related genes and the LongevityMap database of human genetic association studies contains over 500 entries [23]. Thus, validating a true candidate variant will require enormous statistical power [24] and thus very large cohorts of long-lived individuals, something that represents a challenge given the rarity of population prevalence of extremely old people. Nevertheless, prospects will improve as the number of centenarians is set to grow from the current 0.5 million worldwide to ~3.5 million by 2050 [25]. Besides additive genetic effects, it should be noted that non-linear epistatic interactions between variants, antagonistic pleiotropy and environmental interactions can confound attempts to replicate a finding [26].

In this section, we will start with the two most prominent candidate longevity genes to have been consistently replicated in multiple studies and later refer to some of the other candidates that have shown various degrees of promise. We will then move to genome-wide association studies (GWAS) that have the advantage of merely looking for loci showing possible linkage or association with longevity – further work then being required to identify which gene(s) and variant(s) at each locus is responsible.

2.2. TOMM40/APOE/APOC1 cluster

The apolipoprotein E gene (APOE) was first implicated in longevity in a 1994 study of centenarians in whom frequency of the ε4 allele was low and the ε2 allele was high [14]. Subsequent studies either confirmed or failed to replicate the finding (reviewed in: [27]). APOE encodes one of 3 common alleles, ε2, ε3 and ε4, produced from combinations of 2 non-synonymous SNPs, rs7412 (Arg158Cys: ε2 [28,29]) and rs429358 (Cys112Arg heterozygote ε3; Arg112 homozygote ε4 [30-32]). The ε4 variant (allele frequency 6–37% in different populations) provides an example of antagonistic pleiotropy in which benefit and risk depends on stage of life [33]. Antagonistic pleiotropy may be an evolutionary conserved principle of aging [34]. The ε4 allele has a reproductive and survival advantage at younger ages because of its association with higher female fertility and cognitive ability in both sexes in highly infectious equatorial environments [35,36]. In contrast, in modern post-industrial environments in which sanitation is good, the ε4 isoform is associated with elevated risk of aging-related diseases.

On average, individuals with ε4 have lower plasma apolipoprotein E and C-reactive protein (CRP), accompanied by higher plasma cholesterol, low density lipoprotein cholesterol (LDL-C), apolipoprotein B, lipoprotein(a), atherosclerosis and body mass index [37,38]. This puts ε4 carriers at increased risk of cardiovascular disease [39]. A meta-analysis found that compared with the wild-type ε3/ε3 genotype, genotypes ε3/ε4 and ε4/ε4 were associated with increased risk of coronary heart disease of 22% and 45%, respectively, whereas in Caucasians only, the ε2 allele was associated with a 16% decrease in risk [40]. The ε4 allele has long been associated with increased risk of Alzheimer's disease [41]. Besides amyloid-β plaques and tau protein tangles [42], cardiovascular risk factors may be at play in this condition, owing to microhemorrhages in cerebral blood vessels, leading to damage to surrounding neural tissue that eventually presents as Alzheimer's disease in the elderly [43].

Some researchers suggest that data indicate that rather than being a longevity gene, APOE is in fact a “frailty gene” [44]. Over time, greater relative attrition of carriers of the ε4 allele as a consequence of cardiovascular mortality in middle to early old age, would explain the higher prevalence of ε2 carriers in very old age [44]. Japanese aged ≥105 years exhibit extremely low frequency of APOE-ε4 alleles [45]. A study of Canadians aged 85 years who had never been diagnosed with cardiovascular disease, dementia, diabetes, cancer or major pulmonary disease found reduced prevalence of the APOE ε4 allele in comparison with random midlife controls [46].

Network analysis suggested age-related changes in lipid and cholesterol maintenance, particularly in the brain, may be central to healthy aging and longevity [46]. Racial differences were apparent, there being no association of any SNP with longevity in African American women [47]. A meta-analysis of data from different racial groups found a non-significantly higher association of ε2/ε2 compared with ε3/ε3 genotype with longevity [48]. In a subsequent rigorous meta-analysis, the ε2/ε2 association with longevity was significant in Europeans (OR 2.39; 95% CI 0.99, 5.76), being strongest in southern European populations [49]. Variation by ethnicity was suggested as indicating gene x environment or gene x gene interaction effects.

APOE is in the 20-kb TOMM40 (translocase of outer mitochondrial membrane 40 gene), APOE, APOC1 (apolipoprotein C1 gene) cluster – TOMM40/APOE/APOC1. All genes in this cluster are influenced as a group. The variant responsible for contrasting APOE phenotypes has been suggested to be the G-allele of SNP rs2075650 located in the promoter of TOMM40, which is positioned upstream of APOE and APOC1 [50,51]. Via moderate linkage disequilibrium with the APOE ε4-defining SNP rs429358, the TOMM40 SNP tags the deleterious effects of the ApoE ε4 isoform [52]. The TOMM4/APOE/APOC1 locus is, however, genetically complex, with multiple TOMM4/APOE/APOC1 locus cis-elements influencing both APOE and TOMM40 promoter activity according on haplotype and cell type [53]. The G-allele may be tagging various other underlying causal variants with different effects on CAD risk and CRP [54]. A causal relationship between CRP and cardiovascular risk has been questioned [55]. APOE ε4 carriers showed greater cortical thinning [56]. Increasing poly-T lengths of TOMM40 are associated with hippocampal thinning only in subjects lacking the APOE ε4 allele [57]. In a nutshell, the APOE ε2/ε3/ε4 story remains a work in progress.

2.3. FOXO3

The forkhead/winged helix box, group O (FoxO) transcription factors are crucial component(s) of the insulin/insulin-like growth factor (IGF-1) signaling (IIS) pathway. Binding of insulin and IGF-1 to their respective receptors results in activation of adenosine monophosphate-activated protein kinase (AMPK), affecting various signaling networks involved in maintaining cell metabolism in response to a reduction in cellular energy reserves, autophagy, intracellular lipid metabolism, mitochondrial function and aging [58,59]. The IIS pathway was first implicated in longevity when mutations in the Caenorhabditis elegans insulin/IGF-1 receptor gene (daf-2) were found to double lifespan [60]. Replication of this effect in other organisms, as well as of mutation of other genes in the IIS pathway, led to the suggestion that FoxO transcription factor(s) might be crucial to longevity [61]. This was because FoxO proteins regulate expression of a vast array of genes involved in energy metabolism, oxidative stress, apoptosis, G0 to G1 and G1 to S phase progression in the cell cycle, and a diversity of other metabolic processes [61], doing so by dampening the negative effects of the IIS pathway on lifespan [62-64]. Mammals have 4 FoxO proteins – FoxO1, FoxO3, FoxO4 and FoxO6 [65].

Our group was the first to show an association of SNPs in the gene FOXO3 with longevity in a study of American men of Japanese ancestry aged ≥95 years [66]. Since that finding, over a dozen independent studies in geographically and racially divergent populations worldwide have confirmed the FOXO3 finding, with this gene being the next best-replicated (after APOE) longevity gene association [67-74]. A meta-analysis in 2014 of 11 of the studies found 5 SNPs to be associated with longevity, the strongest SNP being rs2802292 (odds ratio [OR] = 1.36; p = 0.005), and which was male-specific [75]. A stratified analysis in 2015 suggested that this SNP might contribute more to longevity in Asians than Europeans [76]. In 2016, male-specific association of other FOXO3 SNPs with longevity was seen in US [77] and Chinese [78] cohorts. In 2017, a more stringent meta-analysis confined to individuals of white ethnicity found that the genetic effect of SNP rs2802292 decreases as more extreme definitions of longevity were considered [49]. Sebastiani and coworkers argued that FOXO3 alleles may be associated with longevity, but not extreme longevity, at least in white males. This was supported by data from a Danish study [79]. In Japanese males, however, we found that the longevity-associated (G) allele of SNP rs2802292 continued to increase in frequency beyond the age of 105 years [66].

In a recent study of 110 SNPs in FOXO3 and 5 kb of its flanking DNA, we found 41 were associated with longevity in American men of Japanese ancestry [80]. Subsequently 17 of 107 FOXO3 SNPs were associated with longevity in 4 European and US cohorts and allelic association with elevated hippocampal expression was seen [73]. In our study of Japanese American men, nucleotide changes in 13 of the 41 SNPs disrupted the binding sites of 18 transcription factors [80]. Two other intronic SNPs strongly associated with longevity in German, French and Danish populations exhibited allele-specific binding of CCCTC-binding factor (CTCF) and serum response factor (SRF), leading to increased FOXO3 expression of the longevity-associated allele of each in luciferase reporter gene assays involving various human tissues [81]. It was reported recently that the transcription factor heat shock factor 1 (HSF1) binds to the enhancer sequence created by the G allele of rs2802292 in FOXO3 intron 2, so explaining why, by conferring increased resilience to stress, this SNP is so strongly associated with longevity [82].

To determine the cause(s) of mortality that FOXO3 protects against we performed a study involving rs2802292 and found that FOXO3's association with longevity was principally via reduction in risk of mortality from coronary artery disease (CAD) [83,84], although with a larger sample other causes may also become evident. FOXO3 SNP rs2802292 was associated with self-rated health in individuals aged 75–87 years, this being influenced by cardiovascular disease reduction, but not mental or cognitive status [85]. No associations were found for SNPs in APOE and TOMM40 in that study [85]. The effect of FOXO3 risk alleles on cardiovascular disease earlier in life might help explain an apparent association with vascular factors and Alzheimer's disease [86]. We found recently that the protective G-allele of rs2802292 was associated with negligible telomere attrition with age in our Okinawan study population, consistent with a contribution of telomere dynamics to longevity [87]. The FOXO3 longevity-associated intronic variant, rs2490272, was the strongest SNP of variants in 52 genes associated positively with intelligence in a meta-analysis of GWAS findings for white individuals of European descent [88].

FOXO1 SNPs were associated with longevity in Han Chinese centenarians [71], but no association was seen in Japanese American men [66].

2.4. Genome-wide association studies

Both the TOMM4/APOE/APOC1 locus [52,89-98] and the FOXO3 locus [74,99] have been validated for longevity based on their loci exceeding the p < 5 × 10−8 threshold for genome-wide significance in large GWAS (Table 1). GWAS of case-control cohorts and sibpair genome-wide linkage studies have also identified other loci exhibiting genome-wide significance for longevity (Table 1).

Abbreviations in column 2: CAMP, Collaborative Aging and Memory Project; CHARGE, Cohorts for Heart and Aging Research in Genomic Epidemiology Consortium; CLHLS, Chinese Longitudinal Health Longevity Survey; GEHA, Genetics of Healthy Aging in Europe; HRS, Health and Retirement Study; LLFS, Long Life Family Study; LLS, Leiden Longevity Study; MCC, multiple combined cohorts; NECS, New England Centenarian Study; UKB, UK Biobank. Other abbreviations: Sibs, number of siblings present in study; Rep, number of replication samples in stidy; Ped, pedigrees; M, found only in male subjects; F, found only in female subjects.

The respective MCC indicated:

aBelfast Elderly Longitudinal Free-Living Aging Study, Calabria cohort, CEPH centenarian cohort, Chinese Longitudinal Healthy Longevity Surveys, Danish longevity study I and II, deCODE, Estonian Biobank, Genetics of Healthy Aging Study, German longevity study, Leiden 85-plus study, Newcastle 85 + Study, PROspective Study of Pravastatin in the Elderly at Risk, Rotterdam Study, TwinGene

bNew England Centenarian Study, 90PLUS Cohort.

cSouthern Italian Centenarian Study, Long Life Family Study, Longevity Gene Project, New England Centenarian Study. cSouthern Italian Centenarian Study, Long Life Family Study, Longevity Gene Project, New England Centenarian Study.

d35 other variants were genome-wide significant; rs1051730 deemed most significant by the authors owing to its previous link to lower smoking and lung cancer risk.

If two genes are separated by dots the SNP is intergenic.

Parts of this Table were adapted from Hook et al. [109].

 

SNPs at candidate gene loci shown in Table 1 included the elongation of very long chain fatty acids protein 6 gene, ELOVL6 [97], the intergenic region between cadherin gene, CLESR2, and the proline and serine rich coiled coil 1 gene, PSRC1 [98], the intergenic region between human leukocyte antigene genes, HLA-DRB1 and HLA-DQA1 [74,98], BEN domain-containing 3 gene, BEND3 [98], epoxide hydrolase 2 gene, EPHX2 [98], prospero homeobox 2 gene, PROX2 [98], melanocortin 2 receptor (adrenocorticotropic hormone receptor) gene, MC2R [98], fucose-1-phosphate guanylyltransferase gene, FPGT/TNN13K [98], ubiquitin specific peptidase 2 (USP2) antisense RNA 1 gene, USP2-AS1 [98], psoriasis susceptibility 1 candidate 3 (non-protein coding) gene, PSORS1C3 [98], thymocyte selection-associated HMG box gene, TOX [98], beta 3-glucosyltransferase gene, B3GALTL (now B3GLCT) [98], major histocompatibility complex (MHC) class I polypeptide-related sequence A gene, MICA, and B gene, MICB [98], zeste-white 10 gene, ZW10 [98], semaphoring 6D gene, SEMA6D [98], intergenic region between egl-9 family hypoxia inducible factor 2 gene, EGLN2, and cytochrome P450 2A6 gene, CYP2A6 [98], exocyst complex component 3-like 2 gene, EXOC3L2 [98], microtubule affinity-regulating kinase 4 gene, MARK4 [98], DNA topoisomerase gene, TOP2B [101], toll-like receptor 4 gene, TLR4 [101], deleted in bladder cancer 1 gene, DBC1 [101], bone morphogenetic protein 1 and 2 genes, BMP4 and BMP5 [103], plasminogen, PLG/mitogen-activated protein kinase kinase kinase 4 gene, MAP3K4 [104], parkin gene, PRKN [104], early B cell factor 1 gene, EBF1 [92], cholinergic receptor nicotinic α3 subunit gene, CHRNA3 [105], CHRNA4 [98], mothers against decapentaplegic homolog 7 gene, SMAD7 [108], interleukin-6 gene, IL6 [106], ankyrin repeat domain 20 family member A9 pseudogene, ANKRD20A9P [106], lipoprotein(a) gene, LPA [96,98], cyclin-dependent kinase 4 inhibitor B anti-sense gene, CDKN2B-AS (also known as ANRIL, a long noncoding [Inc] RNA) [96,98], neuronal acetyl choline receptor subunit α-5 gene, CHRNA5 [96], CHRNA3 [98], the subtilisin-like proprotein convertase, furin gene, FURIN [98], ubiquitin-specific peptidase 42 gene, USP42 [24], transmembrane and TPR repeat-containing protein 2 gene, TMTC2 [24], neuroblastoma breakpoint family 5 and 6 genes, NBPF5 and NBPF6 [100], and capsid protein 9 gene, CAP9 [100]. Others, that have exhibited statistical significance but were unvalidated, were glutamate ionotropic receptor kainite type subunit 2 gene, GRIK2 [93], inflammation and DNA repair protein gene locus, RAD50/IL13 [110], and multiple polyphosphate polyphosphatase 1 gene, MINPP1 [111].

A meta-analysis of 4 GWAS of 2086 cases in which longevity was defined as the oldest 1% of the 1900 birth year cohort identified 37 SNPs that attained genome-wide significance (p < 5E–8) [97]. Instead of using as cases the limited numbers posed by obtaining very old individuals, one of the GWAS ramped up the n value of the data set enormously by taking SNPs associated with aging-related disease and adjusting these for longevity effects in order to detect SNPs associated with longevity [96]. This led to 16 genome-wide significant SNPs, 11 of which were then validated in 5 independent population cohorts [96].

A GWAS has also been conducted in offspring of long-lived parents, finding genome-wide significance for the SMAD7 locus on chromosome 18 [108]. Of 374 SNPs within 50 kb of 6 prior longevity loci, including TOMM4/APOE/APOC1 and FOXO3, only a SNP in the early B-cell factor gene, EBF1, region approached significance [108].

Genome-wide exon sequencing of white individuals aged 98–108 years and controls did not find amino acid mutations associated with exceptional lifespan [112]. A non-significant increase in variant burden was noted for lysosomal trafficking regulator gene, LYST, midasin AAA ATPase 1 gene, MDN1, and RNA-binding motif protein, X-linked like 1 gene, RBMXL1. A GWAS of copy number variants (CNV) in Danish nonagenarians and centenarians found a significant increase in mortality for every 10 kb increase in average CVN length [113]. A study of 21 hypervariable short tandem repeats spread throughout the genome found 6 that were associated with longevity [114].

Using data on genes and pathways associated with five major age-related diseases from 410 GWAS, conserved pathways of aging – in particular apolipoprotein metabolism genes – were found to simultaneously influence multiple age-related diseases [115]. Shared gene ontology terms included nutrient-sensing, signaling, translation, proteostasis, stress response and genome maintenance. Balanced proteostasis is critical for maintaining functional proteins, an adequate stress response is critical for reducing reactive molecules (e.g., superoxide) and their by-products, and genome stability is essential from a mutation load perspective as well for maintaining proper tissue-specific gene regulatory control. Table 2 lists the functions of the genes that reached genome-wide significance in GWAS.

Genetic correlations between cognitive ability and longevity have been noted [138]. A meta-analysis of education-associated variants from a GWAS [139] in 3 large UK and Estonian cohorts showed that a 1 SD higher polygenic education score was associated with a 2.7% and 2.4% lower risk of mortality for mothers and fathers of the subjects; parents in the upper third of the score lived 0.55 years longer than parents of offspring in the lower third [140].

Numerous suggestive loci for longevity have also been found in GWAS and genome-wide linkage studies. The number of these in respective studies was as follows: 2913 [105], 44 [97], 20 [93], 12 [96], 11 [100], 9 [106], and 6 [92]. Among these were loci that attained genome-wide significance in other longevity GWAS.

A large-scale RNA-sequencing-based expression quantitative trait locus (eQTL) study of blood from German and Danish subjects of different ages up to 104 years suggested that longevity-associated biological processes such as altered metabolism are, at least in part, a driving force of longevity, not just a consequence of old age [141]. Lack of replication in different populations may be due to variation in specific population genetic structures [26,142], between-population differences in linkage disequilibrium levels between two causal SNP loci [142], involvement of different genes in similar biological processes [143], and the large influence of non-genetic factors [141]. No biological processes were independent of a nongenetic contribution [141]. Prominent in attainment of longevity might be caloric restriction (CR) and traditional healthy dietary intake [144,145].

A number of GWAS have used mice to validate human longevity gene findings. For example, in the case of RNA binding motif protein 6 gene, RBM6, the mouse ortholog showed an association of lower Rbm6 expression in the prefrontal cortex with longer lifespan, and for sulfotransferase family 1A member 1 gene, SULT1A1, increased expression of Sult1a1was seen in CR mice [96].

Finally, an 8 million-SNP GWAS involving 2,9693 elderly Dutch Europeans, with replication in other Dutch and in UK cohorts, found perceived facial age was most strongly associated with multiple SNPs in the melanocortin receptor 1 gene, MCR1 [146].

2.5. Case-control studies for longevity gene discovery

2.5.1. Other IIS pathway genes 

Disruption in model organisms of genes in the IIS pathway can lead to an up to 2-fold increase lifespan [15,109,147]. A genome-wide meta-analysis identified 7 genome-wide significant loci, including FOXO3, associated with circulating IGF-1 and 4 associated with IGFBP-3 concentrations [148]. An eQTL allele of the transcriptional regulator additional sex combs 2 gene, ASXL2, involved in development, was associated with reduced IGF-1, lower adiposity and longevity [148]. A study of tagging SNPs in other genes in the IIS pathway, besides those for FoxOs, in American men of Japanese ancestry failed to find longevity association for activating transcription factor gene, ATF4, E3 ubiquitin protein ligase Cbl proto-oncogene, CBL, cyclin-dependent kinase inhibitor 2B gene, CDKN2B, exonuclease 1 gene, EXO1 and c-jun proto-oncogene (AP-1 transcription factor subunit), JUN [149].

SNPs in GH/IGF-1/insulin signaling-associated genes influence both longevity and height. In a Japanese study, height-increasing genetic scores based on 30 SNPs were significantly associated with height in controls and inversely with extreme lifespan in women, but not men [150]. A study involving 4 white populations found a significant association of a GH receptor exon 3 deletion variant that increases GH sensitivity with male, but not female, longevity [151]. Homozygotes were 2.5 cm taller and lived 10 years longer. In a study of American men of Japanese ancestry an inverse association of the major longevity-associated FOXO3 SNP, rs2802292, with height was found [152].

2.5.2. TORC genes 

The mechanistic target of rapamycin (mTOR) is a serine/threonine kinase that forms a component of mTOR complex 1 (TORC1) and TORC2 [153]. TORC1 activates ribosomal protein S6 kinase (RPS6KA1) and inhibits eukaryotic translation initiation factor 4E-binding protein 1 (EIF4EBP1), which results in increased mRNA translation and thus protein synthesis [153]. In studies of model organisms, treatment with the mTOR inhibitor rapamycin extends their lifespan [132]. The longevity effect of CR is associated with reduced TORC1 activity [15]. Manipulation of TORC pathway genes is able to modulate lifespan [109]. A case-control study in humans of TORC components failed to find any association with longevity [154]. This involved genotyping 6 tagSNPs in mTOR gene, MTOR, 61 in regulatory-associated protein of mTOR gene, RPTOR, 7 in rapamycin-insensitive companion of mTOR gene, RICTOR, and 5 in ribosomal protein S6 kinase A1 gene, RPS6KA1 [154]. Genotyping of additional SNPs in RPTOR revealed a marginal association with longevity [155].

2.5.3. Sirtuins 

This family, comprising 7 members, is involved in epigenetic modification of multiple intracellular substrates in a NAD+-dependent manner [17]. Sirtuins respond to CR by regulating a wide range of cellular functions that include metabolic and neuronal pathways [17]. The best characterized, sirtuin 1, controls mitochondrial function by deacetylation of FoxOs, transformation-related protein 53 (Trp53), peroxisome proliferator-activated receptor gamma coactivator 1-α (PGC-1α) and others [156]. Increased sirtuin activity helps lower aging-related metabolic dysfunction and cancer risk. While sirtuins 1, 2, 6 and 7 operate in the nucleus, sirtuin 3, 4 and 5 are located in mitochondria and affect mitochondrial proteins. Genetic variation in SIRT1 has not been associated with longevity [157-159], whereas variants in SIRT2 [160], SIRT3 [72,161,162], SIRT4 [155], SIRT5 [72,155], SIRT6 [72,163,164] and SIRT7 [155] have shown associations with longevity.

2.5.4. Telomere genes 

Telomere attrition with age has been implicated in cell senescence and lifespan [165,166]. Most individuals do not reach a “telomeric brink” of 5 kb, denoting high risk of imminent death, during their lifetime, although this phenomenon is more likely to be seen in the oldest old [167]. A common haplotype of telomerase reverse transcriptase gene, TERT, has shown an association with longevity in Ashkenazi centenarians and their offspring [168]. Others found no association for TERT SNPs, whereas a SNP in telomerase RNA component gene, TERC, was associated with longevity [169]. A variant in the 3′UTR of the oligonucleotide/oligosaccharide-binding folds containing 1 gene, OBFC1, whose encoded protein is involved telomere maintenance, was associated with longevity [77].

2.5.5. Genes differentially expressed during CR 

Among genes differentially expressed in response to CR in mice [170], most would have been responding to the CR state. Nevertheless, the possibility that the human homologs could be potential candidates for human longevity led us to perform a case-control study in American men of Japanese ancestry. Two of 12 tagging SNPs in connective tissue growth factor gene, CTGF, and 7 of 41 tagSNPs in epidermal growth factor gene, EGFR, were associated with longevity [171]. A Korean study found haplotypes of EGFR SNPs were associated with longevity in women [172]. Of 459 tagSNPs in 47 genes, as well as 12 other genes, association with longevity was found, after correction for multiple testing, for tagSNPs in mitogen-activated protein kinase kinase kinase 5 gene, MAP3K5, phosphoinositide 3-kinase regulatory subunit 1 gene, PIK3R1, Fms related tyrosine kinase 1 (vascular endothelial growth factor receptor) gene, FLT1, sirtuin 5 gene, SIRT5, and sirtuin 7 gene, SIRT7 [155]. Some of the SNPs in these genes had more significant associations when assuming models of inheritance other than “minor allele recessive”, e.g., “heterozygote advantage”, which could support the concept of “antagonistic pleiotropy”.

2.5.6. Other genes 

Targeted disruption of the angiotensin II type 1 receptor gene, AGTR1, increases mouse lifespan markedly [173]. Allelic variants of promoter SNPs were associated with lower blood pressure and extreme old age in Italian and Japanese centenarians, thus implicating AGTR1 as a longevity gene [174]. A study in 1994 [14] examined another gene in the renin-angiotensin system, ACE, in which the deletion (D) of an insertion/deletion (I/D) polymorphism had been implicated in myocardial infarction [175] and death in subjects at elevated risk of myocardial infarction by having early onet severe essential hypertension [176]. The association of the D allele with longevity in French subjects [14] was then replicated in British [177], Italian [178], Portugese [179] and Uygur Chinese [180,181] centenarians, but not in many other studies. A meta-analysis found significantly higher DD prevalence in centenarians (OR 1.16; 95% CI 1.05–1.28; p < 0.001) [182]. The contrasting association could represent another example of antagonistic pleiotropy. In patients with Alzheimer's disease, those with the insertion (I) allele of ACE were less likely to develop cerebral white matter changes [183].

The minor allele of a longevity-associated SNP, rs2149954, at chromosome 5q33.3, identified in a GWAS [92], was associated with lower hypertension prevalence, decreased risk of myocardial infarction and heart failure, as well as increased physical functioning in long-lived individuals [184]. Genetic variation in the human bactericidal/permeability-increasing fold-containing family member 4 gene, BPIFB4, impairs endothelial nitric oxide synthase activity, so reducing vasorelaxation and increasing diastolic blood pressure [185]. Under a recessive genetic model an Ile229Val polymorphism of BPIFB4, by modulating endothelial function and angiogenesis, was associated with longevity [186]. When overexpressed in hypertensive rats and old mice, the longevity-associated of BPIFB4 protein variant reduced blood pressure, rescued endothelial dysfunction and promoted vascular repair processes [186]. Serum BPIFB4 protein levels were higher in healthy centenarians, but lower in frail centenarians [187], as was the case for BPIFB4 mRNA [188]. The opposite was seen for the ischemia-responding HIF-1α chemokine receptor CXCR4 mRNA [188].

Reduction in frailty is associated with a G–395A promoter polymorphism in KLOTHO, a well-known longevity and aging-suppressor gene [189]. Genetic variation in TXNRD1, the gene encoding thioredoxin reductase, which protects against accumulation of reactive oxidants, modulates physical decline in extreme old age [190]. In other genes, variation in superoxide dismutase 3 gene, SOD3 [72], AKT serine/threonine kinase 1 gene, AKT1 [72] and mitochondrial DNA polymorphisms [191-195] have shown associations with longevity in various populations. A meta-analysis of multiple mitochondrial DNA association studies revealed a significant impact of Caucasian haplogroups H, J and K on longevity, type 2 diabetes, dementia and cancer [99]. Centenarians have a relatively high mitochondrial copy number, and this is evident in their F1 offspring [196]. This was likely mediated by single-stranded DNA-binding protein 4, which was highly expressed in centenarians and offspring, and was significantly associated with DNA copy number [196]. This results in adequate maintenance of energy supply.

The cholesterol ester transfer protein gene, CETP, whose product affects HDL metabolism, was associated with longevity in Ashkenazi Jewish and Danish [197,198], as well as Chinese [199], populations, but not in Greeks [200], American whites [20], and Italians [201]. Genetic variation in syndecan-4 gene, SDC4, whose encoded protein is a central mediator of cell adhesion, was associated with lipid profile and longevity [202].

Interleukin-10 modifies the inflammatory response and an IL10 promoter polymorphism associated with elevated IL-10 production, was associated with longevity in Japanese [203], Italian [204] and Jordanian [205] men, and in both sexes in Bulgaria [206]. Recently, multiple SNPs in the major histocompatibility complex, class II, including the DQ β1 gene, HLA-DQB1, have shown associations with longevity at the p = 10−8 level [207]. Interestingly, HLA genes were the first reported longevity-associated genes in humans [13]. Longevity-associated alleles in 2 SNPs were associated with plasma triglycerides and LDL:HDL ratio. A GWAS of self-rated health involving the UK Biobank found association with 13 independent loci [208]. Most prominent of these were HLA-DQB1, as well as HLA-DQA1, HLA-DRA1 and HLA-DDRB5, and Kruppel-like factor 7 gene, KLF7. The tumor necrosis factor (TNF) superfamily of genes is located in the central class II HLA region. TNF-α is pro-inflammatory, which is opposite to IL-10 which is anti-inflammatory. TNFA variants are not associated with longevity. Targeting of known immune-associated loci using an immunochip identified a novel SNP in the extended region of the inflammation and DNA repair gene locus, RAD50/IL13, with longevity in German, French and Danish long-lived individuals (OR 1.20; 95% CI 1.12–1.28; p = 5.4 × 10−7) [110].

Variation in genes involved in DNA repair, such as LMNA [209], WRN, CDKN2A and CDKN2B [74,210], as well as FOXO3 (discussed earlier), are associated with longevity. Longevity associations were reported for haplotypes of aldehyde dehydrogenase 2 gene, ALDH2, proprotein convertase subtilisin/kexin type 1 gene, PCSK1, V-yes-1 Yamaguchi sarcoma viral related oncogene homolog gene, LYN, and several other genes in Korean nonagenarians [172]. It is often difficult to understand or predict how genes such as these might influence longevity, since all of the pathways a gene influences may not be currently known.

Antagonistic pleiotropy may apply to the spermatogenesis associated 31 gene, SPATA31 [211], which belongs to the core duplicon families important in hominid evolution and is one of the fastest evolving in human evolution [212]. Increased copy number likely helped in protection from UV damage when hominids became diurnal and lost body hair, but at the cost of activation of senescence pathways and DNA repair processes risking elevation in somatic mutations and cancer. While SPATA31 overexpression causes premature senescence by interfering with aging-related transcription pathways, fewer copies are seen in Germans aged > 96 years as compared with younger individuals, consistent with a fitness benefit during the reproductive period of life, but a negative influence on lifespan [211].

An intronic variant of the solute carrier family 1, member 5 gene, SLC1A5, encoding the major glutamine transporter, ASCT2, influences longevity via an effect on splicing [213]. Amino acid transporter gene SNPs have also been shown to be associated with age-related physical decline and survival in the elderly, likely via TORC1 signaling [214].

Longevity associations seen for SIRT2 involved SNPs in microRNA (miRNA) target sites in the 3′untranslated region (3′-UTR), as was also the case for dopamine receptor 2 gene, DRD2 [160]. Disruption of miRNA binding to the 3′-UTR often alters mRNA stability and thus levels of encoded protein synthesized. A rare loss-of-function mutation in the serpin family E member 1 gene, SERPINE1 (c.699–700dupTA), which encodes plasminogen activator inhibitor-1, was associated with longevity in an Amish community, as well as longer telomere length, lower fasting glucose and less type 2 diabetes [215]. Evidence of epistasis exists. For example, an association of fibronectin type III domain-containing 5 gene, FDNC5, with longevity appears to depend on the presence of the FOXO3 rs2802292 T allele and APOE ε2/ε4 [216]. A weak association was found between polymorphisms in the vitamin D receptor gene, VDR, and longevity, as well as health parameters, in centenarians [217].

SNP-SNP interaction analyses have been proposed recently for investigation of the genetics of human longevity [218]. Synergistic interaction by SNPsyn was applied to 3 candidate pathways – IIS, DNA repair, and pro/antioxidant – using 1058 tagSNPs in 140 genes and subjects from the Danish 1905 Birth Cohort Study. The strongest involved specific SNPs in the IGF receptor 1 gene, IGF1R, the tyrosine-protein phosphatase non-receptor type 1, PTPN1 (otherwise known as the proteintyrosine phosphatase 1B gene, PTP1B), DNA repair pathway gene, TP53, and the excision repair cross complementing gene, ERCC. Epistatic interactions were seen between TP53 and pro-oxidant pathways gene TXNRD1, as well as between TP53 with ERCC2, another gene involved in the DNA repair pathway. The growth hormone secretagogue receptor gene, GHSR, interacted with IIS and DNA repair partner genes, pregnancy-associated plasma protein A gene, PAPPA, protein tyrosone phosphatase, non-receptor type 1 gene, PTPN1, Parkinsonism associated deglycase gene, PARK7, and meiotic recombination 11 homolog gene, MRE11A, involved in homologous recombination, telomere length maintenance, and DNA double-strand break repair.

2.5.7. The “gerontome” 

In a large systems-level analysis of the genetics of aging, and that discriminated between pro- and anti-longevity genes, it was shown that genetic links between aging and aging-related diseases appeared to be due to a small fraction of aging-related genes that tend to have a high network connectivity and that aging-related disease genes have faster molecular evolution rates [219]. Aging genes tend to be at network hubs, where, via protein-protein interactions and co-expression networks, communication takes place among different functional modules or pathways, indicating high connectivity [220] (Fig. 1). Close interactions among aging hubs may make aging subnetworks vulnerable, so contributing to aging. There were strong interactions between aging genes through biological sub-networks such that aging genes weremore likely to collaborate with one another than with background, essential, transcription factor, and housekeeping genes (Fig. 2).

 

Fig. 1.

Aging-related protein interaction subnetwork for three well-known aging-related pathways: insulin signaling, AMPK and mTOR signaling. (From Zhang et al. [220])

 

Fig. 2.

The aging subnetwork consists of 192 aging genes and 561 direct interactions among these. The sizes of the nodes are proportional to their degrees of interaction in the entire protein-protein interaction network. Aging genes involved in aging-related pathways and interactions among them are both highlighted. (From Zhang et al. [220])

2.6. Gene neighborhoods in longevity

We have found recently that FOXO3 is located at the hub of an early-replicating neighborhood of 46 genes with which it interacts via CTCF, a transcription factor that regulates chromatin architecture, attracting tissue-specific transcriptional activators, repressors, cohesin and RNA polymerase II [80]. The 46 neighboring genes encode proteins that, like FoxO3, are involved in various processes that contribute to cell resilience, such as autophagy, stress response, energy/nutrient sensing, cell proliferation, apoptosis and stem cell maintenance. Together they may work as a “gene factory” for healthy aging. In response to stress, we showed that FOXO3 moved physically closer to its neighboring genes, the shift being stronger for carriers of the longevity-associated G-allele of FOXO3 SNP rs2802292 compared with homozygotes of the common, T, allele (Fig. 3) [80]. FOXO3 mRNA expression in response to stress was more pronounced for FOXO3 G-allele carriers, as were two other genes in the cluster so far tested. These findings highlight the fact that genotype-phenotype correlations commonly reported in the study of complex traits often focus on single protein-coding genes but ignore gene neighborhoods. It has been suggested [3] that physical interactions between genes themselves might be an additional contributory factor in the omnigenic model proposed recently to explain the “missing heritability” evident from large-scale genome-wide association studies of complex polygenic traits [4]. Confirmation of this will require further research.

 

Fig. 3.

The gene FOXO3 interacts with its neighbors in a 46-gene cell resilience “gene factory” on chromosome 6q21 [80]. Upper panel fluorescent in situ hybridization experiments showing, on the left, position of fluores-cently-labeled FOXO3 (pale blue), HACE1 (green) and LAMA4 (red) in quiescent lymphoblastoid cell lines; and on the right, change in position of the genes in cells after activation by stress, induced by serum deprivation and H2O2 treatment. Lower panel schematics showing the effect; for simplicity only 5 of the 46 neighborhood genes are shown. The sphere denotes a presumed transcription center. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

It would now appear that modulation of FOXO3 activity could have an amplifier effect on genes in its neighborhood. This would complement the transcriptional effects that FoxO3 has on expression of a wide array of specific genes across the genome. Other gene clusters, such as the TOMM4/APOE/APOC1 locus merit similar scrutiny for gene-gene interactions in complex polygenic conditions, including longevity.

2.7. Genes for healthy aging

Healthy aging appears to be a phenotype distinct from exceptional longevity. This was the finding of a whole-genome sequencing study of 1354 individuals in the UK aged 80–105 years reporting no chronic diseases, and termed the “wellderly”, compared with a control group [221]. No locus reached genome-wide significance, but among the top hits were variants associated with cognitive function that were significantly over-represented, most notably in a linkage block at the MHC locus 6p22.1 (p = 6 × 10−7), followed by 5q31.1 containing solute carrier family 22 member 4 gene, SLC22A4, that affects carnitine and carnitine-related metabolite levels, then 2q36.1, containing potassium voltage-gated channel subfamily E regulatory subunit 4 gene, KCNE4, that has a nearby SNP affecting cognitive decline. No association could be found with well-known longevity loci in this cohort, although borderline significance with healthy aging was seen for the major longevity-associated FOXO3 SNP rs2802292 and there was depletion of the APOE-ε4 frailty allele in the “wellderly”. A decreased genetic risk score for Alzheimer's disease and CAD was seen, but not for type 2 diabetes or cancer. The authors suggested that healthy lifestyle might be the main driver of familial clustering of exceptional good health. This factor would help propel individuals into achieving extreme old age.

Variability in red blood cell (RBC) volumes (RBC distribution width: RDW) increases with age and is a strong predictor of mortality, incident CAD and cancer. In a study of 116,666 UK Biobank volunteers, genetic variants explained 29% of RDW individuals aged over 60 years and 33.8% of RDW in those aged < 50 years [222]. RDW was associated with 194 independent genetic signals (119 intronic), 71 implicated in autoimmune disease, body mass index, Alzheimer's disease, longevity, age at menopause, bone density, myostasis, Parkinson's disease and age-related macular degeneration. Pathway analysis showed enrichment for telomere maintenance, ribosomal RNA and apoptosis.

3. Transcriptomics

The transcriptome of centenarians differs from that of septuagenarians [223]. In centenarians 1721 genes were differentially expressed compared with septuagenarians and young people. The most statistically significant associations with biological processes were immune response, followed by cell adhesion and MHC class 1 receptor activity, transport processes, antigen processing and presentation of peptide antigen via MHC class 1, response to drug, ion transport, signal transduction, cell surface receptor linked signaling pathway, small GTPase mediated signal transduction, intracellular signaling pathway, response to wounding, presentation of endogenous peptide antigen. Response to hypoxia, apoptosis, protein transport, T cell activation and processes integral to the plasma membrane [223]. Sub-network analysis converged on 6 genes – interferon-γ gene, IFNG (Fig. 4) – T-cell receptor gene, TCR; tumor necrosis factor gene, TNF; SP1 transcription factor gene, SP1; TGF-β1 gene, TGFB1; and interleukin 32 gene, IL32 – to influence B-cell lymphoma-extra large (Bcl-xL) gene, BCL2L1, Fas and Fas ligand, all involved in the control of apoptosis – Bcl-xL by inhibiting the intrinsic, mitochondrial pathway to apoptosis, and Fas and FasL by controlling the extrinsic pathway to apoptosis. As well as being involved in apoptosis, Bcl-xL is involved in mitochondrial damage protection [224], control of mitochondrial respiration [225], modulation of the immune response [226] and DNA repair [227], all of which are associated with healthy aging. Genes upregulated in centenarians tended to be downregulated in septuagenarians, consistent with activation of those networks in exceptional aging. In Spanish and Sardinian cohorts, BCL2L1 mRNA expression and protein were higher in centenarians than in septuagenarians, but were similar to young individuals, suggesting a major role in healthy aging [223]. In support, transfection of Bcl-xL into mouse embryo fibroblasts and septuagenarian lymphocytes suppressed cell cycle inhibitors, increased cell proliferation, protected against oxidative damage, and delayed the accumulation of senescent cells. A constitutively active mutant of the Caenorhabditis elegans BCL2L1 ortholog, ced-9, increased survival [223]. These findings revealed an important role for BCL2L1 in human aging.

 

Fig. 4.

Example of sub-network analysis result for one of the 6 genes that were prominent in the analysis. Shown are genes regulated by interferon-γ gene, IFNG, in mononuclear cells from (a) centenarians and (b) septuagenarians as compared with young individuals.(From Borras et al. [223])

A transcriptomic prediction model was developed from whole-blood gene expression data for 14,983 individuals of European ancestry from 6 independent cohorts [228]. The 1497 genes differentially expressed with chronological age will be discussed in the epigenetics section below.

RNA profiles of young vs. old human muscle were able to distinguish the age of multiple tissue types [229]. Regulators of the 150 genes identified were identified by reverse genetics and pharmacological methods [230]. Rapamycin perturbed the healthy aging gene expression signature. A degree of direct coordination and a link with mTOR activity pointed to a link between a healthy neuromuscular age biomarker and a major axis of lifespan [230].

4. Epigenetics

4.1. Background

Epigenetics is the study of heritable changes in gene function that do not involve changes in the DNA sequence [231]. The expression of a gene depends not only on the specific sequences but also how it is packaged. DNA can be packaged differently in several ways: (i) DNA is firmly attached to proteins (e.g., histones) responsible for regulating the expression of genes present in this DNA. As a result, half of these proteins end up in the daughter cells after cell division, (ii) DNA can be modified (most commonly at the 5-position of cytosine [C]) and half of these modified sequences end up in the daughter cells following cell division if the hemimethylated DNA is not fully methylated by maintenance methylases following DNA synthesis. DNA modification (e.g., CpG methylation) generally leads to a change in the way the DNA is packaged. As a result, the daughter cells, unless modified in their new environment, will exhibit a similar pattern of gene expression as their mother cells.

Most genes have CpG islands in their promotors that when methylated are fully turned off. When unmethylated, the genes can be active. In the transition from one state to the other there is a cascade of events that occur that involve a host of proteins involved in modifying or maintaining a specific state of chromatin that allows differential access of the DNA to regulatory elements. We now know that the latter process involves not only specific proteins but also RNA that can serve as a “scaffold” to various regulatory proteins. Curiously, several transcription factors preferentially bind CpG methylated sequences [232].

Epigenetic regulation involves posttranslational modification of core histone proteins. This involves the acetylation, phosphorylation, methylation, phosphorylation, sumoylation, ubiquitination and citrullination of histone tails [233,234]. A “histone code” is formed. This involves various combinations of these histone modifications, which differ between organisms. The particular combination influences the pattern of recruitment of transcription factors and coactivators/cosuppressors responsible for regulation of chromatin structure and transcriptional activity of genes. Methylation and acetylation of histone tails appear to be the modifications having the most important effects on gene expression. Histone modifications affect transcriptional activity via two major mechanisms: (i) by changing the structure and conformation of chromatin; (ii) by alerting particular enzymes to recruit transcriptional activators or suppressors. This is a dynamic process. Histone modifications can be added and removed by specific enzymes that target acetyl groups on lysine residues present in the histone N-terminal tails. Addition involves histone acetyltransferases (HATs) and methyltransferases, and removal involves histone deacetylases (HDACs) and lysine demethylases [234-238] (Fig. 5). As a result, transcription of genes can be either activated or repressed.

 

Fig. 5.

Histone modification pathways. (From Li et al. [238])

Histone modifications affect the structure of the nucleosome, which consists of 146-base pairs of DNA wrapped around a histone octamer comprising two sets of H2A, H2B, H3 and H4 monomers [239]. Changes in nucleosome configuration affects the status of chromatin between either compacted (tight-close) or relaxed (loose-open) [236]. The level of chromatin openness affects the degree of gene activity within a DNA region. When lysine residues on histones are deacetylated they have a positive charge, which attracts the negatively charged DNA strand, so resulting in a compact chromatin state that is associated with repression of transcription. On the other hand, when acetylated, the positive charge is lost. This results in an open chromatin structure, so leading to transcriptional activation.

The HDAC family is comprised of at least four classes: class I HDACs (HDAC1, HDAC2, HDAC3 and HDAC8) that are most closely related to the yeast Rpd3 HDAC; class II HDACs (HDAC4, HDAC5, HDAC6, HDAC7, HDAC9 and HDAC10) that share homology domains with the yeast Hda1 enzyme; class III HDACs that comprise sirtuins 1, 2, 3, 4, 5, 6 and 7, which are homologs of the yeast Sir2 enzyme; and lastly, HDAC11, which is the only member of class IV HDACs and is closely related to class I HDACs [238].

In addition to performing deacetylation reactions, HDACs regulate various cellular functions and gene expression by interacting with hundreds of different transcription factors [236,240].

4.2. DNA methylation and aging

DNA methylation data have been used as biomarkers of aging (“epigenetic clocks”), enabling accurate age estimates for any tissue across the lifespan [241]. DNA methylation patterns appear to be gene- and tissue-specific. Globally, DNA in leukocytes has generally been reported to be hypomethylated with age while specific CpC sites in gene promoters tend to be hypermethylated [242,243]. A method has been developed to determine biological age based on the methylation patterns of specific sites in three genes, with an average accuracy of ± 5.2 years [244]. Based on the methylation status of 353 CpGs, accuracy was then improved to an error of ± 3.6 years, with a correlation of 0.96 [245]. Of the 353 CpG sites, 193 became hypermethylated and 160 became hypomethylated with age. The genes are involved in cell death and survival, cell growth and proliferation, organismal and tissue development, and cancer. NAD+-mediated activation of sirtuin-1 deacetylates, and thereby activates, clock-controlled genes as part of circadian expression changes [246]. In the elderly, SIRT1 expression is elevated, possibly as a compensatory response to increased oxidative stress [247].

While monozygous twins are epigenetically indistinguishable during the early years of life, older twins exhibit differences in their overall content and genomic distribution of 5-methylcytosine DNA and histone acetylation, consistent with epigenetic drift with age [248]. Recently, a highly multiplexed mass cytometry analysis was able to show the markedly different cell-type and hematopoietic-lineage-specific chromatin modification patterns in single cells [249]. With aging, chromatin modifications exhibited a marked increase in heterogeneity between individuals and elevation in cell-to-cell variability. Twin studies revealed a genetic component of 30% in variation in chromatin marks, with 70% driven by non-heritable influences. Peripheral blood mononuclear cells of Italian semi-supercentenarians (age 105–109 years) had an 8.6-year younger epigenetic profile than expected, and their offspring (aged 50–89 years) had a 5.1 year lower epigenetic age than age-matched controls [250].

The aging cell is exposed to ROS that can increase inflammation, induce DNA damage and influence DNA methyltransferase (DMT) activity. The damage results in an increase in mutation frequency and affects the DNA methylation of nearby cytosine bases. Hydroxylation of guanine (to yield 8-hydroxy-guanine) causes a > 90% decrease in methylation of neighboring cytosine [251]. It was suggested that this process may help to explain the age-related drift in hypomethylation.

The methylation patterns are distinct in the innate and adaptive arms of the human immune system during hematopoiesis [252]. A progressive loss of CG methylation was found in developing lymphocytes and there was consistent occurrence of non-CG methylation (i.e., C hydroxymethylation) in specific cell types, such as T-cells.

There are many examples of specific genes in the innate immunity pathway that are dysregulated by aging and/or methylation. For example, differential methylation (hypo- vs. hyper-methylation) was seen in 1859 genes in rheumatoid arthritis [253].

A whole-blood gene expression meta-analysis identified 1497 genes that were differentially expressed with chronological age [228]. These genes were enriched for potentially functional CpG methylation sites in enhancer and insulator regions that were associated with both chronological age and gene expressions levels. Gene expression profiles were used to calculate “transcriptomic age” of individuals. Differences between transcriptomic and chronological age were associated with aging-related biological parameters such as blood pressure, total and HDL cholesterol levels, fasting glucose, body mass index, hand grip strength, and mini mental state test score. Transcriptomic prediction complemented epigenetic prediction of age [228]. DNA methylation profile difference between Chinese female centenarians and middle-aged controls revealed 626 differentially methylated regions, within which was enrichment of age-related disease genes [254].

4.3. Chromatin modification and aging

Histone methylation has a great impact on whether a region is active or repressed and depends on the exact location (amino acid) and degree of methylation (mono-, di, tri-). It is dynamically regulated by histone methyltransferases and histone demethylases (see review: [255]). Histone acetylation influences how the DNA will interact with histones. Fig. 6 shows the changes in chromatin state with aging [256].

 

Fig. 6.

The changes in chromatin states with aging. Increased cellular senescence results in a loss of heterochromatin caused by the factors depicted in the diagram. (From Booth and Brunet [256])

Histone methylation H3K4me3 is believed to be activating to genes, while H3K27me3 is repressing to genes, and this pattern has been directly linked to lifespan regulation in mice [257], Drosophila melanogaster [258] and C. elegans [259], wherein these bivalent sites become hypomethylated with age. Life extending treatments such as CR and rapamycin suppress this hypomethylation [257]. Aging in C. elegans results in a global decrease in somatic H3K27me3. Treatments that increase global levels of H3K27me3 extend lifespan in C. elegans, but not consistently in other organisms, so is controversial.

Core histone proteins are known to diminish with age in yeast [260], nematode [261] and human [262] cells. The level of core histone variant, macroH2A, a splice variant of histone H2A, has been reported to increase with age during replicative senescence in cultured human fibroblast cells and in aged mice and primates [263]. MacroH2A is assumed to be a transcriptional repressor [264]. The sirtuin 1 ortholog, Sir2, decreases and subtelomeric locations of histone H4 lysine 16 acetylation increase, along with histone loss, with age in yeast [260]. CR and genetic models of longevity in mice result in a common transcriptional signature that includes mitochondrial energy metabolism, inflammation and ribosomal structure in multiple mouse tissues [265]. In response to CR, mice lacking Sirt3 fail to induce mitochondrial and anti-inflammatory elements of this signature, whereas the inverse of this signature is seen in response to a high fat diet, obesity and metabolic disease [265].

In aging, there is a general loss and disorganization of histones that is assumed to lead to the dysregulation of underlying genes. Evidence for this is an abnormal phasing of histones and induction of repressed genes in yeast [266]. This would suggest an abnormality or limited supply of the components that control gene regulation (i.e., promoter-recognizing proteins/transcription factors).

Age-related DNA hypermethylation has been reported to be enriched at the genomic regions carrying bivalent histone marks (i.e., both H3 K4me3 and H3 K27me3) at promoters, whereas DNA hypomethylation colocalizes with the histone modification marks H3 K9Ac, H3 K27Ac, H3 K4me1, H3 K4me2, and H3 K4me3 that are found largely in enhancer regions [267].

Global H3K27me3 levels increase with age in some organisms [268], while they decrease with age in others [261]. Identifying the locus- and cell-type-specific dynamics will be critical to obtain a better understanding of factors that influence lifespan.

Fig. 7 highlights the physiological processes affected by epistatic changes to the genome during aging [256].

 

Fig. 7.

Epigenomic changes result in dysregulation of gene expression in aging. (From Booth and Brunet [256])

4.4. Epigenetics and longevity

Caloric restriction of 30% extends lifespan of a small primate model by 50% [269]. Aging-related diseases were decreased, white matter was preserved, and although there was a claimed decrease in gray matter (neuronal cell bodies) in the cerebrum, there was no change in cognitive performance, behavior and motor function. With aging, there is a marked drift in both gains and losses of DNA methylation that correlates with lifespan and is attenuated by CR [270]. CR is accompanied by increased HDAC activity. This suggests that in response to nutritional stress global deacetylation may serve to protect cells and thereby influence the aging processes [271]. Two key genes implicated in aging – the p16INK4a gene, CDKN2A, and the human telomerase reverse transcriptase gene, TERT – are upregulated during CR. This is accompanied by enrichment in HDAC1 on the promoters of each, which contributes to longevity [238,271]. During senescence, methyl-CpG-binding protein hypermethylates the SIRT1 promoter and causes histone modification, leading to dysfunction of endothelial progenitor cells, increased apoptosis and reduced angiogenesis [272]. The importance of HDACs in CR point to the potential of drugs or other strategies for modulation of cell epigenetics in order to slow aging and treat diseases of aging.

Sirtuin 1 has an important impact on aging and extension of lifespan in response to CR [17,273,274]. Sirtuin 1 activity is tied to metabolic activity of cells by its dependence on intracellular NAD/NADH ratio, which reflects oxygen consumption, respiratory chain activity and metabolic rate. Histone acetylation is associated with an open chromatin status leading to gene activation. In contrast, histone methylation has differential effects on the binding of various proteins to histones, leading to either activation or silencing of gene expression [235] (Fig. 8). Lysine residues on histones can be mono-, di- or trimethylated, and either activation or repression is dependent upon the particular lysine residue that is modified [275,276]. Such histone modifications modulate CDNK2A and TERT expression, so affecting their expression and thus aging and longevity of human cells [238].

 

Fig. 8.

Caloric restriction regulates epigenetic processes by DNA methylation and histone modification. Via the pathways shown, this helps increase lifespan. (From: Li et al. [238])

During normal aging, changes in gene expression and epigenetic modification occur in a tissue-specific manner. In mammals, 80% of all CpG sites are methylated [277]. Clusters of CpG dinucleotides (CpG islands) are often located near transcription start sites (TSS) of genes [278]. Most CpG islands are normally unmethylated, but during healthy aging of tissues methylation changes occur in a small subset of genes. Such age-related methylated genes have been seen in human whole blood [279-281] and can be used as a biomarker of biological (epigenetic) age [244,280,282]. DNA methylations seems to largely reflect the passage of chronological time, being inferior to a frailty index composed of 34 health items as a predictor of mortality [283]. The epigenetic clock is based on the 353 specific CpG sites spanning the genome [245,284,285].

Erosion of highly organized methylation patterns (“methylation drift”) is negatively correlated with lifespan and is strongly conserved across species [270]. Chronic inflammation, which shortens lifespan, accelerates methylation drift [286]. CR delays aging-related methylation drift in mouse, rhesus monkey and human blood cells and other tissues, and results in a significantly younger “methylation age” [270]. The rate of drift correlated with lifespan. Thus, methylation drift is an excellent biomarker of aging and lifespan. Since methylation drift correlates with changes in gene expression it is most probably a mediator of age-related functional decline and disease. In support, the aging-associated increase in hypomethylation at super-enhancers of highly expressed genes crucial for liver function in ad libitum fed mice was suppressed in CR and Ames dwarf mice, with methylation changes occurring more selectively, but less specifically, in rapamycin-treated mice [257]. Hypermethylation was enriched at CpG islands marked with bivalent activating and repressing histone modifications and resembled hypermethylation seen in liver cancer. Genome-wide methylation profiling in nonagenarians identified 19 mortality-associated CpG sites that mapped to genes whose functions were clustered around the nuclear factor κB complex, indicating an important role for this complex in human longevity [287].

During aging, dynamic changes in DNA methylation in the brain include global hypomethylation [288-291] and site-specific hypermethylation predominantly in promoter regions [292,293]. Many of the genes hypermethylated in the aged brain are involved in neurodevelopment [294,295]. Most hypermethylation correlates with gene silencing [296]. Transcriptional repression rather than induction is seen in the cerebral cortex with aging [292]. The global hypomethylation of the aging brain is associated with decrease in DMT activity [297]. In cortices from Alzheimer's patients 5-methyl cytosine is reduced [298]. Such a reduction correlates with increased neurofibrillary tangle formation [298]. Diminished capacity for re-establishment of normal methylation pattern after DNA demethylation likely affects long-term neuronal survival and degeneration with age [299]. Potential interactions of DMTs with histone modifications could further contribute to the regulation of neuronal function and survival during aging.

Certain regions of the genome are imprinted differently between mother and father, via chromatin modifications that include DNA methylation [300,301]. While it was assumed that all marks were erased during these phases, some methylation marks are not fully erased, and persist. There is epidemiological evidence for intergenerational and transgenerational inheritance of longevity and lifespan-limiting conditions in human populations [302]. Long-lived individuals exhibited a young epigenetic profile compared with their chronological age in the Sydney Centenarian Study [303]. Unique DNA methylation patterns and low overall variability in DNA methylation was observed in high life expectancy region of Costa Rica [304].

Using RNA-seq data sets for 21 somatic cell types and tissues whose cellular turnover (lifespan) ranged from 2 days (monocytes) to lifetime (neurons), the gene expression profile of neurons showed reduced protein metabolism, consistent with survival and CR [305]. A gene expression signature of turnover was negatively correlated with the energetically high-cost cell cycle and factors supporting genome stability. These are risk factors for aging-associated pathologies. Expression of the tumor-suppressor p53 gene (TP53) was 5–30 times lower than in other cells and tissues. The reduced p53 expression in neurones was associated with reduced cell-cycle related proteins. Low p53 expression may contribute to the exceptional lifespan of low-turnover cells and tissues such as neurons, heart muscle and skeletal muscle, and perhaps human longevity.

4.5. Chromatin modification and cell compartmentalization

Heterochromatin is present in the cell in several forms. Constitutive heterochromatin is highly condensed, is composed of highly repetitive sequences. It is found primarily adjacent to chromosome centromeres and positioned near the nuclear membrane during interphase. In contrast, facultative heterochromatin is less condensed, is capable of transitioning between condensed and decondensed states, generally consists of condensed domains interposed with euchromatin and these condensed regions, too, are near the nuclear periphery. The inactive X chromosome in females is an example of facultative heterochromatin, most of the genes being transcriptionally inactive. Most chromosomes comprise combinations of heterochromatin and euchromatin.

Histone modification can act to silence transcription via the formation of facultative heterochromatin (i.e., H3K9me3 and H4K20me2) and to regulate genome stability (involving H3K56ac and H3K14ac). The patterns and changes in histone modification are highly tissue- and gene-specific [306]. Active histone modification H3K4me3 (a mark of promoter accessibility) occurs during aging and in cellular senescence [261].

Lamins and their associated proteins help to form the nuclear lamina which maintains the structural integrity of the nucleus, organizes chromatin, facilitates nuclear assembly/disassembly during mitosis, and acts as a scaffold for DNA repair [307]. Chromatin that is in close proximity to the nuclear lamina, is largely heterochromatic, and is characterized by low gene density, a repressive chromatin configuration, and is flanked by insulator protein (CTCF) binding sites [308]. When mouse embryonic stem cells are induced to differentiate into either astrocytes or adipocytes, many genes that are normally present at the nuclear periphery locate to the nuclear interior where they are primed for expression at a later differentiation state [309,310]. This transition relies on histone methyltransferases and deacetylases [311].

There is a general loss and redistribution of heterochromatin in senescent cells [312]. This leads to cellular dysfunction with age. In the senescent cell > 30% of the chromatin is associated with senescence-associated heterochromatin foci (SAHF), which are regions of highly condensed chromatin associated with heterochromatic histone modifications (H3K4me3 and H3K27me3), heterochromatic proteins, histone variant macroH2A, high-mobility group A (HMGA) proteins and are late replicating regions in the genome [313]. The latter are generally localized over lamin-associated domains. Breakdown of the nuclear lamina is believed to cause a loss of heterochromatin organization, in which there is a transposition of euchromatin, facultative heterochromatin (fHC), and constitutive heterochromatin (cHC), with cHC moving away from the nuclear lamin and to the center of chromosomal territories [313]. While most mutations cause a variety of autosomal dominant disorders, specific mutations in the lamin A gene, LMA, are known to cause the premature aging syndrome Hutchinson-Gilford progeria syndrome (HGPS) [314], in which a point mutation in a cryptic splice site leads to a truncated protein [315]. The same molecular mechanism responsible for HGPS is also active in healthy cells [315]. Age-related changes in histone modification and increased DNA damage have been attributed to sporadic use of this cryptic splice site as inhibition of this site reversed the associated nuclear defects [315].

By high-throughput conformation capture (Hi-C), age-associated differences were demonstrated in local chromatin connectivity between embryonic stem cells, somatic cells, and senescent cells, suggesting that senescence is an endpoint of the continuous nuclear remodeling process during differentiation [316]. With age-related breakdown of the nuclear lamin there was a loss of local interactions within topologically-associated domains (TADs) and an increase in distant interactions between TADs in the senescent state, thus supporting the premise that senescence is associated with a change in physical chromatin compaction. Neighboring genes that were transcriptionally isolated from one another could become co-regulated. It was also found that the regions of the genome that changed the most were the least accessible and were high in AT content. These characteristics are descriptive of “satellite DNA” that is highly repetitive and contains a high concentrations of centromeric constitutive heterochromatin [317]. Fig. 9 depicts changes that occur with cell senescence.

 

Fig. 9.

Breakdown of the nuclear lamina, heterochromatin, and chromatin connections in senescence. With senescence comes a dissolution of the tight control of gene regulation as the inactive chromatin becomes dissociated from the nuclear envelope, topological-associated domains (TADs) lose neighboring connections in exchange for long-range interaction, and the distinction between active and inactive chromatin domains becomes blurred.

5. Non-coding RNA and aging

Only 1.5–1.8% of the mammalian genome is transcribed to yield proteins. The rest is transcribed into noncoding RNAs (ncRNAs) that mostly include small (20–30 nt sncRNAs) and long (> 200 nt lncRNAs) [318,319]. Noncoding RNAs represent a vast pool in the mammalian genome that may provide some of the missing links in the aging process [320].

SncRNAs are comprised mainly of inhibitory RNAs such as microRNAs (miRNAs), small inhibitory RNAs (siRNAs) and piwi-interesting RNAs (piRNAs) involved mainly in post-transcriptional gene regulation, mediated by mRNA degradation or disruption of translation. These have been studied in aging [321-325]. PCR arrays of serum samples from Baltimore Longitudinal Study of Aging subjects identified multiple differentially expressed miRNAs, including 6 miRNAs that correlated with subsequent longevity [326]. Five of these miRNAs targeted 24 aging-associated mRNAs which included PARP1, IGF1R and IGF2R mRNAs.

LncRNAs are transcribed from the intergenic and intronic regions of the mammalian genome [319,327,328]. While the transcriptional regulatory sequences of lncRNAs are evolutionarily conserved, many lncRNAs are species- (including primate-) specific. Just as for mRNA, lncRNAs are transcribed by RNA polymerase II, then introns are removed, 5'methyl-capping takes place, and poly-adenylation occurs. They are characterized by well-defined transcriptional landmarks (histone H3-lysine 4 (H3K4) and H3K36 tri-methylation, H3K9 acetylation and CpG DNA methylation), and are regulated by common transcription factors such as p53, Oct and Nanog [328-330]. Based on their genomic location, orientation and mode of transcription, lncRNAs are arbitrarily classified further into sense, antisense, bi-directional, promoter-associated, enhancer-associated, pseudogene-associated, telomere-associated, circular and repeat-rich lncRNAs in a nonexclusive manner [319,327,331,332]. Most lncRNAs exhibit specific subcellular, cell, tissue and developmental expression patterns in mammals, which supports them having important biological functions such as development and differentiation, cell survival, cell proliferation and apoptosis, dosage compensation and gene imprinting, reprogramming of differentiated cells and maintenance of stem cells [320].

Most of the characterized lncRNAs are nuclear localized and act as enhancer RNAs (eRNAs), chromatin modifiers via recruitment of various DMTs, and histone modifiers via polycomb repressive complexes or histone methyltransferases [320]. Many lncRNAs affect key cellular processes such as proliferation, differentiation, quiescence, senescence, stress and immune response, and many other cellular functions related to aging [333].

A list of lncRNAs involved in regulation of cellular senescence and aging has been compiled [320]. Some are antisense transcripts complementary to mRNA transcripts of the same gene [320]. Among these are lncRNAs differentially expressed in senescent cells [334] and that suppress mRNA expression [320]. Other lncRNAs are encoded by pseudogenes and act to negatively regulate the expression levels of the corresponding mRNA for cell adhesion molecules and translation machinery components in senescent cells. These effects could affect cell morphology, growth and division, as well as expression of various senescence-associated proteins, so contributing to cellular senescence and aging [320]. Yet other lncRNAs are encoded in intergenic DNA and affect a multitude of different intracellular processes, some of which may contribute to senescence and aging [320]. A class of heterogeneous 5′-UUAGGG-3′ repeat-containing lncRNAs – telomere repeat containing RNA (TERRA) – is partly associated with telomeric heterochromatin [335-337]. When strongly upregulated, TERRA can promote telomere dysfunction such as shortening, decreased stability and heterochromatin formation [320].

Genetic variants in noncoding RNAs have the potential to differentially influence aging processes and lifespan. The lncRNA ANRIL, was implicated in longevity in a GWAS [96]. Understanding genetic variants in the noncoding genome is a current challenge. Constrained regions in the noncoding genome associate with the most essential genes and are enriched for pathogenic variants [338].

The immune system changes over time, increasing pro-inflammatory activity by innate immune cells such as monocytes and macrophages, and decreasing the immune response [339]. A progressive change in expression of 69 non-coding RNAs (56 microRNAs and 13 snoRNAs) is seen with chronological age [340]. The age-related miRNAs were found to regulate genes involved in immune, cell cycle and cancer-related pathways.

A number of miRNAs have been reported to have increased expression with age in mouse [341] and rat [342] livers. Up-regulated miRNA targets are associated with detoxification activity and regeneration capacity – functions known to decline in old liver, consistent with the negative regulatory roles of miRNAs [341]. By measuring expression patterns in mouse liver following CR it was found that: (i) the expression of miRNAs, lncRNAs and transposable elements were largely repressed; (ii) the protein-coding mRNAs that demonstrated increased expression in CR are highly targeted by miRNAs; and (iii) the miRNA-targeting sites were enriched for genes having chromatin-related functions. One such gene is chromodomain helicase DNA binding protein 1 gene, Chd1, which is instrumental in chromatin remodeling [343].

6. Longevity gene networks

To understand the genetic and epigenetic landscape of human aging a meta-analysis was performed of 6600 human longevity genes from 35 datasets comprising 8 curated aging data sets (1154 genes), 10 age-related diseases (1207 genes), 4 gene expression sets (2130 genes), and 13 methylation sets (3498 genes) [344]. From this, biological relationships between aging-associated genes were investigated by producing a protein interaction network and characterizing network neighborhoods. Most genes only appeared in a single category, whereas 1050 were associated with two, 159 with three, and 7 with all four categories. Those 7 were APOE, chaperonin containing TCP1 subunit 7 gene, CCT7, erb-b2 receptor tyrosine kinase 1 gene, ERBB2, protein kinase C alpha gene, PRKCA, Ras association domain family member 1 gene, RASSF1, sterol regulatory element binding transcription factor 1 gene, SREBF1, and tumor necrosis factor gene, TNF. The TNF receptor family member ectodysplasmin A receptor (EDAR) associated death domain gene, EDARADD, and lymphocyte activating 3 gene, LAG3, exhibited the strongest evidence for aging-associated DNA methylation changes. Of the 6600 aging-associated genes, 5949 had at least one interaction. There were 1079 human aging clusters in the combined interaction network. The interaction network provided an additional layer for linking proteomic and genomic data. On the basis of “guilt by association”, a previously unsuspected gene may be a candidate if its encoded protein is found to physically interact with a protein known to be involved in the condition or pathway. Clusters with a strong aging association included one containing mTOR signaling pathway members, and one in which Werner syndrome RecQ like helicase gene, WRN, was one of 16 members. Another lacked genes associated previously with aging, but linked 8 genes differentially methylated with age, and 7 differentially expressed in response to CR.

7. Potential anti-aging interventions

Because CR delays aging and ameliorates risk of aging-related diseases, but adherence in human populations is burdensome, attempts have been made to identify natural or synthetic compounds that mimic the effects of CR [238]. “Epigenetic diets” that favourably influence the epigenetic profile of individuals have been described, together with natural compounds able to mediate effects of such diets [345]. Prominent among these is resveratrol, a sirtuin 1 activator, able to promote healthy aging and increase longevity [346-354]. Others include spermidine, the antidiabetic drug metformin, selenium, synthetic sirtuin-activating compounds such as SRT1720 and SRT2104, senolytics (e.g., dasatinib plus quercetin), and the NAD+ booster nicotinamide mononucleotide. Dietary components such as green tea, broccoli sprouts and soybeans, and the bioactive compounds extracted from these diets have received extensive attention due to their ability to favourably alter the epigenetic landscape in cancer cells [355-358]. Long-term consumption of epigenetic diets may alter chromatin profiles, slow aging and reduce risk of degenerative diseases of aging such as cancer, cardiovascular disease, type 2 diabetes and neurodegenerative disorders [359-366], suggesting that these bioactive diets may affect aging processes by altering chromatin profiles that also occur in CR [367]. Global gene expression profiling methods have been developed to identify CR mimetics able to delay aging [368].

A potent specific thiazoquin(az)oli(one)e CD38 inhibitor, 78C, reverses age-related NAD+ decline and improves various aging-related physiological and metabolic parameters [369]. The elevation in NAD+ led to activation of pro-longevity and health-span-related factors such as sirtiuns, AMPK and PPARs, and inhibition of pathways such as mTOR-S6K and ERK, having a negative impact on health span.

8. Future directions

Despite the extensive findings described in this review, contrary to many complex, age-related diseases, consensus on the ultimate set of multi-biomarker aging or lifespan-related phenotypes for genetic and genomic studies has not yet been reached [370]. Awareness is needed of racial differences, sex-specificity of longevity associations, variation in results depending on the definition used for longevity [11], choice of controls, such as whether to use a random population sample or individuals who have died before a certain age [49]. The fact that most studies of human longevity involve long-lived cases and younger controls, has resulted in risk of obtaining false positive or false-negative findings from population stratification artifact and cohort effects, among other epidemiological biases. Large cohort studies with prospectively collected data are optimal since cases and controls come from similar birth cohorts, but such studies are expensive and few in number because they require decades-long follow-up and large biorepositories for the study of the mechanisms of aging and aging-related phenotypes.

There is an interplay between genetic and behavioral risks. Individuals aged ≥75 years with multiple adverse alleles (such as those involving TOMM40/APOE/APOC1, insulin-degrading enzyme gene, IDE, and PI3K catalytic subunit β gene, PI3KCB) had a 62% higher mortality rate than those with none, whereas people with a low-risk behavioral profile had a 65% lower mortality rate than those with a high-risk behavioral profile [371]. A challenge will be interpretation of vast volumes of genetic data that will continue to emerge [26].

More research is needed on interventions able to forestall the “telomere brink”, given the opposing association of short leukocyte telomere length (LTL) and alleles associated with LTL with increase coronary risk, and long LTL being associated with increased cancer risk [167,372].

Epigenetic factors include environmental influences and lifestyle choices, but also the microbiome [373]. The latter in particular merits further research. Dysregulation of transcriptional and chromatin networks is likely a crucial component of aging [256]. Epigenomic changes during aging profoundly affect cellular function and stress resistance [256]. Further work aimed at understanding age-dependent epigenomic changes should lead to key insights into the aging process and development of means of delaying or even reversing the changes and countering age-related diseases.

Acknowledgements

We thank the National Heart, Lung, and Blood Institute, the National Institute on Aging, and the Hawaii Community Foundation for support.

Footnotes

This article is part of a Special Issue entitled: Genetic and epigenetic regulation of aging and longevity edited by Jun Ren & Megan Yingmei Zhang.

Transparency document

The Transparency document associated with this article can be found, in online version.

Go to:

References

  1. McGue M, Vaupel JW, Holm N, Harvald B, Longevity is moderately heritable in a sample of Danish twins born 1870–1880, J. Gerontol48 (1993) B237–B244. [PubMed] [Google Scholar]
  2. Murabito JM, Yuan R, Lunetta KL, The search for longevity and healthy aging genes: insights from epidemiological studies and samples of long-lived individuals, J. Gerontol. A Biol. Sci. Med. Sci67 (2012) 470–479. [PMC free article] [PubMed] [Google Scholar]
  3. Sebastiani P, Perls TT, The genetics of extreme longevity: lessons from the new England centenarian study, Front. Genet3 (2012) 277. [PMC free article] [PubMed] [Google Scholar]
  4. Philippe P, Familial correlations of longevity: an isolate-based study, Am. J. Med. Genet2 (1978) 121–129. [PubMed] [Google Scholar]
  5. Mayer PJ, Inheritance of longevity evinces no secular trend among members of six New England families born 1650–1874, Am. J. Hum. Biol3 (1991) 49–58. [PubMed] [Google Scholar]
  6. Ljungquist B, Berg S, Lanke J, McClearn GE, Pedersen NL, The effect of genetic factors for longevity: a comparison of identical and fraternal twins in the Swedish Twin Registry, J. Gerontol. A Biol. Sci. Med. Sci. 53(1998) M441–M446. [PubMed] [Google Scholar]
  7. Herskind AM, McGue M, Holm NV, Sorensen TI, Harvald B, Vaupel JW, The heritability of human longevity: a population-based study of 2872 Danish twin pairs born 1870–1900, Hum. Genet97 (1996) 319–323. [PubMed] [Google Scholar]
  8. Mitchell BD, Hsueh WC, King TM, Pollin TI, Sorkin J, Agarwala R, Schaffer AA, Shuldiner AR, Heritability of life span in the Old Order Amish, Am. J. Med. Genet102 (2001) 346–352. [PubMed] [Google Scholar]
  9. Kerber RA, O'Brien E, Smith KR, Cawthon RM, Familial excess longevity in Utah genealogies, J. Gerontol. A Biol. Sci. Med. Sci56 (2001) B130–B139. [PubMed] [Google Scholar]
  10. Kaplanis J, Gordon A, Shor T, Weissbrod O, Geiger D, Wahl M, Gershovits M, Markus B, Sheikh M, Gymrek M, Bhatia G, MacArthur DG, Price AL, Erlich Y, Quantitative analysis of population-scale family trees with millions of relatives, Science360 (2018) 171–175. [PMC free article] [PubMed] [Google Scholar]
  11. Sebastiani P, Nussbaum L, Andersen SL, Black MJ, Perls TT, Increasing sibling relative risk of survival to older and older ages and the importance of precise definitions of “aging”, “life span,” and “longevity”, J. Gerontol. A Biol. Sci. Med. Sci71 (2016) 340–346. [PMC free article] [PubMed] [Google Scholar]
  12. Berg NVD, Rodriguez-Girondo M, de Craen AJM, Houwing-Duistermaat JJ, Beekman M, Slagboom PE, Longevity around the turn of the 20th century: life-long sustained survival advantage for parents of today's nonagenarians, J. Gerontol. A Biol. Sci. Med. Sci(2018) (Pub ahead of print Mar 17, 2018). [PMC free article] [PubMed] [Google Scholar]
  13. Takata H, Suzuki M, Ishii T, Sekiguchi S, Iri H, Influence of major histocompatibility complex region genes on human longevity among Okinawan-Japanese centenarians and nonagenarians, Lancet2 (1987) 824–826. [PubMed] [Google Scholar]
  14. Schachter F, Fauredelanef L, Guenot F, Rouger H, Froguel P, Lesueurginot L, Cohen D, Genetic associations with human longevity at the APOEand ACE loci, Nat. Genet 6 (1994) 29–32. [PubMed] [Google Scholar]
  15. Kenyon CJ, The genetics of ageing, Nature464 (2010) 504–512. [PubMed] [Google Scholar]
  16. Barzilai N, Huffman DM, Muzumdar RH, Bartke A, The critical role of metabolic pathways in aging, Diabetes61 (2012) 1315–1322. [PMC free article] [PubMed] [Google Scholar]
  17. Morris BJ, Seven sirtuins for seven deadly diseases of aging, Free Radic. Biol. Med56 (2013) 133–171. [PubMed] [Google Scholar]
  18. Argon Y, Gidalevitz T, Candidate genes that affect aging through protein homeostasis, Adv. Exp. Med. Biol847 (2015) 45–72. [PubMed] [Google Scholar]
  19. Blackburn EH, Epel ES, Lin J, Human telomere biology: a contributory and interactive factor in aging, disease risks, and protection, Science350 (2015) 1193–1198. [PubMed] [Google Scholar]
  20. Novelli V, Viviani Anselmi C, Roncarati R, Guffanti G, Malovini A, Piluso G, Puca AA, Lack of replication of genetic associations with human longevity, Biogerontology9 (2008) 85–92. [PubMed] [Google Scholar]
  21. Shadyab AH, Lacroix AZ, Genetic factors associated with longevity: a review of recent findings, Ageing Res. Rev19 (2015) 1–7. [PubMed] [Google Scholar]
  22. Soerensen M, Nygaard M, Debrabant B, Mengel-From J, Dato S, Thinggaard M, Christensen K, Christiansen L, No association between variation in longevity candidate genes and aging-related phenotypes in oldest-old Danes, Exp. Gerontol78 (2016) 57–61. [PMC free article] [PubMed] [Google Scholar]
  23. Tacutu R, Craig T, Budovsky A, Wuttke D, Lehmann G, Taranukha D, Costa J, Fraifeld VE, de Magalhães JP, Human ageing genomic resources: integrated databases and tools for the biology and genetics of ageing, Nucleic Acids Res. 41(2013) D1027–D1033. [PMC free article] [PubMed] [Google Scholar]
  24. Tan Q, Zhao JH, Iachine I, Hjelmborg J, Vach W, Vaupel JW, Christensen K, Kruse TA, Power of non-parametric linkage analysis in mapping genes contributing to human longevity in long-lived sib-pairs, Genet. Epidemiol26 (2004) 245–253. [PubMed] [Google Scholar]
  25. Stepler R, World's Centenarian Population Projected to Grow Eightfold by 2050. Pew Research Cener, http://www.pewresearch.org/fact-tank/2016/04/21/worlds-centenarian-population-projected-to-grow-eightfold-by-2050/, (2016) (last accessed Apr 19, 2018). [Google Scholar]
  26. Giuliani C, Pirazzini C, Delledonne M, Xumerle L, Descombes P, Marquis J, Mengozzi G, Monti D, Bellizzi D, Passarino G, Luiselli D, Franceschi C, Garagnani P, Centenarians as extreme phenotypes: an ecological perspective to get insight into the relationship between the genetics of longevity and age-associated diseases, Mech. Ageing Dev165 (2017) 195–201. [PubMed] [Google Scholar]
  27. Santos-Lozano A, Santamarina A, Pareja-Galeano H, Sanchis-Gomar F, Fiuza-Luces C, Cristi-Montero C, Bernal-Pino A, Lucia A, Garatachea N, The genetics of exceptional longevity: insights from centenarians, Maturitas90 (2016) 49–57. [PubMed] [Google Scholar]
  28. Rall SC Jr., Weisgraber KH, Mahley RW, Human apolipoprotein E. The complete amino acid sequence, J. Biol. Chem257 (1982) 4171–4178. [PubMed] [Google Scholar]
  29. Gill LL, Peoples OP, Pearston DH, Robertson FW, Humphries SE, Cumming AM, Hardman N, Isolation and characterisation of a variant allele of the gene for human apolipoprotein E, Biochem. Biophys. Res. Commun130 (1985) 1261–1266. [PubMed] [Google Scholar]
  30. Weisgraber KH, Rall SC Jr., Mahley RW, Human E apoprotein heterogeneity. Cysteine-arginine interchanges in the amino acid sequence of the apo-E isoforms, J. Biol. Chem256 (1981) 9077–9083. [PubMed] [Google Scholar]
  31. Das HK, McPherson J, Bruns GA, Karathanasis SK, Breslow JL, Isolation, characterization, and mapping to chromosome 19 of the human apolipoprotein E gene, J. Biol. Chem260 (1985) 6240–6247. [PubMed] [Google Scholar]
  32. Paik YK, Chang DJ, Reardon CA, Davies GE, Mahley RW, Taylor JM, Nucleotide sequence and structure of the human apolipoprotein E gene, Proc. Natl. Acad. Sci. U. S. A82 (1985) 3445–3449. [PMC free article] [PubMed] [Google Scholar]
  33. Williams GC, Pleiotropy, natural selection, and the evolution of senescence, Evolution11 (1957) 398–411. [Google Scholar]
  34. Yanai H, Budovsky A, Barzilay T, Tacutu R, Fraifeld VE, Wide-scale comparative analysis of longevity genes and interventions, Aging Cell16 (2017) 1267–1275. [PMC free article] [PubMed] [Google Scholar]
  35. van Exel E, Koopman JJE, Bodegom DV, Meij JJ, Knijff P, Ziem JB, Finch CE, Westendorp RGJ, Effect of APOE epsilon4 allele on survival and fertility in an adverse environment, PLoS One12 (2017) e0179497. [PMC free article] [PubMed] [Google Scholar]
  36. Trumble BC, Stieglitz J, Blackwell AD, Allayee H, Beheim B, Finch CE, Gurven M, Kaplan H, Apolipoprotein E4 is associated with improved cognitive function in Amazonian forager-horticulturalists with a high parasite burden, FASEB J. 31(2017) 1508–1515. [PMC free article] [PubMed] [Google Scholar]
  37. Guo Y, Lanktree MB, Taylor KC, Hakonarson H, Lange LA, Keating BJ, Genecentric meta-analyses of 108,912 individuals confirm known body mass index loci and reveal three novel signals, Hum. Mol. Genet22 (2013) 184–201. [PMC free article] [PubMed] [Google Scholar]
  38. Chasman DI, Kozlowski P, Zee RY, Kwiatkowski DJ, Ridker PM, Qualitative and quantitative effects of APOEgenetic variation on plasma C-reactive protein, LDL-cholesterol, and apoE protein, Genes Immun. 7 (2006) 211–219. [PubMed] [Google Scholar]
  39. Bennet AM, Di Angelantonio E, Ye Z, Wensley F, Dahlin A, Ahlbom A, Keavney B, Collins R, Wiman B, de Faire U, Danesh J, Association of apolipoprotein E genotypes with lipid levels and coronary risk, JAMA298 (2007) 1300–1311. [PubMed] [Google Scholar]
  40. Xu M, Zhao J, Zhang Y, Ma X, Dai Q, Zhi H, Wang B, Wang L, Apolipoprotein E gene variants and risk of coronary heart disease: a meta-analysis, Biomed. Res. Int2016 (2016) 3912175. [PMC free article] [PubMed] [Google Scholar]
  41. Corder EH, Saunders AM, Strittmatter WJ, Schmechel DE, Gaskell PC, Small GW, Roses AD, Haines JL, Pericak-Vance MA, Gene dose of apolipoprotein E type 4 allele and the risk of Alzheimer's disease in late onset families, Science261 (1993) 921–923. [PubMed] [Google Scholar]
  42. Bloom GS, Amyloid-beta and tau: the trigger and bullet in Alzheimer disease pathogenesis, JAMA Neurol. 71(2014) 505–508. [PubMed] [Google Scholar]
  43. Cullen KM, Kócsi Z, Stone J, Microvascular pathology in the aging human brain: evidence that senile plaques are sites of microhaemorrhages, Neurobiol. Aging27 (2006) 1786–1796. [PubMed] [Google Scholar]
  44. Gerdes LU, Jeune B, Ranberg KA, Nybo H, Vaupel JW, Estimation of apolipoprotein E genotype-specific relative mortality risks from the distribution of genotypes in centenarians and middle-aged men: apolipoprotein E gene is a “frailty gene”, not a “longevity gene”, Genet. Epidemiol19 (2000) 202–210. [PubMed] [Google Scholar]
  45. Arai Y, Sasaki T, Hirose N, Demographic, phenotypic, and genetic characteristics of centenarians in Okinawa and Honshu, Japan: part 2 Honshu, Japan, Mech. Ageing Dev165 (2017) 80–85. [PubMed] [Google Scholar]
  46. Tindale LC, Leach S, Spinelli JJ, Brooks-Wilson AR, Lipid and Alzheimer's disease genes associated with healthy aging and longevity in healthy oldest-old, Oncotarget8 (2017) 20612–20621. [PMC free article] [PubMed] [Google Scholar]
  47. Shadyab AH, Kooperberg C, Reiner AP, Jain S, Manson JE, Hohensee C, Macera CA, Shaffer RA, Gallo LC, Lacroix AZ, Replication of genome-wide association study findings of longevity in White, African American, and Hispanic women: the Women's Health Initiative, J. Gerontol. A Biol. Sci. Med. Sci72 (2017) 1401–1406. [PMC free article] [PubMed] [Google Scholar]
  48. Garatachea N, Marín PJ, Santos-Lozano A, Sanchis-Gomar F, Emanuele E, Lucia A, The ApoE gene is related with exceptional longevity: a systematic review and meta-analysis, Rejuvenation Res. 18(2015) 3–13. [PubMed] [Google Scholar]
  49. Sebastiani P, Bae H, Gurinovich A, Soerensen M, Puca A, Peris TT, Limitations and risks of meta-analyses of longevity studies, Mech. Ageing Dev165 (2017) 139–146. [PMC free article] [PubMed] [Google Scholar]
  50. Deloukas P, Kanoni S, Willenborg C, Farrall M, Assimes TL, Thompson JR, Ingelsson E, Saleheen D, Erdmann J, Goldstein BA, Stirrups K, Konig IR, Cazier JB, Johansson A, Hall AS, Lee JY, Wilier CJ, Chambers JC, Esko T, Folkersen L, Goel A, Grundberg E, Havulinna AS, Ho WK, Hopewell JC, Eriksson N, Kleber ME, Kristiansson K, Lundmark P, Lyytikainen LP, Rafelt S, Shungin D, Strawbridge RJ, Thorleifsson G, Tikkanen E, Van Zuydam N, Voight BF, Waite LL, Zhang W, Ziegler A, Absher D, Altshuler D, Balmforth AJ, Barroso I, Braund PS, Burgdorf C, Claudi-Boehm S, Cox D, Dimitriou M, Do R, Doney AS, Mokhtari N. El, Eriksson P, Fischer K, Fontanillas P, Franco-Cereceda A, Gigante B, Groop L, Gustafsson S, Hager J, Hallmans G, Han BG, Hunt SE, Kang HM, Illig T, Kessler T, Knowles JW, Kolovou G, Kuusisto J, Langenberg C, Langford C, Leander K, Lokki ML, Lundmark A, McCarthy MI, Meisinger C, Melander O, Mihailov E, Maouche S, Morris AD, Muller-Nurasyid M, Nikus K, Peden JF, Rayner NW, Rasheed A, Rosinger S, Rubin D, Rumpf MP, Schafer A, Sivananthan M, Song C, Stewart AF, Tan ST, Thorgeirsson G, van der Schoot CE, Wagner PJ, Wells GA, Wild PS, Yang TP, Amouyel P, Arveiler D, Basart H, Boehnke M, Boerwinkle E, Brambilla P, Cambien F, Cupples AL, de Faire U, Dehghan A, Diemert P, Epstein SE, Evans A, Ferrario MM, Ferrieres J, Gauguier D, Go AS, Goodall AH, Gudnason V, Hazen SL, Holm H, Iribarren C, Jang Y, Kahonen M, Kee F, Kim HS, Klopp N, Koenig W, Kratzer W, Kuulasmaa K, Laakso M, Laaksonen R, Lee JY, Lind L, Ouwehand WH, Parish S, Park JE, Pedersen NL, Peters A, Quertermous T, Rader DJ, Salomaa V, Schadt E, Shah SH, Sinisalo J, Stark K, Stefansson K, Tregouet DA, Virtamo J, Wallentin L, Wareham N, Zimmermann ME, Nieminen MS, Hengstenberg C, Sandhu MS, Pastinen T, Syvanen AC, Hovingh GK, Dedoussis G, Franks PW, Lehtimaki T, Metspalu A, Zalloua PA, Siegbahn A, Schreiber S, Ripatti S, Blankenberg SS, Perola M, Clarke R, Boehm BO, O'Donnell C, Reilly MP, Marz W, Collins R, Kathiresan S, Hamsten A, Kooner JS, Thorsteinsdottir U, Danesh J, Palmer CN, Roberts R, Watkins H, Schunkert H, Samani NJ, Large-scale association analysis identifies new risk loci for coronary artery disease, Nat. Genet45 (2013) 25–33. [PMC free article] [PubMed] [Google Scholar]
  51. Christiansen MK, Larsen SB, Nyegaard M, Neergaard-Petersen S, Ajjan R, Wurtz M, Grove EL, Hvas AM, Jensen HK, Kristensen SD, Coronary artery disease-associated genetic variants and biomarkers of inflammation, PLoS One12 (2017) e0180365. [PMC free article] [PubMed] [Google Scholar]
  52. Deelen J, Beekman M, Uh HW, Helmer Q, Kuningas M, Christiansen L, Kremer D, van der Breggen R, Suchiman HED, Lakenberg N, van den Akker EB, Passtoors WM, Tiemeier H, van Heemst D, de Craen AJ, Rivadeneira F, de Geus EJ, Perola M, van der Ouderaa FJ, Gunn DA, Boomsma DI, Uitterlinden AG, Christensen K, van Duijn CM, Heijmans BT, Houwing-Duistermaat JJ, Westendorp RGJ, Slagboom PE, Genome-wide association study identifies a single major locus contributing to survival into old age; the APOElocus revisited, Aging Cell 10 (2011) 686–698. [PMC free article] [PubMed] [Google Scholar]
  53. Bekris LM, Lutz F, Yu CE, Functional analysis of APOElocus genetic variation implicates regional enhancers in the regulation of both TOMM40 and APOEJ. Hum. Genet 57 (2012) 18–25. [PMC free article] [PubMed] [Google Scholar]
  54. Middelberg RP, Ferreira MA, Henders AK, Heath AC, Madden PA, Montgomery GW, Martin NG, Whitfield JB, Genetic variants in LPLOASLand TOMM40/APOE-C1-C2-C4 genes are associated with multiple cardiovascular-related traits, BMC Med. Genet 12 (2011) 123. [PMC free article] [PubMed] [Google Scholar]
  55. Wensley F, Gao P, Burgess S, Kaptoge S, Di Angelantonio E, Shah T, Engert JC, Clarke R, Davey-Smith G, Nordestgaard BG, Saleheen D, Samani NJ, Sandhu M, Anand S, Pepys MB, Smeeth L, Whittaker J, Casas JP, Thompson SG, Hingorani AD, Danesh J, Association between C reactive protein and coronary heart disease: mendelian randomisation analysis based on individual participant data, BMJ342 (2011) d548. [PMC free article] [PubMed] [Google Scholar]
  56. Rast P, Kennedy KM, Rodrigue KM, Robinson P, Gross AL, McLaren DG, Grabowski T, Schaie KW, Willis SL, APOEe4 genotype and hypertension modify 8-year cortical thinning: five occasion evidence from the Seattle Longitudinal Study, Cereb. Cortex (New York)(2017) 1–12. [PMC free article] [PubMed] [Google Scholar]
  57. Burggren AC, Mahmood Z, Harrison TM, Siddarth P, Miller KJ, Small GW, Merrill DA, Bookheimer SY, Hippocampal thinning linked to longer TOMM40poly-T variant lengths in the absence of the APOE e4 variant, Alzheimers Dement. 13 (2017) 739–748. [PMC free article] [PubMed] [Google Scholar]
  58. Burkewitz K, Zhang Y, Mair WB, AMPK at the nexus of energetics and aging, Cell Metab. 20(2014) 10–25. [PMC free article] [PubMed] [Google Scholar]
  59. Salminen A, Kaarniranta K, Kauppinen A, Age-related changes in AMPK activation: role for AMPK phosphatases and inhibitory phosphorylation by upstream signaling pathways, Ageing Res. Rev28 (2016) 15–26. [PubMed] [Google Scholar]
  60. Kenyon C, Chang J, Gensch E, Rudner A, Tabtiang R, A C. elegansmutant that lives twice as long as wild-type, Nature 366 (1993) 461–464. [PubMed] [Google Scholar]
  61. Kenyon C, The plasticity of aging: insights from long-lived mutants, Cell120 (2005) 449–460. [PubMed] [Google Scholar]
  62. Morris BJ, A forkhead in the road to longevity: the molecular basis of lifespan becomes clearer. (Review), J. Hypertens. 23(2005) 1285–1309. [PubMed] [Google Scholar]
  63. Martins R, Lithgow GJ, Link W, Long live FOXO: unraveling the role of FOXO proteins in aging and longevity, Aging Cell15 (2016) 196–207. [PMC free article] [PubMed] [Google Scholar]
  64. Lee S, Dong HH, FoxO integration of insulin signaling with glucose and lipid metabolism, J. Endocrinol233 (2017) R67–r79. [PMC free article] [PubMed] [Google Scholar]
  65. Tzivion G, Dobson M, Ramakrishnan G, FoxO transcription factors; regulation by AKT and 14–3-3 proteins, Biochim. Biophys. Acta1813 (2011) 1938–1945. [PubMed] [Google Scholar]
  66. Willcox BJ, Donlon TA, He Q, Chen R, Grove JS, Yano K, Masaki KH, Willcox DC, Rodriguez B, Curb JD, FOXO3A genotype is strongly associated with human longevity, Proc. Natl. Acad. Sci. U. S. A105 (2008) 13987–13992. [PMC free article] [PubMed] [Google Scholar]
  67. Flachsbart F, Caliebe A, Kleindorp R, Blanché H, von Eller-Eberstein H, Nikolaus S, Schreiber S, Nebel A, Association of FOXO3Avariation with human longevity confirmed in German centenarians, Proc. Natl. Acad. Sci. U. S. A 106 (2009) 2700–2705. [PMC free article] [PubMed] [Google Scholar]
  68. Pawlikowska L, Hu D, Huntsman S, Sung A, Chu C, Chen J, Joyner AH, Schork NJ, Hsueh WC, Reiner AP, Psaty BM, Atzmon G, Barzilai N, Cummings SR, Browner WS, Kwok PY, Ziv E, Association of common genetic variation in the insulin/IGF1 signaling pathway with human longevity, Aging Cell8 (2009) 460–472. [PMC free article] [PubMed] [Google Scholar]
  69. Broer L, Buchman AS, Deelen J, Evans DS, Faul JD, Lunetta KL, Sebastiani P, Smith JA, Smith AV, Tanaka T, Yu L, Arnold AM, Aspelund T, Benjamin EJ, De Jager PL, Eirkisdottir G, Evans DA, Garcia ME, Hofman A, Kaplan RC, Kardia SL, Kiel DP, Oostra BA, Orwoll ES, Parimi N, Psaty BM, Rivadeneira F, Rotter JI, Seshadri S, Singleton A, Tiemeier H, Uitterlinden AG, Zhao W, Bandinelli S, Bennett DA, Ferrucci L, Gudnason V, Harris TB, Karasik D, Launer LJ, Perls TT, Slagboom PE, Tranah GJ, Weir DR, Newman AB, van Duijn CM, Murabito JM, GWAS of longevity in CHARGE consortium confirms APOEand FOXO3 candidacy, J. Gerontol. A Biol. Sci. Med. Sci 70 (2014) 110–118. [PMC free article] [PubMed] [Google Scholar]
  70. Anselmi CV, Malovini A, Roncarati R, Novelli V, Villa F, Condorelli G, Bellazzi R, Puca AA, Association of the FOXO3Alocus with extreme longevity in a southern Italian centenarian study, Rejuvenation Res. 12 (2009) 95–104. [PubMed] [Google Scholar]
  71. Li Y, Wang WJ, Cao H, Lu J, Wu C, Hu FY, Guo J, Zhao L, Yang F, Zhang YX, Li W, Zheng GY, Cui H, Chen X, Zhu Z, He H, Dong B, Mo X, Zeng Y, Tian XL, Genetic association of FOXO1Aand FOXO3A with longevity trait in Han Chinese populations, Hum. Mol. Genet 18 (2009) 4897–4904. [PMC free article] [PubMed] [Google Scholar]
  72. Tennapel MJ, Lynch CF, Burns TL, Wallace R, Smith BJ, Button A, Domann FE, SIRT6minor allele genotype is associated with > 5-year decrease in lifespan in an aged cohort, PLoS One 9 (2014) e115616. [PMC free article] [PubMed] [Google Scholar]
  73. Bae H, Gurinovich A, Malovini A, Atzmon G, Andersen SL, Villa F, Barzilai N, Puca A, Peris TT, Sebastiani P, Effects of FOXO3polymorphisms on survival to extreme longevity in four centenarian studies, J. Gerontol. A Biol. Sci. Med. Sci (2017) (Epub ahead of print July 18, 2017). [PMC free article] [PubMed] [Google Scholar]
  74. Joshi PK, Pirastu N, Kentistou KA, Fischer K, Hofer E, Schraut KE, Clark DW, Nutile T, Barnes CLK, Timmers P, Shen X, Gandin I, McDaid AF, Hansen TF, Gordon SD, Giulianini F, Boutin TS, Abdellaoui A, Zhao W, Medina-Gomez C, Bartz TM, Trompet S, Lange LA, Raffield L, van der Spek A, Galesloot TE, Proitsi P, Yanek LR, Bielak LF, Payton A, Murgia F, Concas MP, Biino G, Tajuddin SM, Seppala I, Amin N, Boerwinkle E, Borglum AD, Campbell A, Demerath EW, Demuth I, Faul JD, Ford I, Gialluisi A, Gogele M, Graff M, Hingorani A, Hottenga JJ, Hougaard DM, Hurme MA, Ikram MA, Jylha M, Kuh D, Ligthart L, Lill CM, Lindenberger U, Lumley T, Magi R, Marques-Vidal P, Medland SE, Milani L, Nagy R, Ollier WER, Peyser PA, Pramstaller PP, Ridker PM, Rivadeneira F, Ruggiero D, Saba Y, Schmidt R, Schmidt H, Slagboom PE, Smith BH, Smith JA, Sotoodehnia N, Steinhagen-Thiessen E, van Rooij FJA, Verbeek AL, Vermeulen SH, Vollenweider P, Wang Y, Werge T, Whitfield JB, Zonderman AB, Lehtimaki T, Evans MK, Pirastu M, Fuchsberger C, Bertram L, Pendleton N, Kardia SLR, Ciullo M, Becker DM, Wong A, Psaty BM, van Duijn CM, Wilson JG, Jukema JW, Kiemeney L, Uitterlinden AG, Franceschini N, North KE, Weir DR, Metspalu A, Boomsma DI, Hayward C, Chasman D, Martin NG, Sattar N, Campbell H, Esko T, Kutalik Z, Wilson JF, Genome-wide meta-analysis associates HLA-DQA1/DRB1and LPA and lifestyle factors with human longevity, Nat. Commun 8 (2017) 910. [PMC free article] [PubMed] [Google Scholar]
  75. Bao JM, Song XL, Hong YQ, Zhu HL, Li C, Zhang T, Chen W, Zhao SC, Chen Q, Association between FOXO3Agene polymorphisms and human longevity: a meta-analysis, Asian J. Androl 16 (2014) 446–452. [PMC free article] [PubMed] [Google Scholar]
  76. Sun L, Hu C, Zheng C, Qian Y, Liang Q, Lv Z, Huang Z, Qi K, Gong H, Zhang Z, Huang J, Zhou Q, Yang Z, FOXO3 variants are beneficial for longevity in Southern Chinese living in the Red River Basin: a case-control study and meta-analysis, Sci. Rep5 (2015) 9852. [PMC free article] [PubMed] [Google Scholar]
  77. Druley TE, Wang L, Lin SJ, Lee JH, Zhang Q, Daw EW, Abel HJ, Chasnoff SE, Ramos EI, Levinson BT, Thyagarajan B, Newman AB, Christensen K, Mayeux R, Province MA, Candidate gene resequencing to identify rare, pedigree-specific variants influencing healthy aging phenotypes in the long life family study, BMC Geriatr. 16(2016) 80. [PMC free article] [PubMed] [Google Scholar]
  78. Lin R, Zhang Y, Yan D, Liao X, Wang X, Fu Y, Cai W, Genetic association analysis of common variants in FOXO3related to longevity in a Chinese population, PLoS One 11 (2016) e0167918. [PMC free article] [PubMed] [Google Scholar]
  79. Soerensen M, Dato S, Christensen K, McGue M, Stevnsner T, Bohr VA, Christiansen L, Replication of an association of variation in the FOXO3Agene with human longevity using both case-control and longitudinal data, Aging Cell 9 (2010) 1010–1017. [PMC free article] [PubMed] [Google Scholar]
  80. Donlon TA, Morris BJ, Chen R, Masaki KH, Allsopp RC, Willcox DC, Elliott A, Willcox BJ, FOXO3longevity interactome on chromosome 6, Aging Cell 16 (2017) 1016–1025. [PMC free article] [PubMed] [Google Scholar]
  81. Flachsbart F, Dose J, Gentschew L, Geismann C, Caliebe A, Knecht C, Nygaard M, Badarinarayan N, ElSharawy A, May S, Luzius A, Torres GG, Jentzsch M, Forster M, Hasler R, Pallauf K, Lieb W, Derbois C, Galan P, Drichel A, Arlt A, Till A, Krause-Kyora B, Rimbach G, Blanche H, Deleuze JF, Christiansen L, Christensen K, Nothnagel M, Rosenstiel P, Schreiber S, Franke A, Sebens S, Nebel A, Identification and characterization of two functional variants in the human longevity gene FOXO3Nat. Commun8 (2017) 2063. [PMC free article] [PubMed] [Google Scholar]
  82. Grossi V, Forte G, Sanese P, Peserico A, Tezil T, Signorile L, Fasano C, Bagnulo R, Loconte DC, Susca FC, Resta N, Simone C, The longevity SNP rs2802292 uncovered: HSF1 activates stress-dependent expression of FOXO3 through an incronic enhancer, Nucleic Acids Res. (2018). [PMC free article][PubMed]
  83. Willcox BJ, Tranah GJ, Chen R, Morris BJ, Masaki KH, He Q, Willcox DC, Allsopp RC, Moisyadi S, Poon LW, Rodriguez B, Newman AB, Harris TB, Cummings SR, Liu L, Parimi N, Evans DS, Davy P, Gerschenson M, Donlon TA, The FoxO3 gene and cause-specific mortality, Aging Cell15 (2016) 617–624. [PMC free article] [PubMed] [Google Scholar]
  84. Willcox BJ, Morris BJ, Tranah GJ, Chen R, Masaki KH, He Q, Willcox DC, Allsopp RC, Moisyadi S, Gerschenson M, Davy PMC, Poon LW, Rodriguez B, Newman AB, Harris TB, Cummings SR, Liu Y, Parimi N, Evans DS, Donlon TA, Longevity-associated FOXO3 genotype and its impact on coronary artery disease mortality in Japanese, whites, and blacks: a prospective study of three American populations, J. Gerontol. A Biol. Sci. Med. Sci. 72(2017) 724–728. [PMC free article] [PubMed] [Google Scholar]
  85. Zettergren A, Kern S, Ryden L, Ostling S, Blennow K, Zetterberg H, Falk H, Skoog I, Genetic variation in FOXO3is associated with self-rated health in a population-based sample of older individuals, J. Gerontol. A Biol. Sci. Med. Sci. (2018) (Pub ahead of print Feb 5, 2018). [PMC free article] [PubMed] [Google Scholar]
  86. Williams DM, Karlsson IK, Pedersen NL, Hagg S, Circulating insulin-like growth factors and Alzheimer disease: a Mendelian randomization study, Neurology90 (2018) e291–e297. [PMC free article] [PubMed] [Google Scholar]
  87. Davy PMC, Willcox DC, Shimabukuro M, Donlon TA, Torigoe T, Suzuki M, Higa M, Masuzaki H, Sata M, Chen R, Murkofsky R, Morris BJ, Lim E, Allsopp RC, Willcox BJ, Minimal shortening of leukocyte telomere length across age groups in a cross-sectional study for carriers of a longevity-associated FOXO3 allele, J. Gerontol. A Biol. Sci. Med. Sci. (2018) (Pub ahead of print Apr 21, 2018). [PMC free article][PubMed] [Google Scholar]
  88. Sniekers S, Stringer S, Watanabe K, Jansen PR, Coleman JRI, Krapohl E, Taskesen E, Hammerschlag AR, Okbay A, Zabaneh D, Amin N, Breen G, Cesarini D, Chabris CF, Iacono WG, Ikram MA, Johannesson M, Koellinger P, Lee JJ, Magnusson PKE, McGue M, Miller MB, Ollier WER, Payton A, Pendleton N, Plomin R, Rietveld CA, Tiemeier H, van Duijn CM, Posthuma D, Genome-wide association meta-analysis of 78,308 individuals identifies new loci and genes influencing human intelligence, Nat. Genet49 (2017) 1107–1112. [PMC free article] [PubMed] [Google Scholar]
  89. Nebel A, Kleindorp R, Caliebe A, Nothnagel M, Blanche H, Junge O, Wittig M, Ellinghaus D, Flachsbart F, Wichmann HE, Meitinger T, Nikolaus S, Franke A, Krawczak M, Lathrop M, Schreiber S, A genome-wide association study confirms APOE as the major gene influencing survival in long-lived individuals, Mech. Ageing Dev132 (2011) 324–330. [PubMed] [Google Scholar]
  90. Sebastiani P, Solovieff N, Dewan AT, Walsh KM, Puca A, Hartley SW, Melista D, Andersen S, Dworkis DA, Wilk JB, Myers RH, Steinberg MH, Montano M, Baldwin CT, Hoh J, Peris TT, Genetic signatures of exceptional longevity in humans, PLoS One7 (2012) e29848. [PMC free article] [PubMed] [Google Scholar]
  91. Beekman M, Blanche H, Perola M, Hervonen A, Bezrukov V, Sikora E, Flachsbart F, Christiansen L, De Craen AJ, Kirkwood TB, Rea IM, Poulain M, Robine JM, Valensin S, Stazi MA, Passarino G, Deiana L, Gonos ES, Paternoster L, Sorensen TI, Tan Q, Helmer Q, van den Akker EB, Deelen J, Martella F, Cordell HJ, Ayers KL, Vaupel JW, Tornwall O, Johnson TE, Schreiber S, Lathrop M, Skytthe A, Westendorp RG, Christensen K, Gampe J, Nebel A, Houwing-Duistermaat JJ, Slagboom PE, Franceschi C, G. consortium, Genome-wide linkage analysis for human longevity: genetics of healthy aging study, Aging Cell12 (2013) 184–193. [PMC free article] [PubMed] [Google Scholar]
  92. Deelen J, Beekman M, Uh HW, Broer L, Ayers KL, Tan Q, Kamatani Y, Bennet AM, Tamm R, Trompet S, Guethbjartsson DF, Flachsbart F, Rose G, Viktorin A, Fischer K, Nygaard M, Cordell HJ, Crocco P, van den Akker EB, Bohringer S, Helmer Q, Nelson CP, Saunders GI, Alver M, Andersen-Ranberg K, Breen ME, van der Breggen R, Caliebe A, Capri M, Cevenini E, Collerton JC, Dato S, Davies K, Ford I, Gampe J, Garagnani P, de Geus EJ, Harrow J, van Heemst D, Heijmans BT, Heinsen FA, Hottenga JJ, Hofman A, Jeune B, Jonsson PV, Lathrop M, Lechner D, Martin-Ruiz C, McNerlan SE, Mihailov E, Montesanto A, Mooijaart SP, Murphy A, Nohr EA, Paternoster L, Postmus I, Rivadeneira F, Ross OA, Salvioli S, Sattar N, Schreiber S, Stefansson H, Stott DJ, Tiemeier H, Uitterlinden AG, Westendorp RG, Willemsen G, Samani NJ, Galan P, Sorensen TI, Boomsma DI, Jukema JW, Rea IM, Passarino G, de Craen AJ, Christensen K, Nebel A, Stefansson K, Metspalu A, Magnusson P, Blanche H, Christiansen L, Kirkwood TB, van Duijn CM, Franceschi C, Houwing-Duistermaat JJ, Slagboom PE, Genome-wide association meta-analysis of human longevity identifies a novel locus conferring survival beyond 90 years of age, Hum. Mol. Genet23 (2014) 4420–4432. [PMC free article] [PubMed] [Google Scholar]
  93. Broer L, Buchman AS, Deelen J, Evans DS, Faul JD, Lunetta KL, Sebastiani P, Smith JA, Smith AV, Tanaka T, Yu L, Arnold AM, Aspelund T, Benjamin EJ, De Jager PL, Eirkisdottir G, Evans DA, Garcia ME, Hofman A, Kaplan RC, Kardia SL, Kiel DP, Oostra BA, Orwoll ES, Parimi N, Psaty BM, Rivadeneira F, Rotter JI, Seshadri S, Singleton A, Tiemeier H, Uitterlinden AG, Zhao W, Bandinelli S, Bennett DA, Ferrucci L, Gudnason V, Harris TB, Karasik D, Launer LJ, Peris TT, Slagboom PE, Tranah GJ, Weir DR, Newman AB, van Duijn CM, Murabito JM, GWAS of longevity in CHARGE consortium confirms APOEand FOXO3 candidacy, J. Gerontol. A Biol. Sci. Med. Sci 70 (2015) 110–118. [PMC free article] [PubMed] [Google Scholar]
  94. Fortney K, Dobriban E, Garagnani P, Pirazzini C, Monti D, Mari D, Atzmon G, Barzilai N, Franceschi C, Owen AB, Kim SK, Genome-wide scan informed by age-related disease identifies loci for exceptional human longevity, PLoS Genet. 11(2015) e1005728. [PMC free article] [PubMed] [Google Scholar]
  95. Joshi PK, Fischer K, Schraut KE, Campbell H, Esko T, Wilson JF, Variants near CHRNA3/5and APOE have age- and sex-related effects on human lifespan, Nat. Commun 7 (2016) 11174. [PMC free article] [PubMed] [Google Scholar]
  96. McDaid AF, Joshi PK, Porcu E, Komljenovic A, Li H, Sorrentino V, Litovchenko M, Bevers RPJ, Rueger S, Reymond A, Bochud M, Deplancke B, Williams RW, Robinson-Rechavi M, Paccaud F, Rousson V, Auwerx J, Wilson JF, Kutalik Z, Bayesian association scan reveals loci associated with human lifespan and linked biomarkers, Nat. Commun8 (2017) 15842. [PMC free article] [PubMed] [Google Scholar]
  97. Sebastiani P, Gurinovich A, Bae H, Andersen S, Malovini A, Atzmon G, Villa F, Kraja AT, Ben-Avraham D, Barzilai N, Puca A, Perls TT, Four genome-wide association studies identify new extreme longevity variants, J. Gerontol. A Biol. Sci. Med. Sci72 (2017) 1453–1464. [PMC free article] [PubMed] [Google Scholar]
  98. Pilling LC, Kuo CL, Sicinski K, Tamosauskaite J, Kuchel GA, Harries LW, Herd P, Wallace R, Ferrucci L, Melzer D, Human longevity: 25 genetic loci associated in 389,166 UK biobank participants, Aging9 (2017) 2504–2520. [PMC free article] [PubMed] [Google Scholar]
  99. Marom S, Friger M, Mishmar D, MtDNA meta-analysis reveals both phenotype specificity and allele heterogeneity: a model for differential association, Sci. Rep7 (2017) 43449. [PMC free article] [PubMed] [Google Scholar]
  100. Singh J, Minster RL, Schupf N, Kraja A, Liu Y, Christensen K, Newman AB, Kammerer CM, Genomewide association scan of a mortality associated endophenotype for a long and healthy life in the Long Life Family Study, J. Gerontol. A Biol. Sci. Med. Sci72 (2017) 1411–1416. [PMC free article] [PubMed] [Google Scholar]
  101. Boyden SE, Kunkel LM, High-density genomewide linkage analysis of exceptional human longevity identifies multiple novel loci, PLoS One5 (2010) e12432. [PMC free article] [PubMed] [Google Scholar]
  102. Puca AA, Daly MJ, Brewster SJ, Matise TC, Barrett J, Shea-Drinkwater M, Kang S, Joyce E, Nicoli J, Benson E, Kunkel LM, Peris T, A genome-wide scan for linkage to human exceptional longevity identifies a locus on chromosome 4, Proc. Natl. Acad. Sci. U. S. A98 (2001) 10505–11058. [PMC free article] [PubMed] [Google Scholar]
  103. Edwards DR, Gilbert JR, Jiang L, Gallins PJ, Caywood L, Creason M, Fuzzell D, Knebusch C, Jackson CE, Pericak-Vance MA, Haines JL, Scott WK, Successful aging shows linkage to chromosomes 6, 7, and 14 in the Amish, Ann. Hum. Genet75 (2011) 516–528. [PMC free article] [PubMed] [Google Scholar]
  104. Edwards DR, Gilbert JR, Hicks JE, Myers JL, Jiang L, Cummings AC, Guo S, Gallins PJ, Konidari I, Caywood L, Reinhart-Mercer L, Fuzzell D, Knebusch C, Laux R, Jackson CE, Pericak-Vance MA, Haines JL, Scott WK, Linkage and association of successful aging to the 6q25 region in large Amish kindreds, Age (Dordr.)35 (2013) 1467–1477. [PMC free article] [PubMed] [Google Scholar]
  105. Pilling LC, Atkins JL, Bowman K, Jones SE, Tyrrell J, Beaumont RN, Ruth KS, Tuke MA, Yaghootkar H, Wood AR, Freathy RM, Murray A, Weedon MN, Xue L, Lunetta K, Murabito JM, Harries LW, Robine JM, Brayne C, Kuchel GA, Ferrucci L, Frayling TM, Melzer D, Human longevity is influenced by many genetic variants: evidence from 75,000 UK Biobank participants, Aging8 (2016) 547–560. [PMC free article] [PubMed] [Google Scholar]
  106. Zeng Y, Nie C, Min J, Liu X, Li M, Chen H, Xu H, Wang M, Ni T, Li Y, Yan H, Zhang JP, Song C, Chi LQ, Wang HM, Dong J, Zheng GY, Lin L, Qian F, Qi Y, Liu X, Cao H, Wang Y, Zhang L, Li Z, Zhou Y, Wang Y, Lu J, Li J, Qi M, Bolund L, Yashin A, Land KC, Gregory S, Yang Z, Gottschalk W, Tao W, Wang J, Wang J, Xu X, Bae H, Nygaard M, Christiansen L, Christensen K, Franceschi C, Lutz MW, Gu J, Tan Q, Peris T, Sebastiani P, Deelen J, Slagboom E, Hauser E, Xu H, Tian XL, Yang H, Vaupel JW, Novel loci and pathways significantly associated with longevity, Sci. Rep6 (2016) 21243. [PMC free article] [PubMed] [Google Scholar]
  107. Minster RL, Sanders JL, Singh J, Kammerer CM, Barmada MM, Matteini AM, Zhang Q, Wojczynski MK, Daw EW, Brody JA, Arnold AM, Lunetta KL, Murabito JM, Christensen K, Peris TT, Province MA, Newman AB, Genome-wide association study and linkage analysis of the healthy aging index, J. Gerontol. A Biol. Sci. Med. Sci70 (2015) 1003–1008. [PMC free article] [PubMed] [Google Scholar]
  108. Tanaka T, Dutta A, Pilling LC, Xue L, Lunetta KL, Murabito JM, Bandinelli S, Wallace R, Melzer D, Ferrucci L, Genome-wide association study of parental life span, J. Gerontol. A Biol. Sci. Med. Sci72 (2017) 1407–1410. [PMC free article] [PubMed] [Google Scholar]
  109. Hook M, Roy S, Williams EG, Bou Sleiman M, Mozhui K, Nelson JF, Lu L, Auwerx J, Williams RW, Genetic cartography of longevity in humans and mice: current landscape and horizons, Biochim. Biophys. Acta1864 (9 Pt A) (2018) 2718–2732. [PMC free article] [PubMed] [Google Scholar]
  110. Flachsbart F, Ellinghaus D, Gentschew L, Heinsen FA, Caliebe A, Christiansen L, Nygaard M, Christensen K, Blanche H, Deleuze JF, Derbois C, Galan P, Buning C, Brand S, Peters A, Strauch K, Muller-Nurasyid M, Hoffmann P, Nothen MM, Lieb W, Franke A, Schreiber S, Nebel A, Immunochip analysis identifies association of the RAD50/IL13 region with human longevity, Aging Cell15 (2016) 585–588. [PMC free article] [PubMed] [Google Scholar]
  111. Newman AB, Walter S, Lunetta KL, Garcia ME, Slagboom PE, Christensen K, Arnold AM, Aspelund T, Aulchenko YS, Benjamin EJ, Christiansen L, D'Agostino RB Sr., Fitzpatrick AL, Franceschini N, Glazer NL, Gudnason V, Hofman A, Kaplan R, Karasik D, Kelly-Hayes M, Kiel DP, Launer LJ, Marciante KD, Massaro JM, Miljkovic I, Nalls MA, Hernandez D, Psaty BM, Rivadeneira F, Rotter J, Seshadri S, Smith AV, Taylor KD, Tiemeier H, Uh HW, Uitterlinden AG, Vaupel JW, Walston J, Westendorp RG, Harris TB, Lumley T, van Duijn CM, Murabito JM, A meta-analysis of four genome-wide association studies of survival to age 90 years or older: the Cohorts for Heart and Aging Research in Genomic Epidemiology Consortium, J. Gerontol. A Biol. Sci. Med. Sci65 (2010) 478–487. [PMC free article] [PubMed] [Google Scholar]
  112. Nygaard HB, Erson-Omay EZ, Wu X, Kent BA, Bemales CQ, Evans DM, Farrer J, Vilarino-Guell C, Strittmatter SM, Whole exome sequencing of an exceptional longevity cohort, J. Gerontol. A Biol. Sci. Med. Sci(2018) (Pub ahead of print May 10, 2018).
  113. Nygaard M, Debrabant B, Tan Q, Deelen J, Andersen-Ranberg K, de Craen AJ, Beekman M, Jeune B, Slagboom PE, Christensen K, Christiansen L, Copy number variation associates with mortality in long-lived individuals: a genome-wide assessment, Aging Cell15 (2016) 49–55. [PMC free article] [PubMed] [Google Scholar]
  114. Bediaga NG, Aznar JM, Elcoroaristizabal X, Alboniga O, Gomez-Busto F, Artaza Artabe I, Rocandio A, de Pancorbo MM, Associations between STR autosomal markers and longevity, Age (Dordr.)37 (2015) 95. [PMC free article] [PubMed] [Google Scholar]
  115. Johnson SC, Dong X, Vijg J, Suh Y, Genetic evidence for common pathways in human age-related diseases, Aging Cell14 (2015) 809–817. [PMC free article] [PubMed] [Google Scholar]
  116. Hirst J, Edgar JR, Esteves T, Darios F, Madeo M, Chang J, Roda RH, Durr A, Anheim M, Gellera C, Li J, Zuchner S, Mariotti C, Stevanin G, Blackstone C, Kruer MC, Robinson MS, Loss of AP-5 results in accumulation of aberrant endolysosomes: defining a new type of lysosomal storage disease, Hum. Mol. Genet24 (2015) 4984–4996. [PMC free article] [PubMed] [Google Scholar]
  117. Welter D, MacArthur J, Morales J, Burdett T, Hall P, Junkins H, Klemm A, Flicek P, Manolio T, Hindorff L, Parkinson H, The NHGRI GWAS Catalog, a curated resource of SNP-trait associations, Nucleic Acids Res. 42(2014) D1001–D1006. [PMC free article] [PubMed] [Google Scholar]
  118. Khan A, Prasanth SG, BEND3 mediates transcriptional repression and heterochromatin organization, Transcription6 (2015) 102–105. [PMC free article] [PubMed] [Google Scholar]
  119. Aguilo F, Di Cecilia S, Walsh MJ, Long non-coding RNA ANRIL and Polycomb in human cancers and cardiovascular disease, Curr. Top. Microbiol. Immunol394 (2016) 29–39. [PMC free article] [PubMed] [Google Scholar]
  120. Baker DJ, Wijshake T, Tchkonia T, LeBrasseur NK, Childs BG, van de Sluis B, Kirkland JL, van Deursen JM, Clearance of p16Ink4a-positive senescent cells delays ageing-associated disorders, Nature479 (2011) 232–236. [PMC free article] [PubMed] [Google Scholar]
  121. Congrains A, Kamide K, Oguro R, Yasuda O, Miyata K, Yamamoto E, Kawai T, Kusunoki H, Yamamoto H, Takeya Y, Yamamoto K, Onishi M, Sugimoto K, Katsuya T, Awata N, Ikebe K, Gondo Y, Oike Y, Ohishi M, Rakugi H, Genetic variants at the 9p21 locus contribute to atherosclerosis through modulation of ANRIL and CDKN2A/BAtherosclerosis220 (2012) 449–455. [PubMed] [Google Scholar]
  122. Congrains A, Kamide K, Hirose N, Arai Y, Oguro R, Nakama C, Imaizumi Y, Kawai T, Kusunoki H, Yamamoto H, Onishi-Takeya M, Takeya Y, Yamamoto K, Sugimoto K, Akasaka H, Saitoh S, Miura T, Awata N, Kato N, Katsuya T, Ikebe K, Gondo Y, Rakugi H, Disease-associated polymorphisms in 9p21 are not associated with extreme longevity, Geriatr. Gerontol. Int15 (2015) 797–803. [PubMed] [Google Scholar]
  123. Qin B, Minter-Dykhouse K, Yu J, Zhang J, Liu T, Zhang H, Lee S, Kim J, Wang Z Lou, DBC1 functions as a tumor suppressor by regulating p53 stability, Cell Rep. 10(2015) 1324–1334. [PMC free article] [PubMed] [Google Scholar]
  124. Zhang J, Wang C, Chen X, Takada M, Fan C, Zheng X, Wen H, Liu Y, Wang C, Pestell RG, Aird KM, Kaelin WG Jr., Liu XS, Zhang Q, EglN2 associates with the NRF1-PGC1alpha complex and controls mitochondrial function in breast cancer, EMBO J. 34(2015) 2953–2970. [PMC free article] [PubMed] [Google Scholar]
  125. Matsuzaka T, Shimano H, Elovl6: a new player in fatty acid metabolism and insulin sensitivity, J. Mol. Med. (Berlin, Germany)87 (2009) 379–384. [PubMed] [Google Scholar]
  126. Sato K, Emi M, Ezura Y, Fujita Y, Takada D, Ishigami T, Umemura S, Xin Y, Wu LL, Larrinaga-Shum S, Stephenson SH, Hunt SC, Hopkins PN, Soluble epoxide hydrolase variant (Glu287Arg) modifies plasma total cholesterol and triglyceride phenotype in familial hypercholesterolemia: intrafamilial association study in an eight-generation hyperlipidemic kindred, J. Hum. Genet49 (2004) 29–34. [PubMed] [Google Scholar]
  127. Wei Q, Doris PA, Pollizotto MV, Boerwinkle E, Jacobs DR Jr., Siscovick DS, Fornage M, Sequence variation in the soluble epoxide hydrolase gene and sub-clinical coronary atherosclerosis: interaction with cigarette smoking, Atherosclerosis190 (2007) 26–34. [PubMed] [Google Scholar]
  128. Pesu M, Watford WT, Wei L, Xu L, Fuss I, Strober W, Andersson J, Shevach EM, Quezado M, Bouladoux N, Roebroek A, Belkaid Y, Creemers J, O'Shea JJ, T-cell-expressed proprotein convertase furin is essential for maintenance of peripheral immune tolerance, Nature455 (2008) 246–250. [PMC free article] [PubMed] [Google Scholar]
  129. Ren K, Jiang T, Zheng XL, Zhao GJ, Proprotein convertase furin/PCSK3 and atherosclerosis: new insights and potential therapeutic targets, Atherosclerosis262 (2017) 163–170. [PubMed] [Google Scholar]
  130. Turpeinen H, Seppala I, Lyytikainen LP, Raitoharju E, Hutri-Kahonen N, Levula M, Oksala N, Waldenberger M, Klopp N, Illig T, Mononen N, Laaksonen R, Raitakari O, Kahonen M, Lehtimaki T, Pesu M, A genome-wide expression quantitative trait loci analysis of proprotein convertase subtilisin/kexin enzymes identifies a novel regulatory gene variant for FURIN expression and blood pressure, Hum. Genet134 (2015) 627–636. [PubMed] [Google Scholar]
  131. Karim MA, Suzuki K, Fukai K, Oh J, Nagle DL, Moore KJ, Barbosa E, Falik-Borenstein T, Filipovich A, Ishida Y, Kivrikko S, Klein C, Kreuz F, Levin A, Miyajima H, Regueiro J, Russo C, Uyama E, Vierimaa O, Spritz RA, Apparent genotype-phenotype correlation in childhood, adolescent, and adult Chediak-Higashi syndrome, Am. J. Med. Genet108 (2002) 16–22. [PubMed] [Google Scholar]
  132. Kennedy BK, Berger SL, Brunet A, Campisi J, Cuervo AM, Epel ES, Franceschi C, Lithgow GJ, Morimoto RI, Pessin JE, Rando TA, Richardson A, Schadt EE, Wyss-Coray T, Sierra F, Geroscience: linking aging to chronic disease, Cell159 (2014) 709–713. [PMC free article] [PubMed] [Google Scholar]
  133. Meimaridou E, Hughes CR, Kowalczyk J, Chan LF, Clark AJ, Metherell LA, ACTH resistance: genes and mechanisms, Endocr. Dev24 (2013) 57–66. [PubMed] [Google Scholar]
  134. Erraji-Benchekroun L, Underwood MD, Arango V, Galfalvy H, Pavlidis P, Smyrniotopoulos P, Mann JJ, Sibille E, Molecular aging in human prefrontal cortex is selective and continuous throughout adult life, Biol. Psychiatry57 (2005) 549–558. [PubMed] [Google Scholar]
  135. Barrell D, Dimmer E, Huntley RP, Binns D, O'Donovan C, Apweiler R, The GOA database in 2009–an integrated gene ontology annotation resource, Nucleic Acids Res. 37(2009) D396–D403. [PMC free article] [PubMed] [Google Scholar]
  136. Dale BL, Madhur MS, Linking inflammation and hypertension via LNK/SH2B3Curr. Opin. Nephrol. Hypertens25 (2016) 87–93. [PMC free article] [PubMed] [Google Scholar]
  137. Stolfi C, De Simone V, Colantoni A, Franze E, Ribichini E, Fantini MC, Caruso R, Monteleone I, Sica GS, Sileri P, MacDonald TT, Pallone F, Monteleone G, A functional role for Smad7 in sustaining colon cancer cell growth and survival, Cell Death Dis. 5(2014) e1073. [PMC free article] [PubMed] [Google Scholar]
  138. Lam M, Trampush JW, Yu J, Knowles E, Davies G, Liewald DC, Starr JM, Djurovic S, Melle I, Sundet K, Christoforou A, Reinvang I, DeRosse P, Lundervold AJ, Steen VM, Espeseth T, Raikkonen K, Widen E, Palotie A, Eriksson JG, Giegling I, Konte B, Roussos P, Giakoumaki S, Burdick KE, Payton A, Ollier W, Chiba-Falek O, Attix DK, Need AC, Cirulli ET, Voineskos AN, Stefanis NC, Avramopoulos D, Hatzimanolis A, Arking DE, Smyrnis N, Bilder RM, Freimer NA, Cannon TD, London E, Poldrack RA, Sabb FW, Congdon E, Conley ED, Scult MA, Dickinson D, Straub RE, Donohoe G, Morris D, Corvin A, Gill M, Hariri AR, Weinberger DR, Pendleton N, Bitsios P, Rujescu D, Lahti J, Le Hellard S, Keller MC, Andreassen OA, Deary IJ, Glahn DC, Malhotra AK, Lencz T, Large-scale cognitive GWAS meta-analysis reveals tissue-specific neural expression and potential nootropic drug targets, Cell Rep. 21(2017) 2597–2613. [PMC free article] [PubMed] [Google Scholar]
  139. Okbay A, Beauchamp JP, Fontana MA, Lee JJ, Pers TH, Rietveld CA, Turley P, Chen GB, Emilsson V, Meddens SF, Oskarsson S, Pickrell JK, Thom K, Timshel P, de Vlaming R, Abdellaoui A, Ahluwalia TS, Bacelis J, Baumbach C, Bjornsdottir G, Brandsma JH, Pina Concas M, Derringer J, Furlotte NA, Galesloot TE, Girotto G, Gupta R, Hall LM, Harris SE, Hofer E, Horikoshi M, Huffman JE, Kaasik K, Kalafati IP, Karlsson R, Kong A, Lahti J, van der Lee SJ, de Leeuw C, Lind PA, Lindgren KO, Liu T, Mangino M, Marten J, Mihailov E, Miller MB, van der Most PJ, Oldmeadow C, Payton A, Pervjakova N, Peyrot WJ, Qian Y, Raitakari O, Rueedi R, Salvi E, Schmidt B, Schraut KE, Shi J, Smith AV, Poot RA, St Pourcain B, Teumer A, Thorleifsson G, Verweij N, Vuckovic D, Wellmann J, Westra HJ, Yang J, Zhao W, Zhu Z, Alizadeh BZ, Amin N, Bakshi A, Baumeister SE, Biino G, Bonnelykke K, Boyle PA, Campbell H, Cappuccio FP, Davies G, De Neve JE, Deloukas P, Demuth I, Ding J, Eibich P, Eisele L, Eklund N, Evans DM, Paul JD, Feitosa MF, Forstner AJ, Gandin I, Gunnarsson B, Halldorsson BV, Harris TB, Heath AC, Hocking LJ, Holliday EG, Homuth G, Horan MA, Hottenga JJ, de Jager PL, Joshi PK, Jugessur A, Kaakinen MA, Kahonen M, Kanoni S, Keltigangas-Jarvinen L, Kiemeney LA, Kolcic I, Koskinen S, Kraja AT, Kroh M, Kutalik Z, Latvala A, Launer LJ, Lebreton MP, Levinson DF, Lichtenstein P, Lichtner P, Liewald DC, Loukola A, Madden PA, Magi R, Maki-Opas T, Marioni RE, Marques-Vidal P, Meddens GA, McMahon G, Meisinger C, Meitinger T, Milaneschi Y, Milani L, Montgomery GW, Myhre R, Nelson CP, Nyholt DR, Ollier WE, Palotie A, Paternoster L, Pedersen NL, Petrovic KE, Porteous DJ, Raikkonen K, Ring SM, Robino A, Rostapshova O, Rudan I, Rustichini A, Salomaa V, Sanders AR, Sarin AP, Schmidt H, Scott RJ, Smith BH, Smith JA, Staessen JA, Steinhagen-Thiessen E, Strauch K, Terracciano A, Tobin MD, Ulivi S, Vaccargiu S, Quaye L, van Rooij FJ, Venturing C, Vinkhuyzen AA, Volker U, Volzke H, Vonk JM, Vozzi D, Waage J, Ware EB, Willemsen G, Attia JR, Bennett DA, Berger K, Bertram L, Bisgaard H, Boomsma DI, Borecki IB, Bultmann U, Chabris CF, Cucca F, Cusi D, Deary IJ, Dedoussis GV, van Duijn CM, Eriksson JG, Franke B, Franke L, Gasparini P, Gejman PV, Gieger C, Grabe HJ, Gratten J, Groenen PJ, Gudnason V, van der Harst P, Hayward C, Hinds DA, Hoffmann W, Hypponen E, Iacono WG, Jacobsson B, Jarvelin MR, Jockel KH, Kaprio J, Kardia SL, Lehtimaki T, Lehrer SF, Magnusson PK, Martin NG, McGue M, Metspalu A, Pendleton N, Penninx BW, Perola M, Pirastu N, Pirastu M, Polasek O, Posthuma D, Power C, Province MA, Samani NJ, Schlessinger D, Schmidt R, Sorensen TI, Spector TD, Stefansson K, Thorsteinsdottir U, Thurik AR, Timpson NJ, Tiemeier H, Tung JY, Uitterlinden AG, Vitart V, Vollenweider P, Weir DR, Wilson JF, Wright AF, Conley DC, Krueger RF, Davey Smith G, Hofman A, Laibson DI, Medland SE, Meyer MN, Yang J, Johannesson M, Visscher PM, Esko T, Koellinger PD, Cesarini D, Benjamin DJ, Genome-wide association study identifies 74 loci associated with educational attainment, Nature533 (2016) 539–542. [PMC free article] [PubMed] [Google Scholar]
  140. Marioni RE, Ritchie SJ, Joshi PK, Hagenaars SP, Okbay A, Fischer K, Adams MJ, Hill WD, Davies G, Nagy R, Amador C, Lall K, Metspalu A, Liewald DC, Campbell A, Wilson JF, Hayward C, Esko T, Porteous DJ, Gale CR, Deary IJ, Genetic variants linked to education predict longevity, Proc. Natl. Acad. Sci. U. S. A113 (2016) 13366–13371. [PMC free article] [PubMed] [Google Scholar]
  141. Hasler R, Venkatesh G, Tan Q, Flachsbart F, Sinha A, Rosenstiel P, Lieb W, Schreiber S, Christensen K, Christiansen L, Nebel A, Genetic interplay between human longevity and metabolic pathways - a large-scale eQTL study, Aging Cell16 (2017) 716–725. [PMC free article] [PubMed] [Google Scholar]
  142. Yashin AI, Zhbannikov I, Arbeeva L, Arbeev KG, Wu D, Akushevich I, Yashkin A, Kovtun M, Kulminski AM, Stallard E, Kulminskaya I, Ukraintseva S, Pure and confounded effects of causal SNPs on longevity: insights for proper interpretation of research findings in GWAS of populations with different genetic structures, Front. Genet7 (2016) 188. [PMC free article] [PubMed] [Google Scholar]
  143. Yashin AI, Wu D, Arbeeva LS, Arbeev KG, Kulminski AM, Akushevich I, Kovtun M, Culminskaya I, Stallard E, Li M, Ukraintseva SV, Genetics of aging, health, and survival: dynamic regulation of human longevity related traits, Front. Genet6 (2015) 122. [PMC free article] [PubMed] [Google Scholar]
  144. Willcox BJ, Willcox DC, Caloric restriction, caloric restriction mimetics, and healthy aging in Okinawa: controversies and clinical implications, Curr. Opin. Clin. Nutr. Metab. Care17 (2014) 51–58. [PMC free article] [PubMed] [Google Scholar]
  145. Ravussin E, Redman LM, Rochon J, Das SK, Fontana L, Kraus WE, Romashkan S, Williamson DA, Meydani SN, Villareal DT, Smith SR, Stein RI, Scott TM, Stewart TM, Saltzman E, Klein S, Bhapkar M, Martin CK, Gilhooly CH, Holloszy JO, Hadley EC, Roberts SB, A 2-year randomized controlled trial of human caloric restriction: feasibility and effects on predictors of health span and longevity, J. Gerontol. A Biol. Sci. Med. Sci70 (2015) 1097–1104. [PMC free article] [PubMed] [Google Scholar]
  146. Liu F, Hamer MA, Deelen J, Lall JS, Jacobs L, van Heemst D, Murray PG, Wollstein A, de Craen AJ, Uh HW, Zeng C, Hofman A, Uitterlinden AG, Houwing-Duistermaat JJ, Pardo LM, Beekman M, Slagboom PE, Nijsten T, Kayser M, Gunn DA, The MC1Rgene and youthful looks, Curr. Biol 26 (2016) 1213–1220. [PubMed] [Google Scholar]
  147. Fontana L, Partridge L, Longo VD, Extending healthy life span–from yeast to humans, Science328 (2010) 321–326. [PMC free article] [PubMed] [Google Scholar]
  148. Teumer A, Qi Q, Nethander M, Aschard H, Bandinelli S, Beekman M, Berndt SI, Bidlingmaier M, Broer L, Cappola A, Ceda GP, Chanock S, Chen MH, Chen TC, Chen YD, Chung J, Del Greco Miglianico F, Eriksson J, Ferrucci L, Friedrich N, Gnewuch C, Goodarzi MO, Grarup N, Guo T, Hammer E, Hayes RB, Hicks AA, Hofman A, Houwing-Duistermaat JJ, Hu F, Hunter DJ, Husemoen LL, Isaacs A, Jacobs KB, Janssen JA, Jansson JO, Jehmlich N, Johnson S, Juul A, Karlsson M, Kilpelainen TO, Kovacs P, Kraft P, Li C, Linneberg A, Liu Y, Loos RJ, Lorentzon M, Lu Y, Maggio M, Magi R, Meigs J, Mellstrom D, Nauck M, Newman AB, Poliak MN, Pramstaller PP, Prokopenko I, Psaty BM, Reincke M, Rimm EB, Rotter JI, Saint Pierre A, Schumann C, Seshadri S, Sjogren K, Slagboom PE, Strickler HD, Stumvoll M, Suh Y, Sun Q, Zhang C, Svensson J, Tanaka T, Tare A, Tonjes A, Uh HW, van Duijn CM, van Heemst D, Vandenput L, Vasan RS, Volker U, Willems SM, Ohlsson C, Wallaschofski H, Kaplan RC, Genomewide meta-analysis identifies loci associated with IGF-I and IGFBP-3 levels with impact on age-related traits, Aging Cell15 (2016) 811–824. [PMC free article] [PubMed] [Google Scholar]
  149. Morris BJ, Donlon TA, He Q, Grove JS, Masaki KH, Elliott A, Willcox DC, Willcox BJ, Association analyses of insulin signaling pathway gene polymorphisms with healthy aging and longevity in Americans of Japanese ancestry, J. Gerontol. A Biol. Sci. Med. Sci69 (2014) 270–273. [PMC free article] [PubMed] [Google Scholar]
  150. Tanisawa K, Hirose N, Arai Y, Shimokata H, Yamada Y, Kawai H, Kojima M, Obuchi S, Hirano H, Suzuki H, Fujiwara Y, Taniguchi Y, Shinkai S, Ihara K, Sugaya M, Higuchi M, Arai T, Mori S, Sawabe M, Sato N, Muramatsu M, Tanaka M, Inverse association between height-increasing alleles and extreme longevity in Japanese women, J. Gerontol. A Biol. Sci. Med. Sci73 (2018) 588–595. [PubMed] [Google Scholar]
  151. Ben-Avraham D, Govindaraju DR, Budagov T, Fradin D, Durda P, Liu B, Ott S, Gutman D, Sharvit L, Kaplan R, Bougneres P, Reiner A, Shuldiner AR, Cohen P, Barzilai N, Atzmon G, The GH receptor exon 3 deletion is a marker of male-specific exceptional longevity associated with increased GH sensitivity and taller stature, Sci. Adv3 (2017) e1602025. [PMC free article] [PubMed] [Google Scholar]
  152. He Q, Morris BJ, Grove JS, Petrovitch H, Ross W, Masaki KH, Rodriguez B, Chen R, Donlon TA, Willcox DC, Willcox BJ, Shorter men live longer: association of height with longevity and FOXO3genotype in American men of Japanese ancestry, PLoS One 9 (2014) e94385. [PMC free article] [PubMed] [Google Scholar]
  153. Johnson SC, Rabinovitch PS, Kaeberlein M, mTOR is a key modulator of ageing and age-related disease, Nature493 (2013) 338–345. [PMC free article] [PubMed] [Google Scholar]
  154. Morris BJ, Donlon TA, He Q, Grove JS, Masaki KH, Elliott A, Willcox DC, Allsopp R, Willcox BJ, Genetic analysis of TOR complex gene variation with human longevity: a nested case-control study of American men of Japanese ancestry, J. Gerontol. A Biol. Sci. Med. Sci70 (2015) 133–142. [PMC free article] [PubMed] [Google Scholar]
  155. Donlon TA, Morris BJ, Chen R, Masaki KH, Allsopp RC, Willcox DC, Tiirikainen M, Willcox BJ, Analysis of polymorphisms in 59 potential candidate genes for association with human longevity, J. Gerontol. A Biol. Sci. Med. Sci(2018) (Epub ahead of print Dec 30, 2017). [PMC free article] [PubMed] [Google Scholar]
  156. Houtkooper RH, Pirinen E, Auwerx J, Sirtuins as regulators of metabolism and healthspan, Nat. Rev. Mol. Cell Biol13 (2012) 225–238. [PMC free article] [PubMed] [Google Scholar]
  157. Flachsbart F, Croucher PJ, Nikolaus S, Hampe J, Cordes C, Schreiber S, Nebel A, Sirtuin 1 (SIRT1) sequence variation is not associated with exceptional human longevity, Exp. Gerontol41 (2006) 98–102. [PubMed] [Google Scholar]
  158. Razi S, Cogger VC, Kennerson M, Benson VL, McMahon AC, Blyth FM, Handelsman DJ, Seibel MJ, Hirani V, Naganathan V, Waite L, de Cabo RG, Cumming RG, Le Couteur DG, SIRT1 polymorphisms and serum-induced SIRT1 protein expression in aging and frailty: the CHAMP study, J. Gerontol. A Biol. Sci. Med. Sci72 (2017) 870–876. [PMC free article] [PubMed] [Google Scholar]
  159. Lin R, Yan D, Zhang Y, Liao X, Gong G, Hu J, Fu Y, Cai W, Common variants in SIRT1and human longevity in a Chinese population, BMC Med. Genet 17 (2016) 31. [PMC free article] [PubMed] [Google Scholar]
  160. Crocco P, Montesanto A, Passarino G, Rose G, Polymorphisms falling within putative miRNA target sites in the 3′UTR region of SIRT2and DRD2 genes are correlated with human longevity, J. Gerontol. A Biol. Sci. Med. Sci 71 (2016) 586–592. [PubMed] [Google Scholar]
  161. Rose G, Dato S, Altomare K, Bellizzi D, Garasto S, Greco V, Passarino G, Feraco E, Mari V, Barbi C, BonaFe M, Franceschi C, Tan Q, Boiko S, Yashin AI, De Benedictis G, Variability of the SIRT3gene, human silent information regulator Sir2 homologue, and survivorship in the elderly, Exp. Gerontol 38 (2003) 1065–1070. [PubMed] [Google Scholar]
  162. Albani D, Ateri E, Mazzuco S, Ghilardi A, Rodilossi S, Biella G, Ongaro F, Antuono P, Boldrini P, Di Giorgi E, Frigato A, Durante E, Caberlotto L, Zanardo A, Siculi M, Gallucci M, Forloni G, Modulation of human longevity by SIRT3single nucleotide polymorphisms in the prospective study “Treviso Longeva (TRELONG)”, Age (Dordr.) 36 (2014) 469–478. [PMC free article] [PubMed] [Google Scholar]
  163. Hirvonen K, Laivuori H, Lahti J, Strandberg T, Eriksson JG, Hackman P, SIRT6 polymorphism rs117385980 is associated with longevity and healthy aging in Finnish men, BMC Med. Genet18 (2017) 41. [PMC free article] [PubMed] [Google Scholar]
  164. Li Y, Qin J, Wei X, Liang G, Shi L, Jiang M, Xia T, Liang X, He M, Zhang Z, Association of SIRT6gene polymorphisms with human longevity, Iran. J. Public Health 45 (2016) 1420–1426. [PMC free article] [PubMed] [Google Scholar]
  165. Sahin E, Depinho RA, Linking functional decline of telomeres, mitochondria and stem cells during ageing, Nature464 (2010) 520–528. [PMC free article] [PubMed] [Google Scholar]
  166. Sahin E, Colla S, Liesa M, Moslehi J, Muller FL, Guo M, Cooper M, Kotton D, Fabian AJ, Walkey C, Maser RS, Tonon G, Foerster F, Xiong R, Wang YA, Shukla SA, Jaskelioff M, Martin ES, Heffeman TP, Protopopov A, Ivanova E, Mahoney JE, Kost-Alimova M, Perry SR, Bronson R, Liao R, Mulligan R, Shirihai OS, Chin L, DePinho RA, Telomere dysfunction induces metabolic and mitochondrial compromise, Nature470 (2011) 359–365. [PMC free article] [PubMed] [Google Scholar]
  167. Steenstrup T, Kark JD, Verhulst S, Thinggaard M, Hjelmborg JV, Dalgard C, Kyvik KO, Christiansen L, Mangino M, Spector TD, Petersen I, Kimura M, Benetos A, Labat C, Sinnreich R, Hwang SJ, Levy D, Hunt SC, Fitzpatrick AL, Chen W, Berenson GS, Barbieri M, Paolisso G, Gadalla SM, Savage SA, Christensen K, Yashin AI, Arbeev KG, Aviv A, Telomeres and the natural lifespan limit in humans, Aging9 (2017) 1130–1142. [PMC free article] [PubMed] [Google Scholar]
  168. Atzmon G, Cho M, Cawthon RM, Budagov T, Katz M, Yang X, Siegel G, Bergman A, Huffman DM, Schechter CB, Wright WE, Shay JW, Barzilai N, Govindaraju DR, Suh Y, Evolution in health and medicine Sackler colloquium: genetic variation in human telomerase is associated with telomere length in Ashkenazi centenarians, Proc. Natl. Acad. Sci. U. S. A107 (Suppl. 1) (2010) 1710–1717. [PMC free article] [PubMed] [Google Scholar]
  169. Soerensen M, Thinggaard M, Nygaard M, Dato S, Tan Q, Hjelmborg J, Andersen-Ranberg K, Stevnsner T, Bohr VA, Kimura M, Aviv A, Christensen K, Christiansen L, Genetic variation in TERTand TERC and human leukocyte telomere length and longevity: a cross-sectional and longitudinal analysis, Aging Cell 11 (2012) 223–227. [PMC free article] [PubMed] [Google Scholar]
  170. Estep III PW, Warner JB, Bulyk ML, Short-term calorie restriction in male mice feminizes gene expression and alters key regulators of conserved aging regulatory pathways, PLoS One4 (2009) e5242 (5211pages). [PMC free article] [PubMed] [Google Scholar]
  171. Donlon TA, Morris BJ, He Q, Chen R, Masaki KH, Allsopp RC, Willcox DC, Tranah GJ, Parimi N, Evans DS, Flachsbart F, Nebel A, Kim DH, Park J, Willcox BJ, Association of polymorphisms in connective tissue growth factor and epidermal growth factor receptor genes with human longevity, J. Gerontol. A Biol. Sci. Med. Sci72 (2017) 1038–1044. [PMC free article] [PubMed] [Google Scholar]
  172. Park JW, Ji YI, Choi YH, Kang MY, Jung E, Cho SY, Cho HY, Kang BK, Joung YS, Kim DH, Park SC, Park J, Candidate gene polymorphisms for diabetes mellitus, cardiovascular disease and cancer are associated with longevity in Koreans, Exp. Mol. Med41 (2009) 772–781. [PMC free article] [PubMed] [Google Scholar]
  173. Benigni A, Corna D, Zoja C, Sonzogni A, Latini R, Salio M, Conti S, Rottoli D, Longaretti L, Cassis P, Morigi M, Coffman TM, Remuzzi G, Disruption of the Ang II type 1 receptor promotes longevity in mice, J. Clin. Invest119 (2009) 524–530. [PMC free article] [PubMed] [Google Scholar]
  174. Benigni A, Orisio S, Noris M, Iatropoulos P, Castaldi D, Kamide K, Rakugi H, Arai Y, Todeschini M, Ogliari G, Imai E, Gondo Y, Hirose N, Mari D, Remuzzi G, Variations of the angiotensin II type 1 receptor gene are associated with extreme human longevity, Age (Dordr.)35 (2013) 993–1005. [PMC free article] [PubMed] [Google Scholar]
  175. Cambien F, Poirier O, Lecerf L, Evans A, Cambou JP, Arveiler D, Luc G, Bard JM, Bara L, Ricard S, et al., Deletion polymorphism in the gene for angiotensin-converting enzyme is a potent risk factor for myocardial infarction, Nature359 (1992) 641–644. [PubMed] [Google Scholar]
  176. Morris BJ, Zee RY, Schrader AP, Different frequencies of angiotensin-converting enzyme genotypes in older hypertensive individuals, J. Clin. Invest94 (1994) 1085–1089. [PMC free article] [PubMed] [Google Scholar]
  177. Galinsky D, Tysoe C, Brayne CE, Easton DF, Huppert FA, Dening TR, Paykel ES, Rubinsztein DC, Analysis of the apo E/apo C-I, angiotensin converting enzyme and methylenetetrahydrofolate reductase genes as candidates affecting human longevity, Atherosclerosis129 (1997) 177–183. [PubMed] [Google Scholar]
  178. Seripa D, Franceschi M, Matera MG, Panza F, Kehoe PG, Gravina C, Orsitto G, Solfrizzi V, Di Minno G, Dallapiccola B, Pilotto A, Sex differences in the association of apolipoprotein E and angiotensin-converting enzyme gene polymorphisms with healthy aging and longevity: a population-based study from southern Italy, J. Gerontol. A Biol. Sci. Med. Sci61 (2006) 918–923. [PubMed] [Google Scholar]
  179. da Silva AP, Marques NR, Matos A, Gil A, Clara JG, Bicho M, What angiotensin converting enzyme genotype and high blood pressure have in common in Portugene centenarians?J. Hypertens. 34 (2016) e80–e81. [Google Scholar]
  180. Wufuer M, Fang MW, Cheng ZH, Qiu CC, Polymorphism of angiotensin converting enzyme gene and natural longevity in the Xinjiang Uygur people: an association study, Zhonghua Yi Xue Za Zhi84 (2004) 1603–1606. [PubMed] [Google Scholar]
  181. Rahmutula D, Nakayama T, Izumi Y, Ozawa Y, Shimabukuro H, Kawamura H, Zhen-Wang S, Xiong-Wang J, Aisa M, Run-Yang C, Mahmut M, Mahsut R, Hen-Chen Z, Angiotensin-converting enzyme gene and longevity in the Xin Jiang Uighur autonomous region of China: an association study, J. Gerontol. A Biol. Sci. Med. Sci57 (2002) M57–M60. [PubMed] [Google Scholar]
  182. Garatachea N, Marin PJ, Lucia A, The ACEDD genotype and D-allele are associated with exceptional longevity: a meta-analysis, Ageing Res. Rev 12 (2013) 1079–1087. [PubMed] [Google Scholar]
  183. Chou PS, Wu SJ, Kao YH, Chou MC, Tai SY, Yang YH, Angiotensin-converting enzyme insertion/deletion polymorphism is associated with cerebral white matter changes in Alzheimer's disease, Geriatr Gerontol Int17 (2017) 945–950. [PubMed] [Google Scholar]
  184. Nygaard M, Thinggaard M, Christensen K, Christiansen L, Investigation of the 5q33.3 longevity locus and age-related phenotypes, Aging9 (2017) 247–255. [PMC free article] [PubMed] [Google Scholar]
  185. Vecchione C, Villa F, Carrizzo A, Spinelli CC, Damato A, Ambrosio M, Ferrario A, Madonna M, Uccellatore A, Lupini S, Maciag A, Ryskalin L, Milanesi L, Frati G, Sciarretta S, Bellazzi R, Genovese S, Ceriello A, Auricchio A, Malovini A, Puca AA, A rare genetic variant of BPIFB4predisposes to high blood pressure via impairment of nitric oxide signaling, Sci. Rep 7 (2017) 9706. [PMC free article] [PubMed] [Google Scholar]
  186. Villa F, Carrizzo A, Spinelli CC, Ferrario A, Malovini A, Maciag A, Damato A, Auricchio A, Spinetti G, Sangalli E, Dang Z, Madonna M, Ambrosio M, Sitia L, Bigini P, Cali G, Schreiber S, Peris T, Fucile S, Mulas F, Nebel A, Bellazzi R, Madeddu P, Vecchione C, Puca AA, Genetic analysis reveals a longevity-associated protein modulating endothelial function and angiogenesis, Circ. Res117 (2015) 333–345. [PMC free article] [PubMed] [Google Scholar]
  187. Villa F, Malovini A, Carrizzo A, Spinelli CC, Ferrario A, Maciag A, Madonna M, Bellazzi R, Milanesi L, Vecchione C, Puca AA, Serum BPIFB4 levels classify health status in long-living individuals, Immun. Ageing12 (2015) 27. [PMC free article] [PubMed] [Google Scholar]
  188. Spinetti G, Sangalli E, Specchia C, Villa F, Spinelli C, Pipolo R, Carrizzo A, Greco S, Voellenkle C, Vecchione C, Madeddu P, Martelli F, Puca AA, The expression of the BPIFB4 and CXCR4 associates with sustained health in long-living individuals from Cilento-Italy, Aging9 (2017) 370–380. [PMC free article] [PubMed] [Google Scholar]
  189. Hao Q, Wang Y, Ding X, Dong B, Yang M, Dong B, Wei Y, G-395A polymorphism in the promoter region of the KLOTHO gene associates with frailty among the oldest-old, Sci. Rep8 (2018) 6735. [PMC free article] [PubMed] [Google Scholar]
  190. Dato S, De Rango F, Crocco P, Passarino G, Rose G, Antioxidants and quality of aging: further evidences for a major role of TXNRD1gene variability on physical performance at old age, Oxidative Med. Cell. Longev 2015 (2015) 926067. [PMC free article] [PubMed] [Google Scholar]
  191. Niemi AK, Moilanen JS, Tanaka M, Hervonen A, Hurme M, Lehtimaki T, Arai Y, Hirose N, Majamaa K, A combination of three common inherited mitochondrial DNA polymorphisms promotes longevity in Finnish and Japanese subjects, Eur. J. Hum. Genet13 (2005) 166–170. [PubMed] [Google Scholar]
  192. Li L, Zheng HX, Liu Z, Qin Z, Chen F, Qian D, Xu J, Jin L, Wang X, Mitochondrial genomes and exceptional longevity in a Chinese population: the Rugao longevity study, Age (Dordr.)37 (2015) 9750. [PMC free article] [PubMed] [Google Scholar]
  193. De Benedictis G, Rose G, Carrieri G, De Luca M, Falcone E, Passarino G, Bonafe M, Monti D, Baggio G, Bertolini S, Mari D, Mattace R, Franceschi C, Mitochondrial DNA inherited variants are associated with successful aging and longevity in humans, FASEB J. 13(1999) 1532–1536. [PubMed] [Google Scholar]
  194. Ivanova R, Astrinidis A, Lepage V, Kouvatsi A, Djoulah S, Hors J, Charron D, Mitochondrial DNA polymorphism in the French population, Biomed Pharmacother53 (1999) 207–212. [PubMed] [Google Scholar]
  195. Ross OA, McCormack R, Curran MD, Duguid RA, Barnett YA, Rea IM, Middleton D, Mitochondrial DNA polymorphism: its role in longevity of the Irish population, Exp. Gerontol36 (2001) 1161–1178. [PubMed] [Google Scholar]
  196. He YH, Chen XQ, Yan DJ, Xiao FH, Lin R, Liao XP, Liu YW, Pu SY, Yu Q, Sim P, Jiang JJ, Cai WW, Kong QP, Familial longevity study reveals a significant association of mitochondrial DNA copy number between centenarians and their offspring, Neurobiol. Aging47 (2016) (218.e211–218.e218). [PubMed] [Google Scholar]
  197. Soerensen M, Dato S, Tan Q, Thinggaard M, Kleindorp R, Beekman M, Suchiman E, Jacobsen R, McGue M, Stevnsner T, Bohr VA, de Craen AJ, Westendorp RG, Schreiber S, Slagboom PE, Nebel A, Vaupel JW, Christensen K, Christiansen L, Evidence from case-control and longitudinal studies supports associations of genetic variation in APOECETP, and IL6with human longevity, Age (Dordr.) 35 (2013) 487–500. [PMC free article] [PubMed] [Google Scholar]
  198. Barzilai N, Atzmon G, Schechter C, Schaefer EJ, Cupples AL, Lipton R, Cheng S, Shuldiner AR, Unique lipoprotein phenotype and genotype associated with exceptional longevity, JAMA290 (2003) 2030–2040. [PubMed] [Google Scholar]
  199. Pan SL, Wang F, Lu ZP, Liu CW, Hu CY, Luo H, Peng JH, Luo XQ, Pang GF, Lu SH, Wu HY, Huang LJ, Yin RX, Cholesteryl ester transfer protein TaqIB polymorphism and its association with serum lipid levels and longevity in Chinese Bama Zhuang population, Lipids Health Dis. 11(2012) 26. [PMC free article] [PubMed] [Google Scholar]
  200. Kolovou G, Stamatelatou M, Anagnostopoulou K, Kostakou P, Kolovou V, Mihas C, Vasiliadis I, Diakoumakou O, Mikhailidis DP, Cokkinos DV, Cholesteryl ester transfer protein gene polymorphisms and longevity syndrome, Open Cardiovasc. Med. J4 (2010) 14–19. [PMC free article] [PubMed] [Google Scholar]
  201. Cellini E, Nacmias B, Olivieri F, Ortenzi L, Tedde A, Bagnoli S, Petruzzi C, Franceschi C, Sorbi S, Cholesteryl ester transfer protein (CETP) 1405V polymorphism and longevity in Italian centenarians, Mech. Ageing Dev126 (2005) 826–828. [PubMed] [Google Scholar]
  202. Rose G, Crocco P, De Rango F, Corsonello A, Lattanzio F, De Luca M, Passarino G, Metabolism and successful aging: polymorphic variation of syndecan-4 (SDC4) gene associate with longevity and lipid profile in healthy elderly Italian subjects, Mech. Ageing Dev150 (2015) 27–33. [PubMed] [Google Scholar]
  203. Okayama N, Hamanaka Y, Suehiro Y, Hasui Y, Nakamura J, Hinoda Y, Association of interleukin-10 promoter single nucleotide polymorphisms – 819 T/C and – 592 A/C with aging, J. Gerontol. A Biol. Sci. Med. Sci60 (2005) 1525–1529. [PubMed] [Google Scholar]
  204. Lio D, Scola L, Crivello A, Colonna-Romano G, Candore G, Bonafe M, Cavallone L, Franceschi C, Caruso C, Gender-specific association between – 1082 IL-10 promoter polymorphism and longevity, Genes Immun. 3(2002) 30–33. [PubMed] [Google Scholar]
  205. Khabour OF, Bamawi JM, Association of longevity with IL-10 – 1082 G/A and TNF-alpha-308 G/A polymorphisms, Int. I. Immunogenet37 (2010) 293–298. [PubMed] [Google Scholar]
  206. Naumova E, Mihaylova A, Ivanova M, Michailova S, Penkova K, Baltadjieva D, Immunological markers contributing to successful aging in Bulgarians, Exp. Gerontol39 (2004) 637–644. [PubMed] [Google Scholar]
  207. Yang F, Sun L, Zhu X, Han J, Zeng Y, Nie C, Yuan H, Li X, Shi X, Yang Y, Hu C, Lv Z, Huang Z, Zheng C, Liang S, Huang J, Wan G, Qi K, Qin B, Cao S, Zhao X, Zhang Y, Yang Z, Identification of new genetic variants of HLA-DQB1 associated with human longevity and lipid homeostasis-a cross-sectional study in a Chinese population, Aging9 (2017) 2316–2333. [PMC free article] [PubMed] [Google Scholar]
  208. Harris SE, Hagenaars SP, Davies G, David Hill W, Liewald DCM, Ritchie SJ, Marioni RE, Sudlow CLM, Wardlaw JM, McIntosh AM, Gale CR, Deary IJ, Molecular genetic contributions to self-rated health, Int. J. Epidemiol46 (2017) 994–1009. [PMC free article] [PubMed] [Google Scholar]
  209. Conneely KN, Capell BC, Erdos MR, Sebastiani P, Solovieff N, Swift AJ, Baldwin T, Budagov T, Barzilai N, Atzmon G, Puca AA, Peris TT, Geesaman J, Boehnke M, Collins FS, Human longevity and common variations in the LMNAgene: a meta-analysis, Aging Cell 11 (2012) 475–481. [PMC free article] [PubMed] [Google Scholar]
  210. Lunetta KL, D'Agostino RB Sr., Karasik D, Benjamin EJ, Guo CY, Govindaraju R, Kiel DP, Kelly-Hayes M, Massaro JM, Pencina MJ, Seshadri S, Murabito JM, Genetic correlates of longevity and selected age-related pheno-types: a genome-wide association study in the Framingham Study, BMC Med. Genet8 (2007) S13. [PMC free article] [PubMed] [Google Scholar]
  211. Bekpen C, Xie C, Nebel A, Tautz D, Involvement of SPATA31 copy number variable genes in human lifespan, Aging10 (2018) 674–688. [PMC free article] [PubMed] [Google Scholar]
  212. Han MV, Demuth JP, McGrath CL, Casola C, Hahn MW, Adaptive evolution of young gene duplicates in mammals, Genome Res. 19(2009) 859–867. [PMC free article] [PubMed] [Google Scholar]
  213. D'Aquila P, Crocco P, De Rango F, Indiveri C, Bellizzi D, Rose G, Passarino G, A genetic variant of ASCT2 hampers in vitro RNA splicing and correlates with human longevity, Rejuvenation Res. 21(2017) 193–199. [PubMed] [Google Scholar]
  214. Crocco P, Hoxha E, Dato S, De Rango F, Montesanto A, Rose G, Passarino G, Physical decline and survival in the elderly are affected by the genetic variability of amino acid transporter genes, Aging10 (2018) 658–673. [PMC free article] [PubMed] [Google Scholar]
  215. Khan SS, Shah SJ, Klyachko E, Baldridge AS, Eren M, Place AT, Aviv A, Puterman E, Lloyd-Jones DM, Heiman M, Miyata T, Gupta S, Shapiro AD, Vaughan E, A null mutation in SERPINE1protects against biological aging in humans, Sci. Adv 3 (2017) (eaaol617). [PMC free article] [PubMed] [Google Scholar]
  216. Fuku N, Diaz-Pena R, Arai Y, Abe Y, Zempo H, Naito H, Murakami H, Miyachi M, Spuch C, Serra-Rexach JA, Emanuele E, Hirose N, Lucia A, Epistasis, physical capacity-related genes and exceptional longevity: FNDC5gene interactions with candidate genes FOXOA3 [sic!] and APOEBMC Genomics 18 (2017) 803. [PMC free article] [PubMed] [Google Scholar]
  217. Gussago C, Arosio B, Guerini FR, Ferri E, Costa AS, Casati M, Bollini EM, Ronchetti F, Colombo E, Bemardelli G, Clerici M, Mari D, Impact of vitamin D receptor polymorphisms in centenarians, Endocrine53 (2016) 558–564. [PubMed] [Google Scholar]
  218. Dato S, Soerensen M, De Rango F, Rose G, Christensen K, Christiansen L, Passarino, The genetic component of human longevity: new insights from the analysis of pathway-based SNP-SNP interactions, Aging Cell17 (2018) el2755. [PMC free article] [PubMed] [Google Scholar]
  219. Fernandes M, Wan C, Tacutu R, Barardo D, Rajput A, Wang J, Thoppil H, Thornton D, Yang C, Freitas A, de Magalhaes JP, Systematic analysis of the gerontome reveals links between aging and age-related diseases, Hum. Mol. Genet25 (2016) 4804–4818. [PMC free article] [PubMed] [Google Scholar]
  220. Zhang Q, Nogales-Cadenas R, Lin JR, Zhang W, Cai Y, Vijg J, Zhang ZD, Systems-level analysis of human aging genes shed new light on mechanisms of aging, Hum. Mol. Genet25 (2016) 2934–2947. [PMC free article] [PubMed] [Google Scholar]
  221. Erikson GA, Bodian DL, Rueda M, Molparia B, Scott ER, Scott-Van Zeeland AA, Topol SE, Wineinger NE, Niederhuber JE, Topol EJ, Torkamani A, Whole-genome sequencing of a healthy aging cohort, Cell165 (2016) 1002–1011. [PMC free article] [PubMed] [Google Scholar]
  222. Pilling LC, Atkins JL, Duff MO, Beaumont RN, Jones SE, Tyrrell J, Kuo CL, Ruth KS, Tuke MA, Yaghootkar H, Wood AR, Murray A, Weedon MN, Harries W, Kuchel GA, Ferrucci L, Frayling TM, Melzer D, Red blood cell distribution width: genetic evidence for aging pathways in 116,666 volunteers, PLoS One12 (2017) e0185083. [PMC free article] [PubMed] [Google Scholar]
  223. Borras C, Abdelaziz KM, Gambini J, Serna E, Ingles M, de la Fuente M, Garcia I, Matheu A, Sanchis P, Belenguer A, Errigo A, Avellana JA, Barettino A, Lloret-Femandez C, Flames N, Pes G, Rodriguez-Manas L, Vina J, Human exceptional longevity: transcriptome from centenarians is distinct from septuagenarians and reveals a role of Bcl-xL in successful aging, Aging8 (2016) 3185–3208. [PMC free article] [PubMed] [Google Scholar]
  224. Kharbanda S, Saxena S, Yoshida K, Pandey P, Kaneki M, Wang Q, Cheng K, Chen YN, Campbell A, Sudha T, Yuan ZM, Narula J, Weichselbaum R, Nalin C, Kufe D, Translocation of SAPK/JNK to mitochondria and interaction with Bcl-x(L) in response to DNA damage, J. Biol. Chem275 (2000) 322–327. [PubMed] [Google Scholar]
  225. Shimizu S, Eguchi Y, Kamiike W, Waguri S, Uchiyama Y, Matsuda H, Tsujimoto Y, Bcl-2 blocks loss of mitochondrial membrane potential while ICE inhibitors act at a different step during inhibition of death induced by respiratory chain inhibitors, Oncogene13 (1996) 21–29. [PubMed] [Google Scholar]
  226. Opferman JT, Korsmeyer SJ, Apoptosis in the development and maintenance of the immune system, Nat. Immunol4 (2003) 410–415. [PubMed] [Google Scholar]
  227. Hoeijmakers JH, DNA damage, aging, and cancer, N. Engl. J. Med361 (2009) 1475–1485. [PubMed] [Google Scholar]
  228. Peters MJ, Joehanes R, Pilling LC, Schurmann C, Conneely KN, Powell J, Reinmaa E, Sutphin GL, Zhemakova A, Schramm K, Wilson YA, Kobes S, Tukiainen T, Ramos YF, Goring HH, Fomage M, Liu Y, Gharib SA, Stranger E, De Jager PL, Aviv A, Levy D, Murabito JM, Munson PJ, Huan T, Hofman A, Uitterlinden AG, Rivadeneira F, van Rooij J, Stolk L, Broer L, Verbiest MM, Jhamai M, Arp P, Metspalu A, Tserel L, Milani L, Samani J, Peterson P, Kasela S, Codd V, Peters A, Ward-Caviness CK, Herder C, Waldenberger M, Roden M, Singmann P, Zeilinger S, Illig T, Homuth G, Grabe HJ, Volzke H, Steil L, Kocher T, Murray A, Melzer D, Yaghootkar H, Bandinelli S, Moses EK, Kent JW, Curran JE, Johnson MP, Williams-Blangero S, Westra HJ, McRae AF, Smith JA, Kardia SL, Hovatta I, Perola M, Ripatti S, Salomaa V, Henders AK, Martin NG, Smith AK, Mehta D, Binder B, Nylocks KM, Kennedy EM, Klengel T, Ding J, Suchy-Dicey AM, Enquobahrie A, Brody J, Rotter JL, Chen YD, Houwing-Duistermaat J, Kloppenburg M, Slagboom PE, Helmer Q, den Hollander W, Bean S, Raj T, Bakhshi N, Wang QP, Oyston LJ, Psaty BM, Tracy RP, Montgomery GW, Turner ST, Blangero J, Meulenbelt I, Ressler KJ, Yang J, Franke L, Kettunen J, Visscher PM, Neely GG, Korstanje R, Hanson RL, Prokisch H, Ferrucci L, Esko T, Teumer A, van Meurs JB, Johnson AD, The transcriptional landscape of age in human peripheral blood, Nat. Commun6 (2015) 8570. [PMC free article] [PubMed] [Google Scholar]
  229. Sood S, Gallagher IJ, Lunnon K, Rullman E, Keohane A, Crossland H, Phillips E, Cederholm T, Jensen T, van Loon LJ, Lannfelt L, Kraus WE, Atherton PJ, Howard R, Gustafsson T, Hodges A, Timmons JA, A novel multi-tissue RNA diagnostic of healthy ageing relates to cognitive health status, Genome Biol. 16(2015) 185. [PMC free article] [PubMed] [Google Scholar]
  230. Crossland H, Atherton PJ, Stromberg A, Gustafsson T, Timmons JA, A reverse genetics cell-based evaluation of genes linked to healthy human tissue age, FASEB J. 31(2017) 96–108. [PMC free article] [PubMed] [Google Scholar]
  231. Dupont C, Armant DR, Brenner CA, Epigenetics: definition, mechanisms and clinical perspective, Semin. Reprod. Med27 (2009) 351–357. [PMC free article] [PubMed] [Google Scholar]
  232. Yin Y, Morgunova E, Jolma A, Kaasinen E, Sahu B, Khund-Sayeed S, Das PK, Kivioja T, Dave K, Zhong F, Nitta KR, Taipale M, Popov A, Ginno PA, Domcke S, Yan J, Schubeler D, Vinson C, Taipale J, Impact of cytosine methylation on DNA binding specificities of human transcription factors, Science356 (2017). [PMC free article] [PubMed] [Google Scholar]
  233. Kouzarides T, Chromatin modifications and their function, Cell128 (2007) 693–705. [PubMed] [Google Scholar]
  234. Chen Z, Li S, Subramaniam S, Shyy JY, Chien S, Epigenetic regulation: a new frontier for biomedical engineers, Annu. Rev. Biomed. Eng19 (2017) 195–219. [PubMed] [Google Scholar]
  235. Strahl BD, Allis CD, The language of covalent histone modifications, Nature403 (2000) 41–45. [PubMed] [Google Scholar]
  236. de Ruijter AJ, van Gennip AH, Caron HN, Kemp S, van Kuilenburg AB, Histone deacetylases (HDACs): characterization of the classical HDAC family, Biochem. J370 (2003) 737–749. [PMC free article] [PubMed] [Google Scholar]
  237. Gong F, Chiu LY, Miller KM, Acetylation reader proteins: linking acetylation signaling to genome maintenance and cancer, PLoS Genet. 12(2016) el006272. [PMC free article] [PubMed] [Google Scholar]
  238. Li Y, Daniel M, Tollefsbol TO, Epigenetic regulation of caloric restriction in aging, BMC Med. 9(2011) 98. [PMC free article] [PubMed] [Google Scholar]
  239. Luger K, Mader AW, Richmond RK, Sargent DF, Richmond TJ, Crystal structure of the nucleosome core particle at 2.8 A resolution, Nature389 (1997) 251–260. [PubMed] [Google Scholar]
  240. Kadonaga JT, Eukaryotic transcription: an interlaced network of transcription factors and chromatin-modifying machines, Cell92 (1998) 307–313. [PubMed] [Google Scholar]
  241. Horvath S, Raj K, DNA methylation-based biomarkers and the epigenetic clock theory of ageing, Nat. Rev. Genet19 (2018) 371–384. [PubMed] [Google Scholar]
  242. Heyn H, Li N, Ferreira HJ, Moran S, Pisano DG, Gomez A, Diez J, Sanchez-Mut JV, Setien F, Carmona FJ, Puca AA, Sayols S, Pujana MA, Serra-Musach J, Iglesias-Platas I, Formiga F, Fernandez AF, Fraga MF, Heath SC, Valencia A, Gut IG, Wang J, Esteller M, Distinct DNA methylomes of newborns and centenarians, Proc. Natl. Acad. Sci. U. S. A109 (2012) 10522–10527. [PMC free article] [PubMed] [Google Scholar]
  243. Gentilini D, Mari D, Castaldi D, Remondini D, Ogliari G, Ostan R, Bucci L, Sirchia SM, Tabano S, Cavagnini F, Monti D, Franceschi C, Di Blasio AM, Vitale G, Role of epigenetics in human aging and longevity: genome-wide DNA methylation profile in centenarians and centenarians' offspring, Age (Dordr.)35 (2013) 1961–1973. [PMC free article] [PubMed] [Google Scholar]
  244. Bocklandt S, Lin W, Sehl ME, Sanchez FJ, Sinsheimer JS, Horvath S, Vilain, Epigenetic predictor of age, PLoS One6 (2011) el4821. [Google Scholar]
  245. Horvath S, DNA methylation age of human tissues and cell types, Genome Biol. 14(2013) R115. [PMC free article] [PubMed] [Google Scholar]
  246. Tasselli L, Chua KF, Methylation gets into rhythm with NAD +-SIRT1, Nat. Struct. Mol. Biol 22 (2015) 275–277. [PubMed] [Google Scholar]
  247. Kilic U, Gok O, Erenberk U, Dundaroz MR, Torun E, Kucukardali Y, Elibol-Can B, Uysal O, Dundar T, A remarkable age-related increase in SIRT1protein expression against oxidative stress in elderly: SIRT1 gene variants and longevity in human, PLoS One 10 (2015) eOl17954. [PMC free article] [PubMed] [Google Scholar]
  248. Fraga MF, Ballestar E, Paz MF, Ropero S, Setien F, Ballestar ML, Heine-Suner D, Cigudosa JC, Urioste M, Benitez J, Boix-Chomet M, Sanchez-Aguilera A, Ling C, Carlsson E, Poulsen P, Vaag A, Stephan Z, Spector TD, Wu YZ, Plass C, Esteller M, Epigenetic differences arise during the lifetime of monozygotic twins, Proc. Natl. Acad. Sci. U. S. A102 (2005) 10604–10609. [PMC free article] [PubMed] [Google Scholar]
  249. Cheung P, Vallania F, Warsinske HC, Donato M, Schaffert S, Chang SE, Dvorak M, Dekker CL, Davis MM, Utz PJ, Khatri P, Kuo AJ, Single-cell chromatin modification profiling reveals increased epigenetic variations with aging, Cell173 (2018) 1385–1397. [PMC free article] [PubMed] [Google Scholar]
  250. Horvath S, Pirazzini C, Bacalini MG, Gentilini D, Di Blasio AM, Delledonne M, Mari D, Arosio B, Monti D, Passarino G, De Rango F, D'Aquila P, Giuliani C, Marasco E, Collino S, Descombes P, Garagnani P, Franceschi C, Decreased epigenetic age of PBMCs from Italian semi-supercentenarians and their offspring, Aging7 (2015) 1159–1170. [PMC free article] [PubMed] [Google Scholar]
  251. Valinluck V, Sowers LC, Endogenous cytosine damage products alter the site selectivity of human DNA maintenance methyltransferase DNMT1, Cancer Res. 67(2007) 946–950. [PubMed] [Google Scholar]
  252. Schuyler RP, Merkel A, Raineri E, Altucci L, Vellenga E, Martens JHA, Pourfarzad F, Kuijpers TW, Burden F, Farrow S, Downes K, Ouwehand WH, Clarke L, Datta A, Lowy E, Flicek P, Frontini M, Stunnenberg HG, Martin-Subero JI, Gut I, Heath S, Distinct trends of DNA methylation patterning in the innate and adaptive immune systems, Cell Rep. 17(2016) 2101–2111. [PMC free article] [PubMed] [Google Scholar]
  253. Nakano K, Whitaker JW, Boyle DL, Wang W, Firestein GS, DNA methylome signature in rheumatoid arthritis, Ann. Rheum. Dis72 (2013) 110–117. [PMC free article] [PubMed] [Google Scholar]
  254. Xiao FH, He YH, Li QG, Wu H, Luo LH, Kong QP, A genome-wide scan reveals important roles of DNA methylation in human longevity by regulating age-related disease genes, PLoS One10 (2015) e0120388. [PMC free article] [PubMed] [Google Scholar]
  255. Greer EL, Shi Y, Histone methylation: a dynamic mark in health, disease and inheritance, Nat. Rev. Genet13 (2012) 343–357. [PMC free article] [PubMed] [Google Scholar]
  256. Booth LN, Brunet A, The aging epigenome, Mol. Cell62 (2016) 728–744. [PMC free article] [PubMed] [Google Scholar]
  257. Cole JJ, Robertson NA, Rather MI, Thomson JP, McBryan T, Sproul D, Wang T, Brock C, Clark W, Ideker T, Meehan RR, Miller RA, Brown-Borg HM, Adams PD, Diverse interventions that extend mouse lifespan suppress shared age-associated epigenetic changes at critical gene regulatory regions, Genome Biol. 18(2017) 58. [PMC free article] [PubMed] [Google Scholar]
  258. Siebold AP, Banerjee R, Tie F, Kiss DL, Moskowitz J, Harte PJ, Polycomb repressive complex 2 and Trithorax modulate Drosophila longevity and stress resistance, Proc. Natl. Acad. Sci. U. S. A107 (2010) 169–174. [PMC free article] [PubMed] [Google Scholar]
  259. Maures TJ, Greer EL, Hauswirth AG, Brunet A, The H3K27 demethylase UTX-1 regulates C. eleganslifespan in a germline-independent, insulin-dependent manner, Aging Cell 10 (2011) 980–990. [PMC free article] [PubMed] [Google Scholar]
  260. Dang W, Steffen KK, Perry R, Dorsey JA, Johnson FB, Shilatifard A, Kaeberlein M, Kennedy BK, Berger SL, Histone H4 lysine 16 acetylation regulates cellular lifespan, Nature459 (2009) 802–807. [PMC free article] [PubMed] [Google Scholar]
  261. Ni Z, Ebata A, Alipanahiramandi E, Lee SS, Two SET domain containing genes link epigenetic changes and aging in Caenorhabditis elegansAging Cell11 (2012) 315–325. [PMC free article] [PubMed] [Google Scholar]
  262. Ivanov A, Pawlikowski J, Manoharan I, van Tuyn J, Nelson DM, Rai TS, Shah PP, Hewitt G, Korolchuk VI, Passos JF, Wu H, Berger SL, Adams PD, Lysosome-mediated processing of chromatin in senescence, J. Cell Biol202 (2013) 129–143. [PMC free article] [PubMed] [Google Scholar]
  263. Kreiling JA, Tamamori-Adachi M, Sexton AN, Jeyapalan JC, Munoz-Najar U, Peterson AL, Manivannan J, Rogers ES, Pchelintsev NA, Adams PD, Sedivy JM, Age-associated increase in heterochromatic marks in murine and primate tissues, Aging Cell10 (2011) 292–304. [PMC free article] [PubMed] [Google Scholar]
  264. Pasque V, Halley-Stott RP, Gillich A, Garrett N, Gurdon JB, Epigenetic stability of repressed states involving the histone variant macroH2A revealed by nuclear transfer to Xenopus oocytes, Nucleus (Austin, Tex.)2 (2011) 533–539. [PMC free article] [PubMed] [Google Scholar]
  265. Barger JL, Anderson RM, Newton MA, da Silva C, Vann JA, Pugh TD, Someya S, Prolla TA, Weindruch R, A conserved transcriptional signature of delayed aging and reduced disease vulnerability is partially mediated by SIRT3, PLoS One10 (2015) e0120738. [PMC free article] [PubMed] [Google Scholar]
  266. Hu Z, Chen K, Xia Z, Chavez M, Pal S, Seol JH, Chen CC, Li W, Tyler JK, Nucleosome loss leads to global transcriptional up-regulation and genomic in-stability during yeast aging, Genes Dev. 28(2014) 396–408. [PMC free article] [PubMed] [Google Scholar]
  267. Raddatz G, Hagemann S, Aran D, Sohle J, Kulkami PP, Kaderali L, Heilman A, Winnefeld M, Lyko F, Aging is associated with highly defined epigenetic changes in the human epidermis, Epigenetics Chromatin6 (2013) 36. [PMC free article] [PubMed] [Google Scholar]
  268. Liu L, Cheung TH, Charville GW, Hurgo BM, Leavitt T, Shih J, Brunet A, Rando TA, Chromatin modifications as determinants of muscle stem cell quiescence and chronological aging, Cell Rep. 4(2013) 189–204. [PMC free article] [PubMed] [Google Scholar]
  269. Pifferi F, Terrien J, Marchal J, Dal-Pan A, Djelti F, Hardy I, Ch'ahory S, Cordonnier N, Desquilbet L, Hurion M, Zahariev A, Chery I, Zizzari P, Perret M, Epelbaum J, Blanc S, Picq J-L, Dhenain M, Aujard FE, I.L., Caloric restriction increases lifespan but affects brain integrity in grey mouse lemurs, Comm. Biol1 (2018) 30. [PMC free article] [PubMed] [Google Scholar]
  270. Maegawa S, Lu Y, Tahara T, Lee JT, Madzo J, Liang S, Jelinek J, Colman RJ, Issa JJ, Caloric restriction delays age-related methylation drift, Nat. Commun8 (2017) 539. [PMC free article] [PubMed] [Google Scholar]
  271. Li Y, Liu L, Tollefsbol TO, Glucose restriction can extend normal cell lifespan and impair precancerous cell growth through epigenetic control of hTERTand p16 expression, FASEB J. 24 (2010) 1442–1453. [PMC free article] [PubMed] [Google Scholar]
  272. Wang C, Wang F, Li Z, Cao Q, Huang L, Chen S, MeCP2-mediated epigenetic regulation in senescent endothelial progenitor cells, Stem Cell Res Ther9 (2018) 87. [PMC free article] [PubMed] [Google Scholar]
  273. Leibiger IB, Berggren PO, Sirtl: a metabolic master switch that modulates lifespan, Nat. Med12 (2006) 34–36 (discussion 36). [PubMed] [Google Scholar]
  274. Cohen HY, Miller C, Bitterman KJ, Wall NR, Hekking B, Kessler B, Howitz T, Gorospe M, de Cabo R, Sinclair DA, Calorie restriction promotes mammalian cell survival by inducing the SIRT1 deacetylase, Science305 (2004) 390–392. [PubMed] [Google Scholar]
  275. Fischle W, Wang Y, Allis CD, Histone and chromatin cross-talk, Curr. Opin. Cell Biol15 (2003) 172–183. [PubMed] [Google Scholar]
  276. Kouzarides T, Histone methylation in transcriptional control, Curr. Opin. Genet. Dev12 (2002) 198–209. [PubMed] [Google Scholar]
  277. Bird A, DNA methylation patterns and epigenetic memory, Genes Dev. 16(2002) 6–21. [PubMed] [Google Scholar]
  278. Bird AP, CpG-rich islands and the function of DNA methylation, Nature321 (1986) 209–213. [PubMed] [Google Scholar]
  279. Bell JT, Tsai PC, Yang TP, Pidsley R, Nisbet J, Glass D, Mangino M, Zhai G, Zhang F, Valdes A, Shin SY, Dempster EL, Murray RM, Grundberg E, Hedman AK, Nica A, Small KS, Mu TC, Dermitzakis ET, McCarthy MI, Mill J, Spector TD, Deloukas P, Epigenome-wide scans identify differentially methylated regions for age and age-related phenotypes in a healthy ageing population, PLoS Genet. 8(2012) el002629. [PMC free article] [PubMed] [Google Scholar]
  280. Horvath S, Zhang Y, Langfelder P, Kahn RS, Boks MP, van Eijk K, van den Berg LH, Ophoff RA, Aging effects on DNA methylation modules in human brain and blood tissue, Genome Biol. 13(2012) R97. [PMC free article] [PubMed] [Google Scholar]
  281. Hannum G, Guinney J, Zhao L, Zhang L, Hughes G, Sadda S, Klotzle B, Bibikova M, Fan JB, Gao Y, Deconde R, Chen M, Rajapakse I, Friend S, Ideker T, Zhang K, Genome-wide methylation profiles reveal quantitative views of human aging rates, Mol. Cell49 (2013) 359–367. [PMC free article] [PubMed] [Google Scholar]
  282. Garagnani P, Bacalini MG, Pirazzini C, Gori D, Giuliani C, Mari D, Di Blasio AM, Gentilini D, Vitale G, Collino S, Rezzi S, Castellani G, Capri M, Salvioli S, Franceschi C, Methylation of ELOVL2gene as a new epigenetic marker of age, Aging Cell 11 (2012) 1132–1134. [PubMed] [Google Scholar]
  283. Kim S, Myers L, Wyckoff J, Cherry KE, Jazwinski SM, The frailty index out-performs DNA methylation age and its derivatives as an indicator of biological age, GeroScience39 (2017) 83–92. [PMC free article] [PubMed] [Google Scholar]
  284. Marioni RE, Shah S, McRae AF, Chen BH, Colicino E, Harris SE, Gibson J, Henders AK, Redmond P, Cox SR, Pattie A, Corley J, Murphy L, Martin NG, Montgomery W, Feinberg AP, Fallin MD, Multhaup ML, Jaffe AE, Joehanes R, Schwartz J, Just AC, Lunetta KL, Murabito JM, Starr JM, Horvath S, Baccarelli AA, Levy D, Visscher PM, Wray NR, Deary IJ, DNA methylation age of blood predicts all-cause mortality in later life, Genome Biol. 16(2015) 25. [PMC free article] [PubMed] [Google Scholar]
  285. Marioni RE, Shah S, McRae AF, Ritchie SJ, Muniz-Terrera G, Harris SE, Gibson J, Redmond P, Cox SR, Pattie A, Corley J, Taylor A, Murphy L, Starr JM, Horvath S, Visscher PM, Wray NR, Deary IJ, The epigenetic clock is correlated with physical and cognitive fitness in the Lothian Birth Cohort 1936, Int. J. Epidemiol44 (2015) 1388–1396. [PMC free article] [PubMed] [Google Scholar]
  286. Issa JP, Ahuja N, Toyota M, Bronner MP, Brentnall TA, Accelerated age-related CpG island methylation in ulcerative colitis, Cancer Res. 61(2001) 3573–3577. [PubMed] [Google Scholar]
  287. Jylhava J, Kananen L, Raitanen J, Marttila S, Nevalainen T, Hervonen A, Jylha M, Hurme M, Methylomic predictors demonstrate the role of NF-kB in old-age mortality and are unrelated to the aging-associated epigenetic drift, Oncotarget7 (2016) 19228–19241. [PMC free article] [PubMed] [Google Scholar]
  288. Bollati V, Schwartz J, Wright R, Litonjua A, Tarantini L, Suh H, Sparrow D, Vokonas P, Baccarelli A, Decline in genomic DNA methylation through aging in a cohort of elderly subjects, Mech. Ageing Dev130 (2009) 234–239. [PMC free article] [PubMed] [Google Scholar]
  289. Shimoda N, Izawa T, Yoshizawa A, Yokoi H, Kikuchi Y, Hashimoto N, Decrease in cytosine methylation at CpG island shores and increase in DNA fragmentation during zebrafish aging, Age (Dordr.)36 (2014) 103–115. [PMC free article] [PubMed] [Google Scholar]
  290. Lardenoije R, Iatrou A, Kenis G, Kompotis K, Steinbusch HW, Mastroeni D, Coleman P, Lemere CA, Hof PR, van den Hove DL, Rutten BP, The epigenetics of aging and neurodegeneration, Prog. Neurobiol131 (2015) 21–64. [PMC free article] [PubMed] [Google Scholar]
  291. Moore AZ, Hernandez DG, Tanaka T, Pilling LC, Nalls MA, Bandinelli S, Singleton AB, Ferrucci L, Change in epigenome-wide DNA methylation over 9 years and subsequent mortality: results from the InCHIANTI study, J. Gerontol. A Biol. Sci. Med. Sci71 (2016) 1029–1035. [PMC free article] [PubMed] [Google Scholar]
  292. Xu X, Zhan M, Duan W, Prabhu V, Brenneman R, Wood W, Firman J, Li H, Zhang P, Ibe C, Zonderman AB, Longo DL, Poosala S, Becker KG, Mattson MP, Gene expression atlas of the mouse central nervous system: impact and interactions of age, energy intake and gender, Genome Biol. 8(2007) R234. [PMC free article] [PubMed] [Google Scholar]
  293. Numata S, Ye T, Hyde TM, Guitart-Navarro X, Tao R, Wininger M, Colantuoni C, Weinberger DR, Kleinman JE, Lipska BK, DNA methylation signatures in development and aging of the human prefrontal cortex, Am. J. Hum. Genet90 (2012) 260–272. [PMC free article] [PubMed] [Google Scholar]
  294. Siegmund KD, Connor CM, Campan M, Long TI, Weisenberger DJ, Biniszkiewicz D, Jaenisch R, Laird PW, Akbarian S, DNA methylation in the human cerebral cortex is dynamically regulated throughout the life span and involves differentiated neurons, PLoS One2 (2007) e895. [PMC free article] [PubMed] [Google Scholar]
  295. Rakyan VK, Down TA, Maslau S, Andrew T, Yang TP, Beyan H, Whittaker P, McCann T, Finer S, Valdes AM, Leslie RD, Deloukas P, Spector TD, Human aging-associated DNA hypermethylation occurs preferentially at bivalent chromatin domains, Genome Res. 20(2010) 434–439. [PMC free article] [PubMed] [Google Scholar]
  296. Mo A, Mukamel EA, Davis FP, Luo C, Henry GL, Picard S, Urich MA, Nery JR, Sejnowski TJ, Lister R, Eddy SR, Ecker JR, Nathans J, Epigenomic signatures of neuronal diversity in the mammalian brain, Neuron86 (2015) 1369–1384. [PMC free article] [PubMed] [Google Scholar]
  297. Casillas MA Jr., Lopatina N, Andrews LG, Tollefsbol TO, Transcriptional control of the DNA methyltransferases is altered in aging and neoplastically-transformed human fibroblasts, Mol. Cell. Biochem252 (2003) 33–43. [PubMed] [Google Scholar]
  298. Mastroeni D, Grover A, Delvaux E, Whiteside C, Coleman PD, Rogers J, Epigenetic changes in Alzheimer's disease: decrements in DNA methylation, Neurobiol. Aging31 (2010) 2025–2037. [PMC free article] [PubMed] [Google Scholar]
  299. Symmank J, Zimmer G, Regulation of neuronal survival by DNA methyltransferases, Neural Regen. Res12 (2017) 1768–1775. [PMC free article] [PubMed] [Google Scholar]
  300. Monk D, Germline-derived DNA methylation and early embryo epigenetic reprogramming: the selected survival of imprints, Int. J. Biochem. Cell Biol67 (2015) 128–138. [PubMed] [Google Scholar]
  301. Fraser R, Lin CJ, Epigenetic reprogramming of the zygote in mice and men: on your marks, get set, go!, Reproduction152 (2016) R211–R222. [PMC free article] [PubMed] [Google Scholar]
  302. Vaiserman AM, Koliada AK, Jirtle RL, Non-genomic transmission of longevity between generations: potential mechanisms and evidence across species, Epigenetics Chromatin10 (2017) 38. [PMC free article] [PubMed] [Google Scholar]
  303. Armstrong NJ, Mather KA, Thalamuthu A, Wright MJ, Trollor JN, Ames D, Brodaty H, Schofield PR, Sachdev PS Kwok JB, Aging, exceptional longevity and comparisons of the Hannum and Horvath epigenetic clocks, Epigenomics9 (2017) 689–700. [PubMed] [Google Scholar]
  304. McEwen LM, Morin AM, Edgar RD, Maclsaac JL, Jones MJ, Dow WH, Rosero-Bixby L, Kobor MS, Rehkopf DH, Differential DNA methylation and lymphocyte proportions in a costa Rican high longevity region, Epigenetics Chromatin10 (2017) 21. [PMC free article] [PubMed] [Google Scholar]
  305. Seim I, Ma S, Gladyshev VN, Gene expression signatures of human cell and tissue longevity, NPJ Aging and Mechanisms of Disease2 (2016) 16014. [PMC free article] [PubMed] [Google Scholar]
  306. Benayoun BA, Pollina EA, Brunet A, Epigenetic regulation of ageing: linking environmental inputs to genomic stability, Nat. Rev. Mol. Cell Biol16 (2015) 593–610. [PMC free article] [PubMed] [Google Scholar]
  307. Burke B, Stewart CL, The nuclear lamins: flexibility in function, Nat. Rev. Mol. Cell Biol14 (2013) 13–24. [PubMed] [Google Scholar]
  308. Mattout-Drubezki A, Gruenbaum Y, Dynamic interactions of nuclear lamina proteins with chromatin and transcriptional machinery, Cell. Mol. Life Sci60 (2003) 2053–2063. [PubMed] [Google Scholar]
  309. Lund E, Oldenburg AR, Delbarre E, Freberg CT, Duband-Goulet I, Eskeland R, Buendia P Collas, Lamin A/C-promoter interactions specify chromatin state-dependent transcription outcomes, Genome Res. 23(2013) 1580–1589. [PMC free article] [PubMed] [Google Scholar]
  310. Peric-Hupkes D, Meuleman W, Pagie L, Bruggeman SW, Solovei I, Brugman W, Graf S, Flicek P, Kerkhoven RM, van Lohuizen M, Reinders M, Wessels B van Steensel, Molecular maps of the reorganization of genome-nuclear lamina interactions during differentiation, Mol. Cell38 (2010) 603–613. [PMC free article] [PubMed] [Google Scholar]
  311. de Leeuw R, Gruenbaum Y, Medalia O, Nuclear lamins: thin filaments with major functions, Trends Cell Biol. 28(2018) 34–45. [PubMed] [Google Scholar]
  312. Tsurumi A, Li WX, Global heterochromatin loss: a unifying theory of aging?Epigenetics 7 (2012) 680–688. [PMC free article] [PubMed] [Google Scholar]
  313. Chandra T, Kirschner K, Thuret JY, Pope BD, Ryba T, Newman S, Ahmed K, Samarajiwa SA, Salama R, Carroll T, Stark R, Janky R, Narita M, Xue L, Chicas A, Nunez S, Janknecht R, Hayashi-Takanaka Y, Wilson MD, Marshall A, Odom DT, Babu MM, Bazett-Jones DP, Tavare S, Edwards PA, Lowe SW, Kimura H, Gilbert DM, Narita M, Independence of repressive histone marks and chromatin compaction during senescent heterochromatic layer formation, Mol. Cell47 (2012) 203–214. [PMC free article] [PubMed] [Google Scholar]
  314. Eriksson M, Brown WT, Gordon LB, Glynn MW, Singer J, Scott L, Erdos MR, Robbins M, Moses TY, Berglund P, Dutra A, Pak E, Durkin S, Csoka AB, Boehnke M, Glover TW, Collins FS, Recurrent de novopoint mutations in lamin A cause Hutchinson-Gilford progeria syndrome, Nature 423 (2003) 293–298. [PubMed] [Google Scholar]
  315. Scaffidi P, Misteli T, Lamin A-dependent nuclear defects in human aging, Science312 (2006) 1059–1063. [PMC free article] [PubMed] [Google Scholar]
  316. Chandra T, Ewels PA, Schoenfelder S, Furlan-Magaril M, Wingett SW, Kirschner K, Thuret JY, Andrews S, Fraser P, Reik W, Global reorganization of the nuclear landscape in senescent cells, Cell Rep. 10(2015) 471–483. [PMC free article] [PubMed] [Google Scholar]
  317. Thiery JP, Macaya G, Bemardi G, An analysis of eukaryotic genomes by density gradient centrifugation, J. Mol. Biol108 (1976) 219–235. [PubMed] [Google Scholar]
  318. Encode PC, The ENCODE project consortium. An integrated encyclopedia of DNA elements in the human genome, Nature489 (2012) 57–74. [PMC free article] [PubMed] [Google Scholar]
  319. Mattick JS, Rinn JL, Discovery and annotation of long noncoding RNAs, Nat. Struct. Mol. Biol22 (2015) 5–7. [PubMed] [Google Scholar]
  320. Kour S, Rath PC, Long noncoding RNAs in aging and age-related diseases, Ageing Res. Rev26 (2016) 1–21. [PubMed] [Google Scholar]
  321. Boehm M, Slack F, A developmental timing microRNA and its target regulate life span in C. elegansScience310 (2005) 1954–1957. [PubMed] [Google Scholar]
  322. de Lencastre A, Pincus Z, Zhou K, Kato M, Lee SS, Slack FJ, MicroRNAs both promote and antagonize longevity in C. elegansCurr. Biol20 (2010) 2159–2168. [PMC free article] [PubMed] [Google Scholar]
  323. Gombar S, Jung HJ, Dong F, Calder B, Atzmon G, Barzilai N, Tian XL, Pothof J, Hoeijmakers JH, Campisi J, Vijg J, Suh Y, Comprehensive microRNA profiling in B-cells of human centenarians by massively parallel sequencing, BMC Genomics13 (2012) 353. [PMC free article] [PubMed] [Google Scholar]
  324. Noren Hooten N, Abdelmohsen K, Gorospe M, Ejiogu N, Zonderman AB, Evans MK, microRNA expression patterns reveal differential expression of target genes with age, PLoS One5 (2010) el0724. [PMC free article] [PubMed] [Google Scholar]
  325. Serna E, Gambini J, Borras C, Abdelaziz KM, Belenguer A, Sanchis P, Avellana JA, Rodriguez-Manas L, Vina J, Centenarians, but not octogenarians, up-regulate the expression of microRNAs, Sci. Rep2 (2012) 961. [PMC free article] [PubMed] [Google Scholar]
  326. Smith-Vikos T, Liu Z, Parsons C, Gorospe M, Ferrucci L, Gill TM, Slack FJ, A serum miRNA profile of human longevity: findings from the Baltimore Longitudinal Study of Aging (BLSA), Aging8 (2016) 2971–2987. [PMC free article] [PubMed] [Google Scholar]
  327. Batista PJ, Chang HY, Long noncoding RNAs: cellular address codes in development and disease, Cell152 (2013) 1298–1307. [PMC free article] [PubMed] [Google Scholar]
  328. Johnsson P, Lipovich L, Grander D, Morris KV, Evolutionary conservation of long non-coding RNAs; sequence, structure, function, Biochim. Biophys. Acta1840 (2014) 1063–1071. [PMC free article] [PubMed] [Google Scholar]
  329. Quek XC, Thomson DW, Maag JL, Bartonicek N, Signal B, Clark MB, Gloss BS, Dinger ME, IncRNAdb v2.0: expanding the reference database for functional long noncoding RNAs, Nucleic Acids Res. 43(2015) D168–D173. [PMC free article] [PubMed] [Google Scholar]
  330. Wang Y, Xu Z, Jiang J, Xu C, Kang J, Xiao L, Wu M, Xiong J, Guo X, Liu H, Endogenous miRNA sponge lincRNA-RoR regulates Oct4, Nanog, and Sox2 in human embryonic stem cell self-renewal, Dev. Cell25 (2013) 69–80. [PubMed] [Google Scholar]
  331. Ponting CP, Oliver PL, Reik W, Evolution and functions of long noncoding RNAs, Cell136 (2009) 629–641. [PubMed] [Google Scholar]
  332. St Laurent G, Wahlestedt C, Kapranov P, The Landscape of long noncoding RNA classification, Trends Genet. 31(2015) 239–251. [PMC free article] [PubMed] [Google Scholar]
  333. Montes M, Lund AH, Emerging roles of IncRNAs in senescence, FEBS J. 283(2016) 2414–2426. [PubMed] [Google Scholar]
  334. Abdelmohsen K, Panda A, Kang MJ, Xu J, Selimyan R, Yoon JH, Martindale JL, De S, Wood WH 3rd, Becker KG, Gorospe M, Senescence-associated IncRNAs: senescence-associated long noncoding RNAs, Aging Cell12 (2013) 890–900. [PMC free article] [PubMed] [Google Scholar]
  335. Azzalin CM, Reichenbach P, Khoriauli L, Giulotto E, Lingner J, Telomeric repeat containing RNA and RNA surveillance factors at mammalian chromosome ends, Science318 (2007) 798–801. [PubMed] [Google Scholar]
  336. Martadinata H, Heddi B, Lim KW, Phan AT, Structure of long human telomeric RNA (TERRA): G-quadruplexes formed by four and eight UUAGGG repeats are stable building blocks, Biochemistry50 (2011) 6455–6461. [PubMed] [Google Scholar]
  337. Redon S, Reichenbach P, Lingner J, The non-coding RNA TERRA is a natural ligand and direct inhibitor of human telomerase, Nucleic Acids Res. 38(2010) 5797–5806. [PMC free article] [PubMed] [Google Scholar]
  338. di Iulio J, Bartha I, Wong EHM, Yu HC, Lavrenko V, Yang D, Jung I, Hicks MA, Shah N, Kirkness EF, Fabani MM, Biggs WH, Ren B, Venter JC, Telenti A, The human noncoding genome defined by genetic diversity, Nat. Genet. 50(2018) 333–337. [PubMed] [Google Scholar]
  339. Fulop T, Dupuis G, Baehl S, Le Page A, Bourgade K, Frost E, Witkowski JM, Pawelec G, Larbi A, Cunnane S, From inflamm-aging to immune-paralysis: a slippery slope during aging for immune-adaptation, Biogerontology17 (2016) 147–157. [PubMed] [Google Scholar]
  340. Munoz-Culla M, Irizar H, Gorostidi A, Alberro A, Osorio-Querejeta I, Ruiz-Martinez J, Olascoaga J, Lopez De Munain A, Otaegui D, Progressive changes in non-coding RNA profile in leucocytes with age, Aging9 (2017) 1202–1218. [PMC free article] [PubMed] [Google Scholar]
  341. Maes OC, An J, Sarojini H, Wang E, Murine microRNAs implicated in liver functions and aging process, Mech. Ageing Dev129 (2008) 534–541. [PubMed] [Google Scholar]
  342. Mimura S, Iwama H, Kato K, Nomura K, Kobayashi M, Yoneyama H, Miyoshi H, Tani J, Morishita A, Himoto T, Deguchi A, Nomura T, Sakamoto T, Fujita K, Maeda E, Izuishi K, Okano K, Suzuki Y, Masaki T, Profile of microRNAs associated with aging in rat liver, Int. J. Mol. Med34 (2014) 1065–1072. [PubMed] [Google Scholar]
  343. Green CD, Huang Y, Dou X, Yang L, Liu Y, Han JJ, Impact of dietary interventions on noncoding RNA networks and mRNAs encoding chromatin-related factors, Cell Rep. 18(2017) 2957–2968. [PubMed] [Google Scholar]
  344. Blankenburg H, Pramstaller PP, Domingues FS, A network-based meta-analysis for characterizing the genetic landscape of human aging, Biogerontology19 (2018) 81–94. [PMC free article] [PubMed] [Google Scholar]
  345. Meeran SM, Ahmed A, Tollefsbol TO, Epigenetic targets of bioactive dietary components for cancer prevention and therapy, Clin. Epigenetics1 (2010) 101–116. [PMC free article] [PubMed] [Google Scholar]
  346. Howitz KT, Bitterman KJ, Cohen HY, Lamming DW, Lavu S, Wood JG, Zipkin RE, Chung P, Kisielewski A, Zhang LL, Scherer B, Sinclair DA, Small molecule activators of sirtuins extend Saccharomyces cerevisiaelifespan, Nature 425 (2003) 191–196. [PubMed] [Google Scholar]
  347. Wood JG, Rogina B, Lavu S, Howitz K, Helfand SL, Tatar M, Sinclair D, Sirtuin activators mimic caloric restriction and delay ageing in metazoans, Nature430 (2004) 686–689. [PubMed] [Google Scholar]
  348. Bass TM, Weinkove D, Houthoofd K, Gems D, Partridge L, Effects of resveratrol on lifespan in Drosophila melanogasterand Caenorhabditis elegansMech. Ageing Dev 128 (2007) 546–552. [PubMed] [Google Scholar]
  349. Bam JA, Pearson KJ, Price NL, Jamieson HA, Lerin C, Kalra A, Prabhu VV, Allard JS, Lopez-Lluch G, Lewis K, Pistell PJ, Poosala S, Becker KG, Boss O, Gwinn D, Wang M, Ramaswamy S, Fishbein KW, Spencer RG, Lakatta EG, Le Couteur D, Shaw RJ, Navas P, Puigserver P, Ingram DK, de Cabo R, Sinclair DA, Resveratrol improves health and survival of mice on a high-calorie diet, Nature444 (2006) 337–342. [PMC free article] [PubMed] [Google Scholar]
  350. Barger JL, Kayo T, Vann JM, Arias EB, Wang J, Hacker TA, Wang Y, Raederstorff D, Morrow JD, Leeuwenburgh C, Allison DB, Saupe KW, Cartee GD, Weindruch R, Prolla TA, A low dose of dietary resveratrol partially mimics caloric restriction and retards aging parameters in mice, PLoS One3 (2008) e2264. [PMC free article] [PubMed] [Google Scholar]
  351. Agarwal B, Baur JA, Resveratrol and life extension, Ann. N. Y. Acad. Sci1215 (2011) 138–143. [PubMed] [Google Scholar]
  352. Fischer-Posovszky P, Kukulus V, Tews D, Unterkircher T, Debatin KM, Fulda S, Wabitsch M, Resveratrol regulates human adipocyte number and function in a Sirtl-dependent manner, Am. J. Clin. Nutr92 (2010) 5–15. [PubMed] [Google Scholar]
  353. Subramanian L, Youssef S, Bhattacharya S, Kenealey J, Polans AS, van Ginkel PR, Resveratrol: challenges in translation to the clinic-a critical discussion, Clin. Cancer Res16 (2010) 5942–5948. [PMC free article] [PubMed] [Google Scholar]
  354. Patel KR, Scott E, Brown VA, Gescher AJ, Steward WP, Brown K, Clinical trials of resveratrol, Ann. N. Y. Acad. Sci1215 (2011) 161–169. [PubMed] [Google Scholar]
  355. Li Y, Liu L, Andrews LG, Tollefsbol TO, Genistein depletes telomerase activity through cross-talk between genetic and epigenetic mechanisms, Int. J. Cancer125 (2009) 286–296. [PMC free article] [PubMed] [Google Scholar]
  356. Taylor CK, Levy RM, Elliott JC, Burnett BP, The effect of genistein aglycone on cancer and cancer risk: a review of in vitro, preclinical, and clinical studies, Nutr. Rev67 (2009) 398–415. [PubMed] [Google Scholar]
  357. Li Y, Yuan YY, Meeran SM, Tollefsbol TO, Synergistic epigenetic reactivation of estrogen receptor-alpha (ERalpha) by combined green tea polyphenol and histone deacetylase inhibitor in ERalpha-negative breast cancer cells, Mo Cancer9 (2010) 274. [PMC free article] [PubMed] [Google Scholar]
  358. Meeran SM, Patel SN, Tollefsbol TO, Sulforaphane causes epigenetic repression of hTERT expression in human breast cancer cell lines, PLoS One5 (2010) el1457. [PMC free article] [PubMed] [Google Scholar]
  359. Jayagopal V, Albertazzi P, Kilpatrick ES, Howarth EM, Jennings PE, Hepburn DA, Atkin SL, Beneficial effects of soy phytoestrogen intake in post-menopausal women with type 2 diabetes, Diabetes Care25 (2002) 1709–1714. [PubMed] [Google Scholar]
  360. Kao YH, Chang HH, Lee MJ, Chen CL, Tea, obesity, and diabetes, Mol. Nutr. Food Res50 (2006) 188–210. [PubMed] [Google Scholar]
  361. Shanafelt TD, Call TG, Zent CS, Laplant B, Bowen DA, Roos M, Secreto CR, Ghosh AK, Rabat BF, Lee MJ, Yang CS, Jelinek DF, Erlichman C, Kay NE, Phase I trial of daily oral Polyphenon E in patients with asymptomatic Rai stage 0 to II chronic lymphocytic leukemia, J. Clin. Oncol27 (2009) 3808–3814. [PMC free article] [PubMed] [Google Scholar]
  362. Athanasopoulos D, Karagiannis G, Tsolaki M, Recent findings in Alzheimer disease and nutrition focusing on epigenetics, Adv. Nutrition (Bethesda, MD)7 (2016) 917–927. [PMC free article] [PubMed] [Google Scholar]
  363. Block T, El-Osta A, Epigenetic programming, early life nutrition and the risk of metabolic disease, Atherosclerosis266 (2017) 31–40. [PubMed] [Google Scholar]
  364. Hodjat M, Rahmani S, Khan F, Niaz K, Navaei-Nigjeh M, Mohammadi Nejad S, Abdollahi M, Environmental toxicants, incidence of degenerative diseases, and therapies from the epigenetic point of view, Arch. Toxicol91 (2017) 2577–2597. [PubMed] [Google Scholar]
  365. Puca AA, Spinelli C, Accardi G, Villa F, Caruso C, Centenarians as a model to discover genetic and epigenetic signatures of healthy ageing, Mech. Ageing Dev174 (2017) 95–102. [PubMed] [Google Scholar]
  366. Stover PJ, James WPT, Krook A, Garza C, Emerging concepts on the role of epigenetics in the relationships between nutrition and health, J. Intern. Med284 (2018) 37–49. [PubMed] [Google Scholar]
  367. Mathers JC, Nutritional modulation of ageing: genomic and epigenetic approaches, Mech. Ageing Dev127 (2006) 584–589. [PubMed] [Google Scholar]
  368. Dhahbi JM, Mote PL, Fahy GM, Spindler SR, Identification of potential caloric restriction mimetics by microarray profiling, Physiol. Genomics23 (2005) 343–350. [PubMed] [Google Scholar]
  369. Tarrago MG, Chini CCS, Kanamori KS, Warner GM, Caride A, de Oliveira GC, Rud M, Samani A, Hein KZ, Huang R, Jurk D, Cho DS, Boslett JJ, Miller JD, Zweier JL, Passos JF, Doles JD, Becherer DJ, Chini EN, A potent and specific CD38 inhibitor ameliorates age-related metabolic dysfunction by reversing tissue NAD+decline, Cell Metab. 27 (2018) 1081–1095.el010. [PMC free article] [PubMed] [Google Scholar]
  370. Eline Slagboom P, van den Berg N, Deelen J, Phenome and genome based studies into human ageing and longevity: an overview, Biochim. Biophys. Acta1864 (Pt A) (2017) 2742–2751. [PubMed] [Google Scholar]
  371. Rizzuto D, Keller L, Orsini N, Graff C, Backman L, Bellocco R, Wang HX, Fratiglioni, Effect of the interplay between genetic and behavioral risks on survival after age 75, J. Am. Geriatr. Soc64 (2016) 2440–2447. [PubMed] [Google Scholar]
  372. Stone RC, Horvath K, Kark JD, Susser E, Tishkoff SA, Aviv A, Telomere length and the cancer-atherosclerosis trade-off, PLoS Genet. 12(2016) el006144. [PMC free article] [PubMed] [Google Scholar]
  373. Dato S, Rose G, Crocco P, Monti D, Garagnani P, Franceschi C, Passarino G, The genetics of human longevity: an intricacy of genes, environment, culture and microbiome, Mech. Ageing Dev165 ((2017) 147–155. [PubMed] [Google Scholar]

This is an open access article under the terms of the http://creativecommons.org/licenses/by/4.0/ License, which permits use, distribution and reproduction in any medium, provided the original work is properly cited.

Credit also to PUBMED Genetic and epigenetic regulation of human aging and longevity - PMC (nih.gov)

Related Articles